Sunteți pe pagina 1din 417

NEUROPSYCHOLOGICAL TREATMENT AFTER BRAIN INJURY

FOUNDATIONS OF NEUROPSYCHOLOGY
Barbara Uzzell, Series Editor
Ellis, D. W., Christensen, A.L., eds.: Neuropsychological Treatment After Brain II/jury

NEUROPSYCHOLOGICAL TREATMENT
AFTER BRAIN INJURY

Edited by
DAVID W. ELLIS
Mediplex Rehab-Camden, Camden, NJ

ANNE-LISE CHRISTENSEN
Center for Rehabilitation of Brain Damage,
University of Copenhagen, Denmark

"
~.

KLUWER ACADEMIC PUBLISHERS

BOSTONIDORDRECHT/LONDON

Distributors
for North America: Kluwer Academic Publishers, 101 Philip Drive, Assinippi Park, Norwell,
MA 02061, USA
for all other countries: Kluwer Academic Publishers, Distribution Centre, P. O. Box 322, 3300
AH Dordrecht, The Netherlands

Library of Congress Cataloging-in-Publication Data


Neuropsychological treatment after brain injury/edited by David W. Ellis and
Anne-Lise Christensen.
p. cm.-(Foundations of neuropsychology)
Includes bibliographies and index.
ISBN-13: 978-1-4612-8876-3
e-ISBN-13: 978-1-4613-1581-0
DOT: 10.1007/978-1-4613-1581-0
1. Brain damage-Patients-Rehabilitation. 2. Clinical neuropsychology.
David W., Ph. D. II. Christensen, Anne-Lise, Ph. D. III. Series.
[DNLM: 1. Brain Injuries-rehabilitation. 2. Brain
Injuries-therapy. 3. Neuropsychology.
WL 354 N498443]
RC387.5.N49 1989
617' .481044-dc19
DNLM/DLC
for Library of Congress

I. Ellis,

88-13753
CIP

Copyright 1989 by Kluwer Academic Publishers.


Softcover reprint of the hardcover 1st edition 1989
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted in any form or by any means, mechanical, photocopying, recording, or otherwise,
without the prior written permission of the publisher, Kluwer Academic Publishers, 101 Philip
Drive, Assinippi Park, Norwell, MA 02061, USA.

CONTENTS

Contributing Authors
Acknowledgements
1. Introduction

Vll

IX

DAVID W. ELLIS AND ANNE-LISE CHRISTENSEN

I. THEORY AND INTERVENTION

2. Brain Injury: A Pathophysiological Basis for Neuropsychological


Dysfunction

15

JENS ASTRUP

3. Issues in Behavioral Neurology and Brain Injury

39

DA VID F. LONG

4. Psychopharmacological Agents in the Treatment of Brain Injury

91

GREGORY J. O'SHANICK AND DEAN X. PARMELEE

5. Traumatic Brain Injury and the Rehabilitation Process:


A Psychiatric Perspective

105

IRWIN W. POLLACK

6. The Neuropsychological Investigation as a Therapeutic and


Rehabilitative Technique

127

ANNE-LISE CHRISTENSEN

vi Contents

II. REHABILITATION PROGRAMS: APPLICATION OF THEORY

7. Interventions in the Inpatient Setting

157

M. ELIZABETH SANDEL

8. Residential Treatment

183

MEREDITH M. SARGENT

9. Concepts in Day Programming

221

NATHANIEL H. MAYER AND DANIEL J. KEATING

III. NEUROPSYCHOLOGICAL REHABILITATION TECHNIQUES

10. Neuropsychotherapy

241

DAVID W. ELLIS

11. Structured Group Treatment for Brain-Injury Survivors

271

YEHUDA BEN-YISHAY AND PHYLLIS LAKIN

12. Long-Term Family Intervention

297

HARVEY E. JACOBS

13. Management of Aggressive Behavior Following Traumatic Brain Injury

317

WILLIAM J. HAFFEY AND JOHN W. SCIBAK

IV. PROFESSIONAL AND LEGAL ISSUES

14. Professional Issues in Neuropsychological Rehabilitation

363

LANCE E. TREXLER

15. Legal Issues that Commonly Confront Brain-Injury Survivors and


Their Families

379

SIMON H. FORGETTE

Index

409

CONTRIBUTING AUTHORS

Jens Astrup, MD, University of Arhus, Arhus Kommunehospital, 8000 Arhus


C, Denmark.
Yehuda Ben-Yishay, PhD, New York University Medical Center, Head
Trauma Program, 345 E. 24th St., New York, NY 10010.
Anne-Lise Christensen, PhD, Center for Rehabilitation of Brain Damage,
University of Copenhagen, Njalsgade 88, DK-2300 Copenhagen, Denmark.
David W. Ellis, PhD, Mediplex Rehab-Camden, 2 Cooper Plaza, Camden,
NJ 08103.
Simon H. Forgette, JD, 12624 N. E. 85th St., Kirkland, W A 98033.
William J. Haffey, PhD, Executive Director, Rehabilitation, Sharp HealthCare, 7901 Frost St., San Diego, CA 92123.
Harvey E. Jacobs, PhD, Department of Psychiatry and Biobehavioral Sciences,
Neuropsychiatric Institute, UCLA Medical Center, 760 Westwood Plaza,
Los Angeles, CA 90024.
Daniel J. Keating, PhD, Drucker Brain Injury Center, Moss Rehabilitation
Hospital, 12th St. & Tabor Rd., Philadelphia, PA 19141.
Phyllis Lakin, PhD, New York University Medical Center, Head Trauma
Program, 345 E. 24th St., New York, NY 10010.
David F. Long, MD, Bryn Mawr Rehabilitation Hospital, 414 Paoli Pike,
Malvern, P A 19354.
Nathaniel H. Mayer, MD, Drucker Brain Injury Center, Moss Rehabilitation
Center, 12th St. & Tabor Rd., Philadelphia, PA 19141.
Gregory J. O'Shanick, MD, Box 710, Medical College of Virginia, Richmond,
VA 23298.
vii

viii

Contributing Authors

Dean X. Parmelee, MD, Box 489, Medical College of Virginia, Richmond,


VA 23298.
Irwin W. Pollack, MA, MD, The Center for Cognitive Rehabilitation, 172
New St., New Brunswick, NJ 0890l.
M. Elizabeth Sandel, MD, Mediplex Rehab-Camden, 2 Cooper Plaza,
Camden, NJ 08103.
Meredith M. Sargent, PhD, 40 Amherst Ave., Swarthmore, PA 1908l.
John W. Scibak, PhD, New Medico Head Injury System, 78 Maplewood
Shops, Northampton, MA 01060.
Lance E. Trexler, PhD, Center for Neuropsychological Rehabilitation, 8925
North Meridian St., Suite 100, Indianapolis, IN 46260.

ACKNOWLEDGEMENTS

This book is the result of a collective effort towards the integration of current
knowledge about the theory and practice of neuropsychological treatment of
brain-injury survivors. To acknowledge individually all the researchers and
clinicians whose accumulated work has resulted in this text would be impossible, but we have made every effort to trace original material and to give
full credit for ideas and interpretations.
Special acknowledgement must be made to the late Professor Doctor K.
Bash. This book was discussed during a joint visit we made to Dr. Bash in
Zurich. He felt very strongly about the integration of personality and neuropsychological functioning.
Professor Doctor Erik Stromgren must be thanked for his forceful thinking
behind the Center for Rehabilitation of Brain Injury. In addition, the Egmont
Foundation is gratefully acknowledged for its original and continuous support
of the Center for Rehabilitation of Brain Damage.
The preparation of this manuscript would not have been possible without
the assistance of Elsa R. Efran, MS. As our Editorial Assistant, she worked
tirelessly to ensure the clarity and consistency of the material. Apolonia Galie
typed various portions of the manuscript and handled the logistics of the project. Robert Ouaou also provided valuable technical assistance.
The Neuropsychology Department of Mediplex Rehab-Camden, under the
direction of Kathy Lawler, PhD, and Mark Rader, PhD, provided support in
the reading of many chapters. M. Elizabeth Sandel, MD, acted as a medical
consultant for the editing of the chapters of a medical nature.
ix

Acknowledgements

We would like to thank our respective staffs at Mediplex Rehab-Camden


and the Center for Rehabilitation of Brain Damage for their support, encouragement, and collaborative spirit.
Finally, of course, we should acknowledge our patients and their families,
who have inspired us to attempt to develop more effective, more compassionate approaches for the treatment of brain injury.
David W. Ellis, Philadelphia
Anne-Lise Christensen, Copenhagen

NEUROPSYCHOLOGICAL TREATMENT AFTER BRAIN INJURY

1. INTRODUCTION

DAVID W. ELLIS AND ANNE-LISE CHRISTENSEN

1 A BRIEF OVERVIEW

In the past, most people who sustained catastrophic brain injury died. However,
over the past several decades, sophisticated medical diagnostic techniques such
as computerized tomography (CT) and magnetic resonance imaging (MRI),
along with advances in emergency trauma procedures and neurosurgical procedures (e. g., intracranial pressure monitoring), have dramatically increased
the survival rates for people who have survived such trauma. At the same
time, because of population growth, the number of victims of brain trauma
(primarily automobile accidents) has also risen [1].
As a result of their injuries, many of these people have developed severe
disabilities that affect their lives and the lives of everyone around them. For
those who survive, and their families, mere survival is not enough. Attention
must be paid to the quality of their lives after the traumatic event.
During the past 15 years, there has been an increasing focus on the development of treatment techniques for brain injuries. Although the principal focus
of this text is on the neuropsychological (i. e., neurological and psychological)
aspects of treatment-both theory and technique-the book is also directed
towards the broad variety of issues that affect survivors, their families, healthcare professionals, and the social milieu.
First, the nature of the trauma will be described, along with typical posttraumatic neurobehavioral patterns and syndromes. Next, a number of neurorehabilitation theories, programs, and techniques will be considered. Finally,
the professional and legal aspects of neurobehavioral recovery will be examined.
1

The intended audience includes all neurorehabilitation professionals, as well


as family members and others in the community who are directing their efforts
to the treatment and care of the brain-injury survivor.
The ways in which the brain may be injured are numerous: neurodevelopmental disorders (e.g., perinatal or postnatal complications), cerebral vascular
disorders, anoxia, open-brain injury, closed-brain injury, tumors, and so on.
Although this book will touch upon many of these issues, the primary focus
will be on the treatment of adults and older adolescents who have sustained a
brain injury of a traumatic nature. Neurodevelopmental assessment and treatment of children and younger adolescents are specialty areas that demand a
different but also highly complex neuropsychological knowledge base [2].
Traumatic brain injury (TBI) is defined as a blow to the head that results in
diminished abilities subsequent to the injury and that requires rehabilitative
intervention. TBI is primarily caused by motor vehicle accidents and violent
crimes [3]. If the victim survives the catastrophic injury, a typical pattern
of behavior problems may emerge, depending on the type of brain injury.
Although the pattern varies, depending upon the location and severity of the
injury, the pattern-in tandem with the person's premorbid personalitycreates the survivor's new personality matrix. Typical posttrauma personality
problems include difficulties with 1) attention and concentration, 2) hyper- or
hypo-arousability or emotionality, 3) language, 4) acceptance of disabilities, 5)
self-esteem or self-concept, 6) interpersonal relationships, 7) abstractions, and
8) intimacy [4].
Other difficulties may arise, depending on the patient's age when the injury
occurs. For example, catastrophic brain injury in early adolescence will create
neurodevelopmental problems for personality organization. Such a disruption usually results in educational, vocational, family, and interpersonal
complications.
This book is not directed towards the treatment of those TBI survivors who
are comatose or in a "vegetative state" or "persistent vegetative state." Rather,
the treatment approaches that many of the contributors describe can be used
with many brain-injury survivors who can benefit from an active rehabilitation program. This includes survivors with mild, moderate, and severe brain
Injury.
In this book, different approaches and methods for handling the neurobehavioral consequences of brain injury are presented. However, the book is
titled Neuropsychological Treatment After Brain Injury to emphasize the neurological and psychological difficulties that arise for the survivor and his or her
family, as well as the current trends for treatment of the wide range of difficulties that result from such an injury.
The primary focus is on describing the way in which each therapeutic approach is used to actually treat the survivor. Contributors have been asked to
provide some examples of difficulties that their patients have experienced and
to explore methods they have used to deal with those problems. Furthermore,

1. Introduction

what professionals mean by the broader theoretical concepts such as "personality" and "emotions," especially in the context of behavioral problems
encountered posttrauma, is discussed.
During the initial period of rehabilitation interventions in the late 1970s
and early 1980s, workers in the field of rehabilitation of TEl survivors were
enthusiastic about changing the neurological functioning through learning
techniques (e.g., cognitive remediation through computers). This enthusiasm
has been tempered by mixed results [5].
In the United States, the brain-injury rehabilitation movement has been
spurred forward by families of survivors as well as by organized institutions of
health care. Practitioners have created rehabilitation centers for brain-injury
survivors in response to the professional challenge and the need for services.
Rehabilitation programs in England, Denmark, and the rest of the world have
developed along somewhat similar lines. Nevertheless, the question remains:
What is the best way to help the survivor recover from a catastrophic brain
injury? In essence, what techniques are most effective?
Miller [6], reporting on the techniques available to the rehabilitation professional, stated that creativity and ingenuity were the skills most needed. He
felt that intervention techniques were still at such a primitive level that new
methods of intervention must be created and old methods rigorously evaluated.
At that time, Miller challenged professionals to create new ways in which to
enhance the recovery process after brain injury.
The issues in the rehabilitation of brain-injury survivors are numerous and
complex. For example, we all recognize that a clear understanding of the
pathology related to the injury, as well as the technical methods for assisting
recovery, is important in the rehabilitation process. However, the very definitions of terms involved in the therapeutic process have still not been agreed
upon. For instance, a number of professionals may state that they are practicing a certain type of intervention (e.g., cognitive remediation), yet may
mean very different things by such a term. Miller [6] has pointed out that even
such apparently innocuous terms as recovery need to be clearly defined and
agreed upon by rehabilitation professionals.
Over the years, the boundaries that separate professional disciplines within
the brain-injury field have been changing. Interdisciplinary efforts have become a reality. Currently, however, in many interdisciplinary teams, each
specialist is concerned about a different "piece" of the patient or client. One
professional looks at the range of motion in a patient's limb; another focuses
on the phonemic structure of a sound. Very few programs appear to be concerned about the "entire" survivor and his or her family. Another difficulty
with the interdisciplinary approach is that professionals have failed to reach an
agreement about what the clinical practice of rehabilitation for a brain-injury
survivor should consist of. One of our goals in writing this book was to give
rehabilitation professionals from many different fields a common perspective
from which to approach certain treatment techniques. We believe that the term

neuropsychological treatment does not delineate a single discipline. Rather, it


represents an entire body of knowledge and techniques concerned with the
neurological and psychological functioning of the survivor.
Most professionals agree on the use of an interdisciplinary team to treat a
person with brain injury and subsequent difficulties. They disagree on how it
should be done. The need in the field of rehabilitation appears to be for both
1) an overarching theory of rehabilitation or a number of theories of rehabilitation, and 2) empirical evidence of the success or failure of certain interventions
[6]. The theory or theories not only must make sense of the structural aspects
of brain damage, but also must remain consistent with applied techniques.
In 1980, approximately 10 facilities were treating brain-injury survivors. By
1987 [7], well over 500 such facilities were available. How are the professionals
who staff these facilities to be adequately trained? The professional must have
supervised treatment experiences that cut across disciplines (interdisciplinary)
in order to have the knowledge base needed to apply their models of neuropsychological, movement, and performance interventions. These professional
issues will also be addressed in this book.
2 THE EXTENT OF THE PROBLEM

At present, there are no accurate figures (either in the United State or abroad)
that show the actual incidence ofTBI or the nature or extent of the injuries [3].
Based on the percentage of brain-injury survivors today who are disabled to
various degrees, Kraus concluded that 73,724 to 98,325 people each year in the
United States are candidates for rehabilitation-approximately 200 in every
100,000 (i. e., 1 of every 500). In a Danish study, Engberg and Vinterberg [8]
have noted that the number ofTBI patients with persistent sequelae in Denmark
is not precisely known. The patient register for the country as a whole shows
that approximately 15,000 patients a year are sent home from the hospital with
the diagnosis of concussive syndrome; approximately 2,200 patients art sent
home with diagnoses of subarachnoid, extradural, and traumatic hemorrhage.
An additional 1,250 patients are sent home with the diagnosis of TBI.
Hovedcirklen (the Danish National Association of Support Groups for
Sudden Brain Damage) estimates that, in Denmark, approximately 17,000
people each year suffer a head injury to such a degree that hospitalization is
required (Hovedcirklen, private communication). By far the majority come
out of the insult with only a slight concussion. Some few hundred suffer for
months from headache and memory and concentration difficulties. Three
hundred to 500 people each year contract severe brain damage with extended
or irreversible consequences. Young men between the ages of 15 and 24 are
most vulnerable. Because those in the severely affected group are often young
and thus live with their injury for many years, we are talking about a group
of many thousands of people, even though the exact number is not precisely
known.

1. Introduction

In 1972, Field [9] reported that in England and Wales the incidence for brain
injury was 430 per 100,000. On the other hand, according to Jennett et al. [10],
in 1974 the British rate was approximately 270 per 100,000 (i.e., for England/
Wales) and 313 for Scotland. Selecki et al. [11] reported that for Australia and
New South Wales, the rate in 1977 was 377 per 100,000. What can account for
these discrepancies (both within countries and between countries) in the reported incidence ofbrain-injury cases? Kraus has argued that the higher figures
may be due to different criteria for reporting cases (including a broader list of
International Classification of Diseases codes), inaccurate case ascertainment,
inclusion of the same case several times, and confusion of what was diagnosed
as brain injury [3].
Obviously, the results of these studies suggest that a more accurate method
for collecting data is needed. However, if even Kraus's conservative estimates
are correct, the number of brain-injury survivors represents an enormous
health-care problem-as well as an economic burden-for ma'1Y societies.
3 REHABILITATION

What do rehabilitation and recovery offunction really mean? To understand these


terms, we need to look at how recovery after brain injury occurs. This is often
expressed as a "recovery curve," which varies from patient to patient. The
basic outline of the recovery curve has been tentatively assembled through
some initial studies [12, 13].
The age at which the person was injured and the severity of the injury appear
to be the most important factors that affect recovery. However, other factors
also appear to influence recovery, including 1) the time that has elapsed since
the injury; 2) the environment from which the TBI survivor has come, and to
which he or she has returned; 3) how the lesion was acquired (i. e., serial or
interactional); 4) the patient's psychological characteristics before the injury;
and 5) the effects of drugs or alcohol [14].
Cope and Hall [15] have shown that the relationship between time since
injury and aggressiveness of rehabilitation interrelate with outcome. Because
neuropsychological, functional behavioral, and research data must be used to
direct treatment programs, a major issue is how to integrate the data from all
available sources. How, for instance, does the professional understand and
treat the patient's deficits in memory, reaction time, attention/concentration,
abstract reasoning, or activities of daily living (e.g., problems of morning
care, making breakfast, work, lunch, afternoon and evening activities, and
preparing for bed)? In the final assessment, how do we provide rehabilitation
and treatment programs?
3.1 Methodological Problems

The methodological problems in the study of brain-injury rehabilitation are


comparable to those in any new clinical field. In America, researchers claim

that empirical evidence has been insufficient regarding either the behavioral
characteristics of brain-injury survivors or the effects of varied treatment interventions on such survivors.
Miller [6] has categorized models of recovery along three general theoretical
lines, which he designated as 1) artifact, 2) anatomical reorganization, and 3)
functional adaptation. The artifact theory refers to the result of the traumatic
insult-a suppression or reduction in the physiological or mechanical processes
(e.g., temporary reduction in brain function)-and claims that the processes
have not truly been lost. Therefore, the term recovery is not considered accurate
because the TBI survivor's ability to perform a particular behavior has not
been lost, but rather suppressed. The anatomical reorganization theory implies
that-after brain injury-other areas of the brain take over the functions of
those areas that have been damaged. Functional adaptation reflects the survivor's
ability to relearn, by means other than those originally employed, behaviors
that were lost through the damage.
Miller noted that the model of recovery outlined in the functional adaptation
theory appeared to be of greatest merit. This concept of recovery of function
is directed towards explicit goals, with the overall objective of bringing the
person to the highest level of adaptive functioning that is possible posttrauma.
In addition, he suggested that the goals and interventions that professionals
choose should be consistent with the current understanding of the characteristic
problems after brain injury [6].
3.2 Outcome Studies

Researchers generally agree about what information is critical for evaluating


outcome after brain injury and for developing a prognosis for the patient.
Recently, discoveries based on new outcome data have been altering some of
the accepted tenets and traditions of brain-injury rehabilitation. As a result of
this research, brain-injury professionals are currently adapting concepts from
the field of general rehabilitation, including retraining and remediation of
various skills, acceptance of disability, theories of recovery, and prevention of
morbidity, and are creating techniques specifically designed for brain-injury
survivors. These techniques now include 1) specialized individual treatment,
2) adapted group therapy, 3) cognitive remediation, 4) neuropsychological
assessment, 5) day treatment, 6) inpatient treatment, 7) behavioral neurological
interventions, 8) family interventions, 9) behavioral therapy, and 10) neuropsychiatric interventions. Many of these areas are addressed in this book.
Recognition of the long-term problems encountered by brain-injury survivors has begun to alter the role of rehabilitation in the United States and
Western Europe. The specific combination of deficits associated with closedhead injury has created new challenges for treatment and placement. The field
of closed-head injury rehabilitation appears to be transforming and evolving
towards an integrated base of information.

1. Introduction

3.3 Examples of Outcome Studies

In one of the few outcome studies that have looked at the effectiveness of treatment, Prigatano et al. [13] examined the effects of a neuropsychological rehabilitation program on 18 closed-head injured (CHI) patients (compared
with a control group of 17 untreated CHI subjects). Prigatano's team found
that the treated patients had an increase (although it was not statistically significant) in neuropsychological functioning, as well as a statistically significant
decrease in emotional distress. The treated patients were more productive and
showed fewer disturbances of personality than did patients in the untreated
group. In addition, the treated patients showed an increase in learning and
memory after their treatment.
Oddy and Humphrey [16] and Weddell, Oddy, and Jenkins [17] studied
patients' social adjustment after brain injury. Fifty brain-injured patients were
studied over a 12-month period [16], and 44 were studied over a 2-year time
frame [17]. This research led Oddy to conclude that vocational adjustment and
cognitive functioning appear to return at a greater rate than does the ability to
initiate and perform leisure and social activities [16, 17]. According to these
studies, personality changes and social isolation were the two most severe
problems impacting on family adjustment.
Ben-Yishay et al. [18] studied the vocational outcome of a sample of 94
patients who were followed for three years after their discharge from the New
York Program. At the point of discharge, 63% were competitively employed,
21 % were employed in a subsidized capacity, and 16% were not employed.
Followup data on 36 patients revealed that after three years only 50% of
patients were employed (22% at noncompetitive levels), and 28% were not
employed. The researchers found that the major reasons for the decrease in
employability were
(1) social isolation coupled with the absence of adequate maintenance and support
systems, (2) forgetting to apply the acquired "rehabilitation algorithms" and/or use
compensatory mnemonic aids, also due to the absence of adequate maintenance and
support systems, and (3) financial disincentives to work (p. 44).

They found that the predominant reasons for maintaining employment were
(1) improvement in self-awareness, discipline and regulation of emotional responses,
(2) increase in the effectiveness of functional application of the residual informationprocessing abilities (rather than an increment in the capacity levels per se), and (3) significant improvements in the acceptance by patients of their existential situation (p. 45).

The employability or non employability of a brain-injury survivor becomes a


complicated index of successful rehabilitation.
Outcome studies such as these appear to indicate that rehabilitation can have
substantial impact on the vocational adjustment, emotional and personality

alterations, and family adjustment of brain-injury survivors. However, there


is no agreement about which treatment variables have had the most impact.
Until recently, studies of treatment outcome have primarily focused on the
acute phase of the trauma and the initial assessment of patients. Based on such
factors, survivors have been categorized as having severe disability, moderate
disability, mild disability, or no disability. However, the Glasgow Outcome
Scale [19], which is used in the acute setting, is considered too general to be
used in postacute programming. The Disability Rating Scale [20] has also been
suggested as a viable instrument for reporting the recovery of disability after
brain injury. Nevertheless, predictability of outcome and evaluation of
treatment is still in a rather primitive phase.
4 NEUROPSYCHOLOGICAL TREATMENT

As Harry Stack Sullivan [21] stated about all human beings, "We are much
more alike than we are different. " This holds true for brain-injury survivors as
well. In some ways, a brain-injury survivor is like any other person who has
suffered catastrophic injury and is in need of treatment and rehabilitation.
Some methods of treatment used for other disability groups may be partially
effective. However, the majority of treatments must be adapted to take into
account the specialized problems and needs of brain-injury survivors. As we
have noted, this book is designed to address many of these special needs and
concerns.
Although a professional may intervene in a survivor's life in many ways,
the initial intervention is always from a medical perspective. After medical
stability has been obtained, then the focus can be upon what constitutes the
secondary characteristics of "brain injury" and the sequelae.
4.1 Neuropsychological Treattnent

The neuropsychological treatment and rehabilitation of brain-injured people


draws upon 1) the general body of knowledge about patterns of posttrauma
recovery, 2) the relationship between the brain and behavior, and 3) methods
for aiding recovery. Brain injury is defined as any structural damage to the
brain that causes the survivor to be dysfunctional in his or her daily life, compared to how he or she was functioning before the injury. Theories and techniques of neuropsychological treatment are constantly being revised to keep
pace with the influx of new information about brain injury, recovery, and
rehabilitation techniques.
Across the spectrum of interventions, from emergency medicine to rehabilitation, certain practices and applications are accepted. Unfortunately, neuropsychological treatment and rehabilitation have not yet reached such a point of
knowledge, application, quantification, and specificity. In fact, the structure of
many treatment/rehabilitation programs continues to be marked by incomplete
intervention strategies or controversy regarding claims of methods of" cures. "
However, in one of the few well-designed studies on outcome, Prigatano and

1. Introduction

his colleagues [13] have outlined certain programmatic features for successful
outcome. His data suggested that individual and family emotional adjustment
can be enhanced through interventions and that this had a positive effect on the
survivor's final adjustment to the disability. Moreover, Stein's work [14] suggests that rehabilitation should be undertaken as soon after injury as possible;
this is seconded by Cope and Hall's research [15] on recovery after injury.
As rehabilitation specialists, we need to address the effects of the injury on
the survivor's long-term life pattern (e.g., social life, leisure activities, work,
and personality functioning). This needs to be done not only in our day-to-day
clinical work but also in long-term outcome studies that will add to the body
of knowledge about such effects. In addition, we need to understand the type of
pathology that produced the injury (e.g., blunt diffuse trauma) and the specific
characteristics of the' brain-injured person (i. e., age, intelligence, educational
level, premorbid personality).
4.2 Programs of Intervention

Research has suggested [18] that rehabilitation of brain-injury survivors can


best be accomplished if a series of symptoms or behaviors are addressed hierarchically: 1) attention, 2) concentration, 3) reaction time, 4) activities of daily
living, and 5) concept formation or abstract reasoning. Consequently, a number
of different intervention modules-indeed, entire treatment programs-have
been established to provide a systematic approach to treatment.
Many rehabilitation programs are based upon goals of behavior change.
Whether the programs focus upon an "activity of daily living" (ADL) approach, or upon training directed toward strengths or weaknesses, they are
still directed toward behavior change.
4.3 Specific Techniques of Treatment

Any treatment technique, whether neuropsychological, chemical, or performance-based, must be grounded in the current scientific knowledge about
brain-injury rehabilitation. However, the addition of a theoretical foundation
to the established applied model of intervention is crucial if this new field of
"neuropsychological treatment and rehabilitation" is to progress.
4.4 Professional Issues

The professional involved in neuropsychological treatment and rehabilitation


of brain-injury survivors needs to perform as a researcher as well as a clinician.
Clearly, it is not enough to simply know the anatomy of the brain or tests or
diagnoses involved in the assessment of brain functioning and the clinical application of neuropsychology. The professional must also strive to improve
the understanding of the relationship between brain and behavior, to further
aid in the posttrauma rehabilitation process.
Furthermore, to be considered a specialist in neuropsychological treatment,
the rehabilitation professional must develop and maintain an expertise in such

10

varied fields as individual neuropsychotherapy, group psychotherapy, cognitive psychology, developmental neuropsychology, clinical neurorehabilitation,
neuroanatomy, recovery of function, personality theory, personality assessment, psychoneuroendocrinology, and general clinical and preclinical research
methods.
Rehabilitation appears to be the fastest growing field in U.S. health care at
the present time, and brain-injury rehabilitation is one of the major segments
of that field. Unfortunately, there is currently a lack of properly trained professionals, and no basic structure for training is in place. This constitutes a
health-care crisis.
4.5 Legal Issues

Legal issues facing brain-injury survivors, their families, and treatment professionals overshadow much of the treatment process. Because the brain injury
often results from the actions of another person, many survivors are embroiled
in litigation to obtain compensation for their medical expenses and loss of
income. Survivors may also be faced with such issues as competency. Although
the specifics in this book are addressed from American and Danish perspectives, the basic concepts can be applied in most countries.
5 SUMMARY

The professional community is in agreement that attempts must be made to


help survivors of brain injury reach the highest level of physical, emotional,
and cognitive functioning. A substantial number of these survivors need rehabilitation. Choices must be made about who receives treatment, what types
of treatment are appropriate, and how long treatment should be continued.
We believe that it would be valuable to work toward an overarching theory
(or number of theories) that can bring together the varied components of treatment. The intent in this book has been to present both historical and current
information about neuropsychological treatment of the brain-injury survivor.
Neuropsychological treatment is a blending of many fields. It falls beyond
the rubrics of neuropsychological assessment, diagnosis, or lesion localization.
Furthermore, neuropsychological treatment does not fit into the framework
of traditional rehabilitation. The detailed and slow individualized process of
rehabilitation after brain injury, as outlined in the following chapters, is a
costly, time-consuming rehabilitative process that involves highly trained
personnel and labor-intensive efforts.
The use of both qualitative and quantitative assessment is crucial for baseline
measurement, problem identification, and strategic planning. Treatment occurs
within the evaluation phase, and assessment becomes treatment.
Professionals must master the concepts underlying the process of neuropsychological recovery and treatment before they can adequately care for the
brain-injury survivor. It is hoped that this book will serve as a guide to understanding and mastering those concepts. In addition, this text may be used as a
basic primer for the theory and application of treatment techniques.

1. Introduction

11

It is also hoped that the information in this book will be a valuable resource
for survivors and their families, as well as for attorneys, insurance advisors,
educators, and social agency personnel-indeed, for anyone whose life has
been touched in some way by the tragedy of traumatic brain injury.
REFERENCES
1. Klauber, M.R., Marshall, L.F., Toole, B.M., Knowlton, S.L. and Bowers, S.A. (1985).
Cause of decline in head injury mortality rate in San Diego County, California. J. Neurosurg.
62, 528-531.
2. Rourke, B., Fisk, J. and Strang,]. (1986). Ncuropsychological Assessmellt ofChildrfll. Guilford
Press, New York.
3. Kraus, ].F. (1987). Epidemiology of brain injury. In Head III;ury, 2nd ed., Cooper, P.R., cd.,
Williams and Wilkins, Baltimore, pp. 1-19.
4. Diller, L. (1987). Neuropsychological rehabilitation. In Neuropsychological Rehabilitatioll,
Meier, M., Benton, A. and Diller, L., cds., The Guilford Press, New York, pp. 3-17.
5. Prigatano, G. (1984). Neuropsychological rehabilitation after closed head injury in young
adults. J. Neurol. Neurosurg. Psychiatry 47, 505-513.
6. Miller, E. (1984). Recovery alld Mallagemellt of Nellropsychological Impairmellts. John Wiley,
Chichester UK.
7. National Head Injury Foundation. (1987). Directory of Head III;lIry Rehabilitatioll SfI'viccs.
Southborough, MA, National Head Injury Fonndation.
8. Engberg and Vinterberg. (1987). The need for follow-up of patients with brain damage of an
acute origin. Udeskrift for Laeger. 149/23 Utllle}.
9. Field, ].H. (1976). Epidellliology of Head III;lIl'ies ill ElIglalld and Wales. Department of Health
and Social Security, Her Majesty's Stationary OffICe, London.
10. Jennett, B., Murray, A., Carlin,]. et al. (1979). Head injuries in three Scottish neurological
units. Br. Med. J. 2, 955-958.
11. Selecki, B.R., Ring, LT., Simpson, D.A. et al. (1981). Injuries to the head, spine and peripheral nerves. Unpublished report. The Neurological Society of Australasia and the Health
Commission of New South Wales and S.A., Sydney, Australia.
12. Levin, H.S., Grossman, R.G., Rose, ].E. et al. (1979). Long term neuropsychological ontcome of closed head injury. J. Neurosurg. 50, 412-422.
13. Prigatano, G.P., Fordyce, D.J., Zeiner, H.K. et al. (1984). Neuropsychological rehabilitation
after closed head injury in young adults. J. Neurol. Neurosurg. Psychiatry 47, 505-513.
14. Finger, S. and Stein, D.G. (1982). Braill Dallla,qc alld Recovery. Academic Press, New York.
15. Cope, D.N. and Hall, K. (1982). Head injury rehabilitation: Benefit of early intervention.
Arch. Phys. Med. Rehabil. 63, 433-437.
16. Oddy, M. and Humphrey, M. (1980). Social recovery during the year following severe head
injury. J. Neurol Neurosurg. Psychiatry 43, 798-802.
17. Weddell, R., Oddy, M. and Jenkins, D. 1980. Social adjustment after rehabilitation: a two
year follow-up of patients with severe head injury. Psychol. Med. 10, 257-263.
18. Ben-Yishay, Y., Silver, S.M., Piasetsky, E.B., et al. (1987). Relationship between employability and vocational outcome after intensive holistic cognitive rehabilitation. J. Head Trauma
Rehabil. 2, 35-48.
19. Jennett, B., Snock,]., Bond, M.R. and Brooks, N. (1981). Disability after severe head injury:
Observations on the use of the Glasgow Outcome Scale. J. Neurol. Neurosurg. Psychiatry
44, 285-293.
20. Malkmus, D., Booth, B.]. and Kodimer, C. (1980). Rehabilitation of the Head b1jl/red Adult:
Comprehellsive Cogllitive Mallagemellt. Professional Staff Assoc. of Rancho Los Amigos
Hospital, Inc., Downey, CA.
21. Sullivan, H.S. (1954). The Psychiatric Illterview. W.W. Norton and Co., New York.

I. THEORY AND INTERVENTION

2. BRAIN INJURY: A PATHOPHYSIOLOGICAL BASIS FOR


NEUROPSYCHOLOGICAL DYSFUNCTION

JENS ASTRUP

1 INTRODUCTION

This chapter outlines the main types of brain injury and their sequelae from
a medical/physiological perspective [1,2]. The goal is to indicate and describe
how brain injury contributes to the survivor's posttraumatic behavior and
neuropsychological funcion. Interventions are aimed at the sequelae that directly
result from the types of injuries described here.
The chapter is addressed to the team of therapists that takes over the care of
brain-injury survivors after primary neurosurgical care has been completed. It
is not addressed to the forum of neurosurgical colleagues, so do not expect to
find a detailed analysis of, for example, the question of the use of dexamethasone, or of prednisone-or of steroids at all-in the treatment of brain
edema. Although the principles of primary care naturally are outlined in the
text, the main emphasis in this chapter is on the clinical and pathophysiological
description of the type of lesions to the brain tissue arising primarily from the
injury itself, as well as secondarily from complicating hypoxia, ischemia, brain
edema, intracranial hematoma, and high intracranial pressure.
Some of these types of lesions are focal, and some are diffuse; some mainly
affect white matter, and others mainly affect gray matter. It is the extent of
these lesions that sets the limits for outcome in the individual patient. Diffuse
lesions usually severely impair cognitive functions, producing dementia, or, in
some cases, a vegetative state. Focal lesions cause more specific deficits, such
as hemianopia, hemiparesis, and aphasia. Clearly, these lesions and their com15

16

I. Theory and Intervention

binations and their affects on brain function form the basis for designing an
individual rehabilitation program.
2 BRAIN INJURIES: THEIR EPIDEMIOLOGY AND PREVENTION

Statistics on brain injury from Western countries are quite consistent countryto-country. In Denmark, about 300 of every 100,000 people are admitted to
the hospital with head injuries each year. Eighty percent of these injuries are
simple concussion; 20% are more severe head injuries. Three percent of the
patients die. Two-thirds of those injured are men, and more than half of both
sexes are young adults. Road accidents account for a little less than half of the
injuries, and falls account for one-third. Alcohol is very often implicated-in
about three-fourths of assaults, one-third of road accidents, and more than half
of falls.
Accordingly, the typical brain-injury patient is a young male injured by a
motor vehicle accident. The epidemiology of motorcycle accidents is particularly descriptive of a specific social behavior. Motorcycle accidents occur most

f r e e s pee ,d
1400
Deaths

by

1200

road
accidents 1000
in
Denmark
800

Iraffic safely campaigns

II

"oil crisis" speed Iimils 80/8016 0 ~m/h


speed Iimils 11 0/90/60 ~m/h
-seal bells
rhelmels

r~M'''I' to~'~:'Wh

600

65 66 67 68 69 70 71 7Z 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87
speed limits refer to highest allowed speed on
free way/high way/ci~ area
Figure 2-1. High and increasing number of deaths by traffic accidents in Denmark during the
years with no speed limits, until the "oil crisis" in 1973. Some traffic campaigns started a few years
earlier corresponded with some decrease in the number of deaths. The speed limits enforced by the
"oil crisis" in 1973, however, had the most significant effect ever recorded by the Danish traffic
statistics. When speed limits were re-increased one year later, the number of deaths increased
accordingly. The enforcement of seat-belt and helmet regulations in 1976-1977 probably had
some damping effect on this increase. Campaigns for safer traffic were intensified, and the speed
limits were again reduced in 1979; a reduction in the number of deaths slowly followed toward a
new low, steady level. The unfortunate re-increase from 1985 is unexplained, but a general
increase in alcohol consumption may be one cause (J.B. Dahlgaard, unpublished data, 1987).

2. Brain Injury

17

frequently at night during weekends; 8 of 10 victims are young male drivers;


one-half have been drinking; and one-third are underage or otherwise without
a valid driver's license.
As some statistics from Denmark will indicate, it is beyond doubt that
legislative enforcements of speed limits and seat-belts and helmet regulations,
together with traffic safety campaigns, better cars, and better road construction, have reduced the incidence and severity of the road accidents. In particular, the effect of speed limits is evident (Figure 2-1).
Another area of concern in modern societies is the sport of boxing. Competitive boxing has the aim of producing a brain injury and should be abandoned as an acknowledged sport activity with Olympic recognition. It carries
the risk of a multiple concussion syndrome with accumulative effects leading
to dementia. Although ending the promotion of boxing may be a worthwhile
goal, only fanatics will go as far as to restrict football, soccer, handball,
horseback riding, bicycling, and the like-all of which are quite often associated with head injury. Enjoying life carries risks, but these can certainly be
minimized by reasonable interventions without the loss of excitement.
3 THE PRIMARY STRUCTURAL INJURY

The primary insult can cause two types of lesions. One is the diffiHe axonal
injury, also called the "shearing" lesion, or simply the diffuse injury. The other
is the contusion, which is a localized lesion. Often, these two types oflesions are
combined; however, one of the types is usually the dominating lesion. The
diffuse injury is indicated by loss of consciousness, brief or prolonged, and
either no changes or only modest changes are seen on computerized tomography (CT) scans. The contusion usually produces focal deficits without loss
of consciousness. The CT scan often shows focal hemorrhagic lesions. Accordingly, a deeply comatose patient with a normal or almost normal CT scan has a
"clean" diffuse injury, whereas the awake and oriented patient with a depressed skull fracture and underlying contusion and a corresponding focal
deficit has a "clean" contusion. A concussion is a mild diffuse injury, but
occasionally minor frontal or temporal contusions are observed on CT scans.
In severe diffuse injuries, the brainstem is also affected, and apnea of a
significant but varying duration is probably present immediately following the
injury. Persons who are dead on impact usually have brains tern lacerations.
Falling blood pressure-either due to the stem lesion or to other body injuries-is often present. These factors add primary global cerebral hypoxia/
ischemia to the structural injury.
Because observations of the patient that are made at the site of the accident
are usually poor, the initial determination of the impact of these insults in
individual cases is uncertain. But they are probably of importance in all severe
cases, placing "on-the-spot" lifesaving measures (freeing of airways, mouthto-nose ventilation, stopping of bleeding, rapid transportation to nearest
hospital) in focus.

18

I. Theory and Intervention

4 LATE AND SECONDARY LESIONS DUE TO ISCHEMIA

A decrease in blood flow to the brain due to compression from diffuse edema,
vascular congestion, and/or intracranial hematomas results in secondary ischemia. Such lesions cause the intracranial pressure (ICP) to rise and, ultimately, cause the cerebral perfusion pressure to fall. Aggravated by systemic
hypotension (due to shock, or other lesions), pulmonary hypoxia (due to
pulmonary contusion, collapse of the lungs, aspiration), disturbed cerebral vasomotor reactivity, and venous congestion, blood flow will fall, and ischemia
will threaten to extend the injury.
Areas at particular risk are in the cortex, in the so-called watershed zones
(i.e., the borders are "on the end of the pipeline" between the areas perfused
by the middle cerebral artery, the anterior and the posterior cerebral arteries).
Severe ischemic lesions to the cerebral cortex may follow. Lateral brain compression (e.g., from an epidural or subdural hematoma), causing a significant
midline shift, may compress the posterior cerebral artery against the tentorial
edge, leading to a posterior infarction.
5 THE DIFFUSE BRAIN INJURY

5.1 Pathology and Repair

The diffuse injury is usually caused by a blunt blow to the head. It may be mild
(with a briefloss of consciousness), severe (producing prolonged deep coma),
or of intermediate severity. If you have watched boxing on TV, you may have
noticed that a "knockout" follows a blow to the head that gives it an abrupt
acceleration and an axial rotation. Deep coma, however brief, is the immediate
consequence. Such an incident is an example of a mild diffuse injury (concussion). It is assumed that multiple interneuronal connections that are diffusely distributed throughout the brain are stretched by the mechanical force
and suffer a brief transmission failure.
Imagine a more severe blow to a front-seat passenger's head when it smashes
into the windshield or dashboard, or the head of a child when he or she has been
hit by a car and is thrown against the street pavement, or imagine a drunk
falling and smashing his head heavily against the sidewalk concrete. When
such violent acceleration-deceleration movements with simultaneous rotation
are being transferred to the soft and plastic brain, the brain undergoes a change
in form. Some parts are compressed, some are stretched-resulting in a diffuse
"shearing" effect that overstretches or disrupts longitudinal structures (axons,
dendrites) in the brain.
In fresh injuries, pathological studies of brain tissue have shown multiple
disrupted axons with retraction bulbs, and multiple capillary disruptions indicated by numerous petechial hemorrhages. If, after suffering severe diffuse
injuries, the patient survives, significant atrophy of the entire brain may
appear. The diffuse injury offers a good explanation of the clinical findings in
brain injury. In mild injury (concussion), a large number of connections fail
when they are stretched. This produces the instant deep coma, but because no

2. Brain Injury

19

disruptions-or few disruptions-occur, function rapidly recovers, and the


patient wakes up after minutes or even seconds. The prolonged coma, however, is due to multiple disruptions.
There is experimental evidence that damaged axons (and therefore, perhaps,
dendrites) may regrow along their previous paths. Neurogenesis may also
occur in brain injury, provided that the neuron cell body survives not only the
shearing injury to its axon and dendrite tree, but also the threatening hypoxial
ischemia associated with the injury or caused later by increased ICP.
Regrowth may explain the often dramatic recoveries that are especially seen
in children and young adults. Other hypothetical mechanisms in transmitter
synthesis, such as temporary disturbances in axonal transport, have been
suggested. As long as the basic mechanisms of the late recoveries are unknown,
therapeutic promotions (e. g., by nerve growth-factor treatment, or by mental
stimulation or the opposite-total rest) remain speculative and should continue to be a subject for clinical trials.
5.2 Clinical Findings in Diffuse Brain Injury

The clinical presentation of diffuse injury covers a wide spectrum of severityfrom the very mild cerebral concussion without loss of consciousness but with
brief amnesia and perhaps confusion (which in boxing leads to "counting"), to
the classic cerebral concussion with a brief coma but recovery within minutes
(which in boxing is the "knockout"), to the more severe acceleration-deceleration-rotation blow that produces instant deep coma, from which victims either
recover more or less completely over hours, days, weeks, or months, or do
not recover (i.e., they die or survive in a vegetative state).
Clinical assessment of the degree of coma is described by Jennett and
Teasdale's coma scale, the Glasgow Coma Scale (GCS), which has gained
wide acceptance, due to its practicality and reproducibility.
Table 2-1. Glasgow Coma Scale'
Verbal response

Eye opening

Best motor response

Coma score = 3-15 points.

Oriented
Confused
Inappropriate words
Sounds
Nil
Spontaneous
To speech
To pain
Nil
Obeys
Localize
Withdraw
Abnormal flexion
Extension
Nil

5
4

3
2
4

3
2
1
6
5
4

3
2
1

20

I. Theory and Intervention

<

2. Brain Injury

21

c
Figure 2-2. CT scan of the brain of a 31-year-old man suffering from a diffuse injury after a
traffic accident. At admission, he was unconscious, with flexion as his best response. The CT scan
was normal, and he eventually made a good recovery.

According to the scale, a patient who is awake after a mild concussion and
who is again oriented and alert will score 15, while the deeply unconscious
patient without any motor reactions and without verbal response or eye
opening to any stimulus scores the minimum of 3 points. The scale describes
the level of coma objectively and with little interobserver variation. It transforms the coma level to a numerical score that is well-defined. It has proven of
invaluable benefit as a tool for comparison among clinical studies of head
trauma and is widely used in observation of patients by the nursing staff.

22

l. Theory and Intervention

As early as one to two days after injury, the level of coma as described by the
scale provides a good index of outcome. The scale attempts a description of the
diffuse injury, not focal lesions (i.e., the response of the best arm is noted, not
of the paretic arm). The presence of aphasia is not taken into account, either.
An unconscious patient has no verbal response or eye opening and is assessed
solely by his or her motor response (i. e., for all practical purposes, by the
motor response in his or her arms as a reaction to pain).
Compression of the nail bed, for example, is used to provoke withdrawal,
flexion, extension, or the ability to localize pain. Abnormal flexion of the arms
is sometimes termed decorticate rigidity, and extension is called decerebrate rigidity.
Observing the pupillary response to light, however, provides other important
information such as signs of progressing lateral trans tentorial herniation that
is compressing the oculomotor nerve.
5.3 CT in Diffuse Brain Injury

CT scanning cannot successfully image a diffuse brain injury, since It IS an


injury that occurs mainly at the cellular level, leaving gross anatomy unchanged. Petechial hemorrhages cannot be imaged, although in the more
severe injuries it is not uncommon to see hemorrhages of a few millimeters in
size scattered in the white matter, particularly in the corpus callosum. Overall,
the brain looks surprisingly normal on CT, considering the often serious
clinical condition of deep coma (see Figures 2-2 and 2-3A,B).
Occasionally, the brain may look "swollen," with compressed basal cisterns
and small lateral ventricles. The tissue may appear slightly hyperdense, due to
diffuse hyperemia and scattered petechial hemorrhages. These patients have
elevated ICP and a high risk of transtentorial herniation. In survivors, when
primary healing is completed, some degree of diffuse central and cortical
atrophy can usually be imaged, depending on the degree of neuronal degeneration and the severity of the lesion (Figure 2-3A).
5.4 Blood Flow and Metabolism in Diffuse Brain Injury

In comatose patients, blood flow and brain metabolism are usually low,
and oxygen extraction, indicated by the arterial-venous oxygen difference
(A VD0 2 ), is within the normal range. The electroencephalogram (EEG) is
suppressed and dominated by low-frequency activity, indicating severely reduced synaptic activity.
The metabolic requirements of synaptic transmission account for about 60%
to 70% of total brain oxygen consumption in the normal awake state. This
explains why oxygen consumption may be as low as 30% in the deeply
comatose state and why blood flow often is similarly reduced. This is the usual
picture in the severe diffuse brain injury, and it docs not appear that traditional
treatments are helpful (e.g., vasoconstriction by hyperventilation or metabolic
inhibition by metabolic depressants-especially since metabolism and flow are
already primarily depressed).

Figure 2-3. (A) CT scan of a 21-year-old man suffering from a severe diffuse injury with primary apnea after a traffic accident. At admission, he was
unconscious, without movement. He survived in the vegetative state. His initial CT scan was normal except for very small scattered parenchymal
hemorrhages. (B) After four months, his CT scan showed severe diffuse brain atrophy.

g
c
-<
....

.,....1:1:1

!'"

24

I. Theory and Intervention

Autoregulation of cerebral blood flow in these patients is often abnormal,


whereas cerebral vascular responsivity to CO 2 is somewhat better preserved.
Occasionally, global hyperemia, with flows much higher than necessary for
metabolic requirements, is observed, particularly in children and young adults.
This indicates global vasoparalysis and a pressure-passive flow; probably these
cases are at risk for developing high ICP due to increased vascular blood
volume and edema formation ("malignant" brain edema). Hyperventilation
to induce vasoconstriction and normalize flow in accordance with metabolic
requirements seems a rational therapy in these patients.
Similarly, metabolic depression induced with barbiturates can be used to
treat some patients who have a metabolic requirement in excess offlow (e.g.,
due to continuous seizure activity), which in the muscle-relaxed patient can be
diagnosed by EEG only.
5.5 ICP in Diffuse Brain Injury

ICP is usually normal in diffuse injury. However, brain swelling, contusions,


or complicating hematomas may cause ICP elevations. Diffuse brain swelling
still carries a poor prognosis, and, in spite of much effort to treat these cases
(ICP monitoring, flow and metabolism measurements, hyperventilation, barbiturate sedation, steroids, mannitol, etc.), most of the patients still die of
high ICP.
5.6 Recovery and Prognosis in Diffuse Injury

Prognosis mainly relates to two factors: patient age and level of coma. It is
likely that the ability to repair the diffuse lesion is age-dependent. Older
patients remaining in coma for days with flexion as the best score do not
recover (death or vegetative survival), while this is true for only 30% of
children with the same clinical score.
The state of prolonged deep coma followed by a dramatic recovery after
weeks or even months is not uncommon in children or young adults but is
never observed in older patients. These clinical facts seem to indicate that the
ability to recover is gradually lost with age. The explanation for this is
unknown. Suggestions have been many, from lack of ability to repair and
regrow (a "neurobiological" explanation) to a lack of surplus and plasticity in
the old brain (a "neurofunctional" explanation). Clinically, it is often noted
that an uncomplicated mild concussion makes old patients dependent, whereas
most young patients are well and without sequelae after a few days.
The mechanisms of recovery from a diffuse injury are unknown. The
complete recovery from a concussion within minutes seems to indicate "overstretching" without significant structural damage. But which mechanism is
involved in the momentary, deep, but brief failure of brain function, and how
does it recover?
Clearly, the processes behind prolonged coma and its slow and often incomplete recovery are different. Regrowth may explain recovery over months,

2. Brain Injury

25

but the degree of neuronal (or glial) cell death, the functional plasticity of the
surviving structures, and relearning and refinement of regrowth must influence the outcome. Perhaps repair of intra-axonal processes damaged without actual axonal or dendrite disruptions is faster; this may explain the recovery over days in the intermediate cases. The answers to these questions
may in the future tell us how recovery may be influenced by pharmacological
therapy, medical care, and rehabilitation.
5.7 Assessment of Outcome from Diffuse Injury

The widely applied Glasgow Outcome Scale (GOS) recognizes five categories:
1) death, 2) vegetative survival, 3) severe disability, 4) moderate disability, and
5) good recovery. Severe disability indicates dependence on other people
(often in an institution but sometimes at home). This is often due to a combination of severe neurological and mental disability, but one or the other
may be the sole determinant of dependence. Moderately disabled patients
are independent and can look after themselves and can go out in public, and
some can work. Typically, they have some cognitive dysfunctions or have
undergone personality changes, often in combination with focal neurological
deficits. Patients with good recovery may have some deficits, but they are
able to resume a normal social life. They are able to return to work, although
for a number of other reasons they may not do so. The candidates for
neuropsychological rehabilitation will typically be in the moderately disabled
group.
6 THE CEREBRAL CONTUSION

The term contusion refers to the mechanical destruction of tissue at the site of
the blow. It may be a skull fracture depressed into the brain tissue. It may be
the result of a gunshot wound. More often it is a combination of a focal lesion
and a diffuse injury, with the focal lesion appearing at sites where the brain
impacts on the inside of the cranium. This occurs directly under the blow and
across from that site, where the brain by a contre-coup effect hits the inside of
the skull. In particular, it occurs where the temporal or frontal poles are being
"trapped" in bony corners of the skull, or along the sharp sphenoidal wing,
and above the rough orbital roofs. Accordingly, the common sites where one
might expect to find contusions are the temporal and frontal poles, the area
around the sylvian fissure, and the basal parts of the frontal lobes.
A contusion is a focal lesion, and as such it causes a focal deficit. A
contusion does not by itself explain unconsciousness. If the patient is unconscious, it is due to a concomitant diffuse injury. Usually these two types of
lesions are combined, but even in lesions penetrating the head (e.g., knife or
bullet wounds) preserved consciousness may be observed. Similarly, it is not
uncommon to find a frontal pole contusion in combination with cerebral concussion (Figure 2-4).

26

I. Theory and Intervention

Figure 2-4. CT scan of a 16-year-old girl suffering from a concussion. She was hit by a car while
she was walking. She was unconscious for some minutes, but upon arrival at the emergency room
she was awake and oriented, although she was drowsy, complained of a headache, and was vomiting. Due to her continuous headache and vomiting, she was transferred to the neurosurgical
department. Her CT scan shows a rather large right basofrontal contusion. She recovered rapidly
and was discharged home after one week; there were no sequelae.

Figure 2-5. (A) Skull film of a five-year-old boy who had been hit by a motorbike. He suffered a
mild diffuse injury with a brief loss of consciousness and an open skull fracture with dilacerated
brain tissue in the wound. He underwent acute surgery with removal of bone fragments and dura
repair. He made a quick recovery without apparent focal deficits. The skull film indicates the
depressed fracture. (B) CT scan after eight months indicates healing with atrophy at the site of
contusion underlying the fracture.

28

I. Theory and Intervention

As indicated in Figure 2-4, contusions are seen as localized hemorrhages on


CT, and they heal with a focal atrophy (Figures 2-SA, B). Larger contusions
may swell due to additional hemorrhage and edema formation, causing mass
effect and making evacuation through craniotomy necessary for decompression.
7 MONITORING AND INVESTIGATIONS IN
THE UNCONSCIOUS BRAIN-INJURED PATIENT

After an unconscious brain-injury survivor has undergone acute primary treatment, assessment, and observation, additional observations should be undertaken. A CT scan should be performed. A lateral view of the cervical spine
(CI-C7) must be made so that fractures or dislocations in that area are not
missed.
If significant contusions or signs of brain swelling are indicated by CT, ICP
is usually monitored. This may be accomplished in a number of ways, but the
most widely used method is the "subarachnoid bolt" or the "Richmond
screw." This is a metal screw with a central bore which, when fixed in a burrhole and a dura-arachnoid opening, provides a connection between the
subarachnoid space through the bolt and a water-filled line to the pressure
transducer. The patient's EEG may be continuously recorded and analyzed by
a cerebral function monitor for seizure activity and wave frequency spectrum.
Respiration and ventilation are carefully observed. Prolonged mechanical
ventilation and intubation require that a tracheostomy be performed within
one to two weeks. As already discussed, measurements of cerebral blood flow
and metabolism are often performed with the aim of selecting cases with
relative hyperemia or relative hypermetabolism to specific therapy.
It is beyond doubt that intensive care has improved outcome, but definite
evidence of a beneficial effect of specific measures such as ICP monitoring and
cerebral blood flow measurements, and therapeutic regimens with hyperventilation, barbiturate coma, or osmotherapy, has not yet been provided by
clinical trials.
8 COMPLICATIONS TO BRAIN INJURY

8.1 Skull Fractures

Depressed fractures with underlying contusion (Figure 2-SA) usually require


elevation and dural repair. These fractures carry a high risk of posttraumatic
epilepsy. Otherwise, skull fractures require no specific treatment. They are of
interest in view of their associated complications such as cranial nerve lesions,
dural tears with otorrhea or rhinorrhea, and meningeal artery tear with epidural hematoma formation. Dislocated fractures through the facial cranium
often require repositioning and fixation.
8.2 Cranial Nerve Lesions

The cranial nerves may be injured by fractures of the skull base. Most common is damage to the vestibulocochlear nerve and the inner ear by a fracture

2. Brain Injury

29

through the petrosal bone. Recovery of the vestibulocochlear nerve is unlikely. The lesion may be associated with otorrhea (see Section 8.3). The facial
nerve may be damaged as well, but it may recover.
The optic nerve and the oculomotor nerve may occasionally be damaged by
fractures through the orbit leading to ipsilateral vision impairment, diplopia,
or blindness. The oculomotor nerve may recover, but the optic nerve rarely
does. The olfactory nerves are often damaged bilaterally, leading to anosmia
(the lack of ability to detect odors). This may follow an occipital blow
overstretching the filae that perforate the lamina cribrosa, or it may follow a
fracture through this structure, and then it may be associated with rhinorrhea.
Anosmia is hardly ever reversed. Other cranial nerves are seldom damaged.
8.3 Otorrhea or Rhinorrhea

Otorrhea or rhinorrhea indicates that there has been a fracture with a duralarachnoid tear through which cerebrospinal fluid (CSF) can pass from the
subarachnoid space either to the middle ear and out through the external ear (if
a tear in the tympanic membrane is present), or through the eustachian tube
to the pharynx, or to the nasal cavity through the frontal, sphenoidal, or
ethmoidal sinuses, or through the lamina cribrosa. Otorrhea or rhinorrhea in
the acute state is quite common; it is treated prophylactically with antibiotics
in an attempt to prevent bacterial meningitis. Most cases arrest spontaneously,
but a few persist and require continuous antibiotic treatment until the tear is
surgically closed. If the otorrhea or rhinorrhea remains unnoticed or untreated,
the risk of bacterial meningitis, often caused by pneumococci, is very high.
Even the smallest amount of rhinorrhea noticed when the patient bends
forward should be taken seriously.
8.4 Intracranial Helllatollla

Intracranial hematomas can be extradural, subdural, intracerebral, or combinations thereof. The subdural hematoma has an acute and a chronic form.
Hematomas form when a torn intracranial blood vessel bleeds into the surrounding tissue. Extradural bleeding may result from damage to the meningeal artery and its branches in the dura. Subdural bleeding is usually caused by
tears in the bridging veins or the veins or arteries on the brain surface-in
particular, the middle cerebral artery branches in the sylvian fissure; and
intracerebral bleeding results when parenchymal vessels are damaged in a
contusion. The larger the damaged vessel, the more rapid is the hematoma
formation, and the more acute and serious are the clinical symptoms. Bleeding
from a vein is a low-pressure bleeding and may not cause any acute symptoms,
but may eventually cause a chronic subdural hematoma. Only about 1 % -2%
of all patients admitted to the hospital for head injury develop a "surgical"
hematoma. Complicating intracranial hematoma carries a high risk of posttraumatic epilepsy.

30

1. Theory and Intervention

Before describing the various types of hematomas, it is reasonable to discuss


the pathophysiology and clinical signs of an expanding intracranial mass lesion,
including the clinical syndrome oflateral trans tentorial herniation and its acute
treatment.
8.4.1 Expa/ldin/? Intracranial Mass Lesion

Because the intracranial volume is fixed (except in infants with open skull
sutures and fontanelles), a mass lesion can only expand if there is a corresponding reduction of other intracranial volumes. The intracranial volume
buffer is provided by the CSF volume and the volume of venous blood,
allowing a mass to expand to about 30 ml to 50 ml until the ICP begins to rise
(Figure 2-6).
The volume buffer capacity is somewhat higher in chronic alcoholics and in
old people who have cerebral atrophy, and it is smallest in children and young
adults. When the volume buffer capacity is exhausted, the patient leaves the
"flat" part and enters the "steep" part of the pressure-volume curve.
Therapy accordingly aims at bringing the patient from the steep to the flat
part of the curve. This is accomplished by removing the hematoma or, until
that can be done, by providing symptomatic relief by either hyperventilation
to constrict the cerebral arterioles and thereby reduce the intracranial arterial
blood volume, or by intravenous osmotherapy (mannitol) to extract water
from the brain's extracellular volume.

ICP

PATV

30-50ml

Figure 2-6. Pressure-volume (P-V) curve (i.e., the relation between an expanding pathological
intracranial volume, for example, a hematoma, and the subsequent rise in intracranial pressure
[ICP]). There is an almost perfect volume-buffer effect for the first 30 ml to 50 ml in volume
expansion without any significant rise in ICP (the "flat" part of the P-V curve). When the volumebuffer mechanisms are exhausted, the ICP rises steeply (the "steep" part of the P-V curve). Lethal
volume expansion is about 80 ml, or more in case of cerebral atrophy (alcoholism and old age).

2. Brain Injury 31

Clinical signs of increasing ICP (e.g., in a patient with a cerebral concussion)


are enhanced headache, nausea and vomiting, a fall in level of consciousness
(see Glasgow Coma Scale [Section 5.2]), a progressing bradycardia, and usually
also a rise in blood pressure (known as the "Cushing response").
An expanding hematoma (e.g., an extradural hematoma) will force the
mesial part of the temporal lobe (uncus and gyrus hippocampus) over the
tentorial edge, thereby leading to ipsilateral compression of the oculomotor
nerve and compression of the cerebral peduncle and the posterior cerebral
artery. Signs of oculomotor nerve compression are ipsilateral pupil dilatation
and sluggish-to-absent contraction to light, ptosis, and lateral deviation of the
eyeball.
Compression of the cerebral peduncle results in abnormal flexion, which
progresses to extension (decerebrate rigidity) of arms and legs as a response to
pain. Signs of oculomotor and peduncle compression are late signs, indicating
imminent completed trans tentorial herniation with development of brain death.
Clear signs of an expanding intracranial mass lesion represent an emergency,
requiring tracheal intubation, hyperventilation, osmotherapy, urgent CT, and
a craniotomy for hematoma evacuation. A temporal burr-hole, ipsilateral to
the side of pupil dilatation, may be indicated for decompression if transportation to the neurosurgical service will take a long time.
8.4.2 Extradural Hematoma

The meningeal artery and those branches that are part of the dura (i. e., just
under the skull) may be damaged by a skull fracture. Bleeding from an artery
expands in the extradural space by dissecting off the dura from the skull bone.
The fact that the dura is loosely connected to the skull in children and becomes
more firmly connected with increasing age explains the decreasing incidence of
extradural hematomas with age; consequently, extradural hematoma in the
elderly population is extremely rare. Low temporal fractures running into the
skull base may damage the main trunk of the artery and cause profuse bleeding, whereas fractures that are situated high on the cranial vault may only
damage the smaller arterial branches, causing bleeding that is slower and less
severe, and which is sometimes self-limiting.
An extradural hematoma may occur without a visible skull fracture, particularly in infants. Extradural hematomas occur most frequently in people
injured in their 20s and 30s, but in general they are rare and comprise only 10%
to 20% of all traumatic intracranial hematomas. In patients with a lucid "free"
interval (i.e., who have a relatively intact brain), the prognosis may be excellent, but it depends on early diagnosis and treatment. Still, 20% to 30% of
patients with an extradural hematoma die or have a poor outcome.
An extradural hematoma may be indicated by symptoms of increasing ICP,
and-ultimately-by signs of lateral trans tentorial herniation. It is confirmed
by CT (Figures 2-7 A and 2-7B). Symptomatic therapy that reduces ICP,
such as hyperventilation, osmotherapy, and occasionally a decompressing

32

I. Theory and Intervention

2. Brain Injury

33

burrhole, are temporary lifesaving procedures that can be performed until a


craniotomy can be undertaken.
8.4.3 Subdural Hematoma

An acute subdural hematoma is usually caused by bleeding from one of the


arterioles branching in the sylvian fissure; such damage results from a contusion or dilaceration of the tissue as part of a severe head injury. Accordingly,
the prognosis is rather poor and varies with the severity of the initial injury.
Diagnosis is confirmed by a CT scan (Figures 2-SA and 2-SB).
A chronic subdural hematoma arises from a minor and otherwise insignificant head trauma that has often been forgotten by the patient by the time the
hematoma symptoms appear. Blood from a ruptured bridging vein running
from the brain surface toward the superior sagittal sinus accumulates in the
subdural space. The bleeding itself rapidly stops and usually causes no initial
symptoms, but as the hematoma disintegrates, it accumulates water by osmosis, and minor rebleeding in the surrounding membranes often occurs (Figure
2-9).
This volume expansion leads to slowly progressing pressure symptoms and
often fluctuating focal signs over several weeks. The risk is enhanced by
cerebral atrophy that is due to chronic alcoholism or old age, as well as by
blood coagulopathy (e. g., from alcoholic liver disease) or by anticoagulation
treatment. Prognosis upon evacuation is usually good, because the brain is not
primarily injured. About 10% of patients need reevacuation.
8.4.4 Intracerebral Hematoma

These hematomas develop by intracerebral bleeding in a contusion as part of a


severe head injury. As a rule, their locations follow the contusions (temporal
lobe, frontal pole, basofrontal area). Treatment is primarily conservative
with monitoring ofICP, but surgical evacuation may be necessary for decompreSSIOn. Prognosis is rather poor and varies with the severity of the initial
lllJury.
8.5 Hydrocephalus

Posttraumatic hydrocephalus (Figure 2-10) is caused by a disturbance in CSF


dynamics with an increased resistance to fluid flow through the subarachnoid
space over the hemispheres and its passage into the venous sinuses. PostFigure 2-7. (A) CT scan of a sevcn-year-old girl with an extradural hematoma located in the
temporal region. She was hit by a car and suffered a mild diffuse injury with brief unconsciousness. After she regained consciousness, she con tinned to have a headache and vomiting, and within
hours her pulse rate fell to 55, her level of consciousness decreased, and her right pupil dilated and
showed a sluggish reaction to light. She made a quick and good recovery after the hematoma was
evacuated. (B) CT scan of a 20-year-old man hit in the frontal region by a wine bottle. He had no
initial loss of consciousness, but after a few hours, his level of consciousness progressively decreased. He made a quick and good recovery after the hematoma was evacuated.

34

I. Theory and Intervention

2. Brain Injury

35

Figure 2-9. CT scan of a 76-year-old man who slipped on an icy sidewalk and hit his head
mildly. Over the next weeks, he had persisting headache, and gradually he developed gait instability and an unattended mild left hemiparesis. The CT scan indicates a chronic subdural
hematoma with membranes extending from "pole to pole" over the right hemisphere. He made a
good recovery after the hematoma was evacuated through a small craniotomy.
Figure 2-8. (A) CT scan of a 38-year-old woman who jumped from the second floor of a building. She was awake when brought to the emergency department, but her level of consciousness
rapidly decreased, and her right pupil dilated. She was immediately intubated and hyperventilated
and was transferred to neurosurgery. After CT scans were made and the acute subdural hematoma
on the right side was evacuated, multiple left costal fractures with pneumothorax and lung
contusion as well as fractures of the mandible and pelvis were diagnosed. She made a good
recovery. (B) CT scan of a 48-year-old chronic alcoholic man found at the foot of his kitchen
stairway. He was deeply unconscious without motor response and with bilateral fixed, dilated
pupils. With these signs of completed trans tentorial herniation and a hopeless cerebral prognosis,
evacuation of the hematoma was not considered justified.

36

I. Theory and Intervention

2. Brain Injury

37

traumatic hydrocephalus occurs in approximately 10% of the patients who


have significant amounts of blood in the subarachnoid space. It develops
slowly over weeks or months, usually without signs of increased ICP, but
with diffuse cerebral symptoms such as arrest of progress during rehabilitation,
gait disturbances, incontinence, and dementia. Distinguishing hydrocephalus
from cerebral atrophy can be difficult. Treatment by a ventriculoperitoneal
shunt may relieve the symptoms.
8.6 Carotid-Cavernous Fistula

This rare complication develops if the intracavernous part ofthe carotid artery
is damaged by a skull base fracture. The arterial pressure is transmitted to the
venous cavernous sinus, resulting in a systolic bruit from blood flow through
the fistula, vascular congestion (mainly in the eye orbit, with pulsating exophthalmos), ophthalmoplegia, deterioration of vision, and pain. Balloon
closure of the fistula, or the trapping of blood by ligation of the carotid artery
below and above the fistula, is usually effective.
8.7 Disturbances of Pituitary and Hypothalamic Functions

In spite of fractures through the sella, the pituitary gland is resistant to trauma.
Some instances of posttraumatic hypopituitarism have been described. Hypopituitarism should be considered in patients with massive fatigue, asthenia,
weight loss, pallor, hair loss, amenorrhea, loss of libido, and, in children,
growth retardation.
Differentiating clinically between hypopituitarism and mild dementia with
"neurasthenia" and mental depression is difficult, but in most cases the
pituitary function tests are normaL Because head injury that is combined with
signs of hypothalamic contusion, such as diabetes insipidus (lack of antidiuretic hormone [ADH] with polyuria and water loss), or syndrome of in appropriate ADH secretion ([SIADH] [release of ADH with water retention, edema,
and hyponatremia]) carries a poor prognosis, these syndromes are seldom seen
permanently in survivors.
8.8 Posttraumatic Epilepsy

Posttraumatic epilepsy is a serious complication of head injury. It should be


diagnosed and treated early (see Chapter 3).
Uncomplicated mild diffuse brain injury carries a low risk of permanent
posttraumatic epilepy-only 1% to 2%. However, complications such as a

Figure 2-10. (A) CT scan of a 24-year-old man who suffered a severe diffuse brain injury in a
traffic accident. The initial CT scan indicates blood in the subarachnoid space on the brain surface
and in the posterior horns of the ventricles. (B) After six weeks of slow recovery, there was an
arrest in progress followed by deterioration. Four months after the injury, the CT scan shows
hydrocephalus with periventricular lucency indicative of transependymal fluid absorption.
Ventriculoperitoneal shunting improved his condition.

38

I. Theory and Intervention

depressed fracture with underlying cerebral contusion, or an intracranial hematoma (any kind), increase the risk of epilepsy to about 35%. If an early seizure
(within the first week) occurs in patients in these risk groups, or if their coma
is prolonged (as indicated by a posttraumatic amnesia of more than 24 hours),
approximately three out of four patients will develop permanent epilepsy.
Generalized, focal, and temporal lobe (i.e., "partial complex") seizures are
seen, but not petit mal seizures.
8.9 Posttraumatic Dementia

This is the most serious complication to head injury. It is a general experience


that even quite severe focal deficits (e.g., hemiparesis, aphasia, hemianopia)
may be well-tolerated as long as intellect and personality are preserved. Cognitive dysfunction and personality changes are usually the result of a diffuse
injury, not a focal lesion. However, bilateral frontal lobe contusions may lead
to specific frontal lobe syndromes.
Sometimes dementia may be the only symptom of posttraumatic hydrocephalus, which should be considered if there is an arrest of the patient's
progress or a regression in his or her accomplishments. Unrecognized temporallobe epilepsy with frequent attacks may imitate dementia. Occasionally,
posttraumatic pituitary insufficiency may add symptoms somewhat similar to
the posttraumatic dementia syndrome.
In addition to their dementia, the patients usually suffer focal deficits (from
localized brain contusions), cranial nerve lesions, epilepsy, liquorrea or hydrocephalus, all adding to the list of handicaps burdening the individual patient.
The burden can be relieved by proper medical treatment and rehabilitation.
REFERENCES
1. Jennett, B. and Teasdale, G. (1981). MilIIagclllellf of Head Illillyies, F. A. Davis Co., Philadelphia.
2. Cooper, P. R., ed. (1987). Head IlIillYY, 2nd ed, Williams & Wilkins, London.

3. ISSUES IN BEHAVIORAL NEUROLOGY AND BRAIN INJURY

DA VID F. LONG

1 INTRODUCTION TO BERA VIORAL NEUROLOGY

1.1 Definition and History

Behavioral neurology is a subspecialty of neurology concerned with the diagnosis and management of patients who have linguistic, perceptual, cognitive,
or behavioral impairments due to brain injury or disease. The field has an
extensive heritage in the original common historical roots of neurology and
psychiatry. However, the modern growth of behavioral neurology can be
traced from the pioneering work of Norman Geschwind in the 1960s [1, 2].
Behavioral neurology has also shared a particularly intimate relationship with
neuropsychology, both clinically and in the scientific investigation of the
neurological basis of human behavior. The tremendous recent growth of
behavioral neurology and the closely related field of neuropsychiatry demonstrates the increasing recognition of the importance of brain function in
determining behavior.
1.2 The Neurobehavioral Approach

Several key tenets distinguish the neurobehavioral approach from other approaches. One is the use of the detailed mental status examination, which is
correlated with underlying brain anatomy and function. Although the different
The author wishes to express appreciation to Carole Wyman, Sherry Burnes, and Doreen Lampert for secretarial
assistance; to Lester S. Dewis, M.D., and Tahereh Ahdieh, M.D., for reviewing the manuscript and providing
helpful suggestions; and to his wife, Kathy, for her support and understanding.

39

40

I. Theory and Intervention

mental status components are systematically addressed, the examination is the


antithesis of a fixed test battery. Rather than relying on rote questions, the
examiner may glean the most information from a direct practical approach.
For instance, asking a patient about future plans is often a more effective way
of learning about the person's insight and judgment than asking him or her
standard questions about discovering smoke in a crowded theater. During the
exam, potential problem areas are probed in depth, and hypotheses are tested
and refined.
Examination techniques must be precise. For example, even if a patient does
not comprehend a verbal command, he or she could respond appropriately,
due to an inadvertent accompanying gesture. This does not preclude the
examiner's purposely providing a combined visual-verbal stimulus on another
occasion, for therapeutic purposes.
Sometimes precise examination allows remarkable insight into the nature of
a patient's difficulty. In 1962, Geschwind and Kaplan described a patient who,
when blindfolded, was unable to verbally name objects placed in his left hand
[3]. The patient could name objects similarly placed in his right hand or presented visually. He could also use his left hand to draw an object previously
held in the left hand, or to select it from a group of objects. The patient could
use his left foot to reproduce a pattern drawn on the left hand but not the right.
These and related examination findings allowed the authors to postulate the
callosal lesion that was subsequently confirmed at a postmortem examination.
Observations must be distinguished from interpretations. For example, one
can observe that a patient is crying but cannot interpret whether he or she is
depressed without obtaining further information, including asking about how
he or she feels. Projection of how another person might feel in a similar situation only obscures the issue and cannot take the place of direct information
(N. Geschwind, personal communication). On the other hand, during an
examination it is also helpful to notice which objective behaviors are generating certain impressions or reactions. For instance, an examiner who is beginning to feel impatient may come to recognize that one of the patient's major
deficits is slowness.
The neurobehavioral examination places heavy emphasis on the sequence of
tasks performed and the types of errors made. A far better understanding of
the nature of a patient's difficulty can be gained from such an emphasis on process, as well as on achievement, than from reliance on numerical test scores
alone [4]. Furthermore, understanding the process often helps identify an
appropriate remediation approach for the patient. For example, a patient who
solves math problems incorrectly because he or she neglects the numbers on
the left side of the problems should work on spatial attention skills rather than
on arithmetic skills per se.
The behavioral neurologist organizes and interprets the examination with
known clinical-pathological correlations in mind. The classic approach, localizing the lesion, is probably less important than is the use of available anatomi-

3. Issues in Behavioral Neurology

41

Figure 3-1. CT scan showing extensive exencephalomalacia in the left temporal region
surrounding Wernicke's area, as well as diffuse atrophy. The scan was performed late after
traumatic brain injury including left temporal intracerebral hemorrhage. Clinically, the patient
exhibited persistent Wernicke's aphasia.

42

I. Theory and Intervention

cal information (including scans) to direct the examination, estimate prognosis,


or direct therapy efforts. The increased likelihood of alexia after a known left
occipital lobe infarction might prompt an especially detailed assessment of the
patient's reading skills. Recognition of involvement of Wernicke's area and
adjacent regions might lead to the prediction of comprehension impairment
(see Figure 3-1) [5]. A lesion in Broca's area without involvement of the right
hemisphere or posterior left hemisphere might enhance a patient's candidacy
for Melodic Intonation Therapy [5-7].
1.3 Behavioral Neurology in Head Injury

Behavioral neurology is assuming a unique role in the management of brain


injuries from head trauma for several reasons. The cognitive, personality, and
emotional deficits caused by these injuries have consistently been the most
devastating long-term problems faced by patients and their families [8-10].
Disentangling the complex symptomatology requires precise examination and
an understanding of the underlying brain pathophysiology and connectivity.
In this regard, it is important to note that multiple lesions are not simply additive but may cause specific syndromes not explainable by anyone of the contributing lesions [1]. Examples of this phenomenon include pure alexia (impaired
reading) [1], prosopagnosia (impaired facial recognition) [11], and one form of
pure word deafness [12]. Most importantly, "Neurobehavioral disorders have
changed from diagnostic curiosities to remediable conditions, and the recognition and employment of appropriate treatments have become a reality" [13].
This is especially true for traumatic brain injury, for which the "impure" diagnostic lesions of the past era have become the management challenges of the
present.
In keeping with this emphasis on treatment, the behavioral neurologist typically functions as part of a brain-injury treatment team. Detailed assessment of
the brain-injured person allows not only specific deficit recognition but also
identification of areas of preserved relative strength that may be utilized in
specific treatment techniques. Even more importantly, understanding the
nature of a patient's disorder and its prognosis allows better integration with
personal and family goals and with adjustment and rehabilitation issues.
2 CLINICAL-ANATOMICAL CONSIDERATIONS IN BRAIN INJURY

2.1 Pathological Complexity

Much of the early work in the field of brain injury tended to look at outcomes
for groups of patients without considering the underlying pathophysiology.
Recent major advances have included the recognition of the epidemic proportions of the problem [14], the development of overall measures of severity
[15-16], and the delineation of stages of recovery [17-19]. Consideration of
the underlying pathology is necessary for optimum individual patient management. Given the pathological complexity in traumatic brain injury [17], comparing groups of pathologically unselected head-trauma patients is similar to

3. Issues in Behavioral Neurology

43

comparing groups of "neurological" patients with intermixed strokes, multiple sclerosis, and Alzheimer's disease. Disentangling subgroups of braininjured patients is more difficult, however, since multiple processes and lesions
are typically interacting within a single patient. One useful approach is to
divide injuries into focal, diffuse, and mixed groups [20-22]. Further subdivision can be by brain-injury location and size in focal injuries and by process
and severity in diffuse injuries [20-24]. It is important to remember, however,
that management decisions for a particular patient are best made on the basis of
individual assessment.
2.2 Diffuse Axonal Injury

At one time, patients with major clinical impairments but no evidence of


hemorrhage were considered to have a "brainstem contusion." We now know
that the vast majority of patients with brainstem lesions not secondary to herniation have diffuse axonal injury (DAI). This is the pathological constellation
of widespread shearing injury to axons in association with focal lesions in the
corpus callosum and rostral dorsolateral brainstem [25, 26]. Axonal damage
without associated brainstem lesions has been described in milder injures

[26, 27].
Clinically, DAI is characteristically seen following motor vehicle accidents,
probably because the duration of the inertial acceleration in such accidents is
longer than in falls. Loss of consciousness is immediate at impact. In primates,
the duration of unresponsiveness parallels the extent of D AI, and a similar correlation of duration of both coma and posttraumatic amnesia with DAI severity
in humans seems likely, although data are limited [22, 28, 29].
Recognition of the inertial mechanism of injury and the correlation of coma
duration with DAI severity have prompted speculation that less severe forms
ofDAI may be the pathophysiological basis for concussion [25,26]. The finding of internal axonal disruption without axonal transection in milder experimental injuries lends support to this view [27, 30].
Problems with identifying DAI in survivors have limited our knowledge of
its characteristic clinical sequelae. The CT (computerized tomography) scan
has sometimes allowed identification of DAI when specific shearing hemorrhages have been observed (see Figure 3-2) [31, 32]. However, DAI has been
found at postmortem examinations in patients with either normal scans or
diffuse swelling [33]. Technical and anatomical factors often limit the ability to
visualize the characteristic callosal or brainstem hemorrhages [24, 32, 34, 35].
However, the group of patients with shearing hemorrhages documented by
CT scans have fared worse than the group with only diffuse swelling seen on
their CT scans [24].
In a small series of patients with no focal CT lesion, two groups were identified [22]. Those patients with long periods of unresponsiveness typically exhibited quadriparesis, dysarthria, ocular brainstem signs, slowness, and temper
outbursts; those with shorter unresponsive periods showed good motor re-

44 I. Theory and Intervention

Figure 3-2. CT scan demonstrating a hemorrhage in the region of the corpus callosum and
columns of the fornix. Callosal hemorrhage is characteristic in diffuse axonal injury (DAI),
although it is frequently difficult to demonstrate by CT scan.

3. Issues in Behavioral Neurology

45

co very but exhibited impulsivity, decreased insight, and mental control difficulties. The widespread axonal involvement would certainly be logically
compatible with attention-related deficits. However, it has been difficult to
separate DAI from frontotemporal contusion in the "core syndrome" of mental control impairment [21]. Of particular interest is the finding that patients
with even mild degrees of presumed diffuse axonal injury fail to suppress the
vestibulo-ocular reflex [36]. This finding is consistent with the component of
midbrain involvement that is known to occur in DAI, and it would not be
anticipated with isolated cortical contusion [37].
Magnetic resonance imaging (MRI) is proving to be useful in demonstrating
lesions not well seen on CT scans (see Figure 3-3). Both cortical lesions consistent with contusions and white matter lesions consistent with DAI have been
described on acute MRI scans [38, 39]. Late MRI scans in a small number of
patients with severe diffuse injury showed ventricular enlargement and parasagittal lesions compatible with gliosis [40]. Thus, MRI scans may prove particularly useful in distinguishing DAI from other pathological conditions, such
as contusions not well-visualized by CT scans.
2.3 Diffuse Cerebral Swelling

Acute CT scans often demonstrate evidence of diffuse cerebral swelling. Sometimes this swelling can be due to increased vascular perfusion or "hyperemia"
rather than edema [41, 42]. Diffuse cerebral swelling is characteristically seen
in children, in whom it is the most common CT finding [41]. In contrast to
patients with DAI, who are comatose from onset, patients with diffuse swelling
can exhibit a lucid interval.
The basal cisterns are usually well-visualized on CT scans. When brain swelling becomes severe, with increased intracranial pressure (ICP), an attenuation
or obliteration of these basal cisterns can occur, which carries a poor prognosis

[43-45].
2.4 Hypoxic Ischemic Injury

The importance of hypoxic and ischemic injury in determining outcome from


traumatic brain injury (TBI) is often underestimated. In a recent consecutive
pathological series of 151 cases of nonmissile head injury, 91 % of the patients
had evidence of ischemic brain damage [46]. Although about 50% of patients
with extensive generalized anoxic damage had a known history of cardiopulmonary arrest or status epilepticus, 74% of patients with no such history had
pathological evidence of ischemic brain damage. Ischemia was often focal and
related to increased ICP, but its impact cannot be ignored. The hippocampus
and basal ganglia were most frequently affected, with related potential effects
on memory and motor functions, respectively.
It is often difficult to determine the relative contributions of anoxiclischemic
injury and primary traumatic injury in severely impaired head-trauma patients.
However, in nontraumatic patients with known anoxia, early serial examin-

3. Issues in Behavioral Neurology

47

ations-including level ofconsciousness, pupillary responses, and oculocephalic


maneuvers-are highly predictive of outcome [47]. Brainstem structures are
characteristically more resistant to anoxia than is the cerebral cortex, whereas
structures such as the hippocampus, globus pallidus, and cerebellar Purkinje
cells are particularly sensitive to anoxia [48, 49].
The duration, severity, and type of anoxia (e.g., decreased blood flow versus decreased oxygen saturation) are all relevant to the pattern of pathological
involvement. Disturbances of attention and memory are common after anoxia,
but the prognosis is generally not as good as after trauma. Severe apraxia has
been encountered in a number of postanoxic patients and presumably is attributable to the greater amount of parietal involvement in these patients, in
contrast with the predominance of frontotemporal involvement after head
trauma.
2.5 Focal Parenchymal Contusions and Hemorrhages

In missile injuries, the brain can be injured in virtually any part, with the nature
of sequelae limited primarily by the patient's ability to survive. In closed-head
injuries, however, contusions occur consistently in characteristic locations,
specifically, the frontopolar, orbitofrontal, and anterior temporal regions [17,
50, 51]. Contusions less frequently occur in other locations, regardless of the
direction of the blow, although they are often larger on one side of the brain
than on the other.
A change in behavior is the hallmark of frontal lobe injury [52]. The nature
of this change has recently been reviewed extensively [53-56]. Orbitofrontal
lesions typically are associated with impulsivity, distractibility, and impaired
social competence or a so-called "pseudopsychopathic" syndrome. Dorsolateral
lesions are more typically associated with apathy, perseveration, motor sequencing difficulty, and a "pseudodepressed" appearance [52]. Mesial frontal
lesions are typically associated with akinesia [57]. Severe and lasting deficits are
most often seen with bilateral involvement [54].
Orbitofrontal contusions are most characteristic in brain injury, but more
extensive lesions and syndromes are not uncommon. It is also important to
recognize that "frontal" type behavioral disorders following brain injury are
not always indicative of anatomical lesions of the frontal lobes. These symptoms may be the result of focal frontal hypometabolism [42] or disruption of
frontal lobe connections with neocortical, subcortical, brainstem, or limbic
structures. Frontolimbic connections are of key importance in emotional and
personality alterations [53]. The orbitofrontal regions are particularly related

Figure 3-3. CT and magnetic resonance imaging (MRI) are compared in a patient late after
severe traumatic brain injury. (A) CT scan shows dilatation of the left frontal horn of the lateral
ventricle and subtle adjacent left frontal low density. Subtle low density also involves the left
globus pallid us. (B) Comparable MRI section shows extensive increased signal in the left frontal
white matter as well as increased signal in the right frontal white matter.

48

I. Theory and Intervention

to the basolaterallimbic system, including anterior temporal structures [53]. In


contrast, it has been hypothesized that DAI may affect structures that are more
functionally related to the medial limbic system and dorsolateral frontal lobes
[58].
In a recent series of head-trauma patients who had sustained marked frontal
lobe parenchymal damage, but who had lost consciousness for less than one
hour, prominent impairments included amnesia, mental control difficulty,
superficial and labile affect, and behavioral outbursts. In contrast, findings
on a fundamental neurological examination and basic activities of daily living
(ADL) skills were generally preserved even in patients with moderate or severe
disability [22].
Anterior temporal lobe contusions are likely to be associated with amnesia
[17], although in some cases the lesions are lateral to the hippocampal region
[17, 21]. Seizures are also a particularly significant risk for patients with these
lllJunes.
The sequelae of left temporal injury are somewhat better known than the
sequelae of right temporal injury. A slow return of the verbal (V) score on the
Glasgow Coma Scale and impaired cognitive outcome have been correlated with left temporal electroencephalographic (EEG) abnormalities [59].
Wernicke's or anomic aphasia and a tendency to exhibit paranoid ideation
have been described in patients with temporal lobe hematomas that are more
posterior [60].
Basal ganglia hemorrhages are less frequent than cortical contusions. They
can be associated with characteristic deficits, including hemiparesis and mutism
[61]. Children who have had an initial hemiparesis have generally shown rapid
improvement and good outcome [62]. The outcome for adults with basal ganglia hemorrhages has generally been worse than for those with cortical hemorrhages, and a pathophysiological association between DAI and basal ganglia
hemorrhages has been described [63].
2.6 Herniation and Brainstent Involventent

The amount and location of parenchymal damage are the most significant anatomical contributors to the late sequelae of brain injury. Subdural hematomas
are often accompanied by underlying parenchymal contusions, whereas epidural hematomas typically are not. However, in both of these extra cerebral
hematomas, the secondary damage to other brain structures (from increased
ICP and herniation) primarily determines outcome.
Herniation can be central or lateralized. In central herniation, diffuse downward pressures are exerted on brainstem structures. In lateralized (uncal) herniation, mesial temporal structures-including the amygdala and frequently
the hippocampal region-are displaced through the tentorial aperture. In uncal
herniation, third-nerve palsy is typically followed by brainstem compression
and coma. With lateralized herniation, it is not uncommon to see ipsilateral
hemiparesis (i. e., on the same side of the patient's body as the lesion). This fre-

3. Issues in Behavioral Neurology

49

quently represents the so-called "Kernohan's notch" phenomenon, in which a


mass lesion on one side pushes the deeper structures towards the opposite side,
where corticospinal tract fibers in the opposite cerebral peduncle of the midbrain are compressed against the free tentorial margin of the dura [64-66]. At a
more advanced stage oflateralized herniation, ipsilateral decerebration and contralateral decortication have been described, especially with epidural hematoma.
In addition to direct compressive effects on brain tissue, deformation of the
posterior cerebral artery (PCA) branches and the anterior choroidal artery
(AChA) during herniation can cause infarction [65, 66]. Unilateral infarction
of PCA territory causes a contralateral field cut. Bilateral involvement can result in cortical blindness with denial of blindness (Anton's syndrome) [11]. Left
occipital infarction can be associated with alexia without agraphia if fibers crossing from the right hemisphere through the corpus callosum are also affected.
(Recent findings show that these white matter fibers need not specifically be
affected in the corpus callosum per se, but that involvement of the adjacent
forceps major or para ventricular white matter can be sufficient to cause the
alexia [67].) Bilateral PCA infarction may result in severe persistent amnesia,
whereas amnesia from unilateral infarction is more often temporary [68]. Of
course, in brain injury, the combination of unilateral PCA infarction with other
pathological conditions in the opposite hemisphere may also result in more persistent amnesia. AChA infarction most frequently causes hemiparesis through
internal capsular involvement [69, 70].
Stretching of brain stem-penetrating vessels can cause brainstem hemorrhages
and infarcts. The location of brain stem involvement in herniation differs from
that in DAI [71], but its occurrence in either instance impacts significantly on
functional outcome [72, 73]. Lesions involving the periaqueductal gray matter
and the floor of the third ventricle have been correlated with lasting unresponsiveness [74, 75]. Brainstem involvement relates to the motor impairments,
dysarthria, limb ataxia, ocular motor abnormalities, and arousal/attentional
difficulties that are seen in brain-injured patients.
2.7 Hydrocephalus

Hydrocephalus is always a consideration in patients who do not progress adequately after brain injury. Hydrocephalus can be associated with increased ICP
and may necessitate shunting. Alternatively, normal pressure hydrocephalus
(NPH) can be extremely difficult to distinguish from ex vacuo ventricular dilatation (central atrophy). Table 3-1 shows features helpful in determining if
dynamic hydrocephalus is present [76-87]. Even if true NPH is present, sometimes fluid dynamics can equilibrate, and shunting may not be necessary.
Of the classic clinical triad of dementia, urinary incontinence, and gait difficulty, the last is probably most important [76, 77] for diagnosing hydrocephalic
complications. For example, a slow, small-stepping, "magnetic," "frontal"
gait with imbalance and falls, rather than cerebellar ataxia, is characteristic of
NPH. Akinetic mutism, apathy, and attentional or memory difficulties may

Opening pressure high or normal


Positive therapeutic response to fluid removal
Ventricular pooling
Decreased

Present
Increased

Decreased
Early 2 mos)
Progressive ventricular enlargement
Present (trans ependymal fluid)

Hemorrhage-especially IVH or SAH


Dementia, gait disturbance, urinary incontinence
Bulging
Worsening, intermittent or static course

IVH = intraventricular hemorrhage; SAH = subarachnoid hemorrhage.


Sources: [76-871.

Cisternography
CSF outflow conductance

Lumbar puncture

INVASIVE PROCEDURES

Ventricular rimming
Aqueductal flow void

MRI SCANS

Sulci, fissures, cisterns


Ventricular enlargement
Serial scans
Peri ventricular lucency

CT SCANS

Antecedent history
Clinical deficits
Skull flap (when present)
Clinical course

CLINICAL FEATURES

Dynamic hydrocephalus

Table 3-1. Factors helpful in identifying dynamic hydrocephalus

Progression to convexity
Normal

Opening pressure normal

Absent
Decreased

Prominent
Late (>2 mos)
Stable ventricular size
Absent (or frontal contusions)

Diffuse axonal injury or anoxia


Focal or c/w injury severity
Depressed
Chronic slow improvement

Cerebral atrophy or ex vacuo dilatation

::s

S"

0-

::s

.,

o
-.';l

(b

;l

til

<:>

3. Issues in Behavioral Neurology

51

be prominent. In addition, urinary incontinence rna y be associated with urgency


or with lack of interest and initiation. Clinical worsening and progressive ventricular enlargement seen on serial CT scans are diagnostically useful. Periventricular lucency on CT scans [78] or ventricular rimming on MRI scans [79,
80] can represent trans ependymal fluid and can be an important clue. Further
tests in equivocal cases can include lumbar puncture [77, 81], cisternography
[77], ICP monitoring, and measures of outflow conductance (which are the
most reliable, but require placement of an intraventricular catheter) [82, 83].
In some instances, shunting may represent a last hope for noticeable improvement in a patient. However, we must bear in mind that most brain-injured
patients with true hydrocephalus usually have only a limited response to shunting, since other pathological change is usually present. Complications of shunting include infection, shunt malfunctioning, seizures, and subdural hematoma.
3 POSTTRAUMATIC EPILEPSY

3.1 Risks

Posttraumatic seizures are generally divided into "early" (within the first week
postinjury) and "late" varieties. Early seizures often occur in a setting of increased ICP and do not have the same implications for ongoing epilepsy that
later seizures have. Nonetheless, late epilepsy occurs in about one third of adult
patients with early seizures [88].
The risk of late seizures varies widely and depends on the nature of the injury. Missile injuries have consistently been shown to be associated with a late
seizure risk of over 30% and with a risk of over 40% when dural penetration
has occurred [89]. Although the overall incidence of seizures for patients admitted to a hospital after closed-head injury is approximately 5%, the risk is
clearly much higher in certain patient subgroups. Depressed skull fractures,
intracranial hematomas, and early posttraumatic seizures all markedly raise the
risk oflate seizures [88, 90]. Combinations of these factors with each other or
with a postraumatic amnesia (PTA) of greater than 24 hours confer greater risk
than does a single factor alone [88]. For instance, in one series, the risk oflate
seizures after intracranial hematoma rose from 26% to 44% when PTA exceeded 24 hours [88]. Although PTA of greater than 24 hours by itself does not
appear to significantly increase seizure risk, coma of over three weeks' duration has been reported to do so in the absence of intracranial hematoma or early
posttraumatic seizure [91]. In general, focal injury in combination with severe
diffuse injury appears to be worse than focal injury alone [92].
Although a first posttraumatic seizure may occur many years after an injury,
at about five years postinjury the incidence of such seizures in head-trauma
patients approaches that of the general population. Over 50% of first late
seizures occur within the first year and between 70% to 80% within the first
two years [93]. Methods exist for calculating the residual risk of seizures, given
the four worst injury factors and the length of time postinjury [94, 95].

52

I. Theory and Intervention

In addition to considering the risk of a first late seizure, we need to consider


the relative likelihood of the occurrence oflasting epilepsy. Cessation of fits is
observed in approximately 50% of those patients with late seizures [92]. The
most important factor in persistence of seizures appears to be the frequency of
attacks before control is achieved. Increasing seizure frequency and focal attacks
indicate a greater likelihood of recurrence [92].
The disappointing unreliability of the EEG for predicting late seizures has been
generally acknowledged [88]. In particular, a normal EEG does not provide any
security that seizures will not subsequently develop. An irritative abnormality
on a late EEG, however, might discourage discontinuation of anticonvulsant
therapy in some clinical contexts.
3.2 Kindling

The mechanism of epileptogenesis after brain trauma remains incompletely


understood. Relevant factors may include iron deposition secondary to hemorrhage, biochemical alterations, glial reaction, and meningocerebral cicatrix
[96]. Because the kindling model in experimental animals may have significant
implications for seizure prophylaxis and behavioral alterations in epileptic patients, it will be reviewed briefly. Kindling is a phenomenon whereby repeated
brain electrical stimulation induces a lasting propensity to spontaneous seizures
[97, 98]. This occurs despite the fact that stimuli are at levels below the seizure
threshold. Thus it is not that the stimuli directly propagate a seizure, but rather
that repeated stimuli cause a lasting alteration in neuronal firing patterns. The
limbic structures most often kindled are the amygdala, surrounding pyriform
lobe, and hippocampus.
3.3 Anticonvulsant Prophylaxis

There is a substantial risk of developing epilepsy after severe head injury. Unfortunately, there is little convincing evidence for decreasing this risk by anticonvulsant prophylaxis. Studies of posttraumatic epilepsy in animals, as well
as early uncontrolled studies in humans, support the view that anticonvulsants
have been beneficial [99]. However, three controlled double-blind studies have
failed to show any demonstrable decreases in seizure development in patients
treated prophylactically with phenytoin or combined phenytoin and phenobarbital [100-102].
Unfortunately, it is difficult to interpret the results of these studies because
of problems with patient compliance, drug dosage, the failure to maintain therapeutic blood levels, insufficient sample size, and lower-than-anticipated seizure
frequency in controls. In one study, for example, all patients who had a first
seizure had a blood phenytoin level less than 12 [101]. In another study, blood
levels were not measured, and the standardized drug dosages used may not
have produced therapeutic levels in many cases [100]. These results have raised
speculation as to whether prophylaxis might be more effective if drug levels
were consistently maintained in the high therapeutic range.

3. Issues in Behavioral Neurology

53

Some anticonvulsants carry the risk of side effects such as rashes, liver abnormalities, and blood dyscrasias. For this reason, a seizure risk in the range of
15% to 20% has been used as threshold for treatment [88, 93, 99]. In addition,
phenobarbital has well-known sedative effects, and phenytoin has recently been
shown to have adverse effects on cognitive function [103]; these side effects are
not trivial for brain-injury patients-many of whose major deficits are in the
cognitive realm-and also should be taken into account when considering
prophylaxis.
Interestingly, carbamazepine has been associated with less cognitive impairment than phenobarbital and phenytoin [103]. Only limited information is currently available regarding its use prophylactically [104, 105]. However, in one
randomized prospective study, head-injury patients at high risk for seizures
received prophylaxis with carbamazepine or placebo for 11/2 to 2 years postinjury. The carbamazepine group showed a lower total incidence of posttraumatic seizures, although this finding was not statistically significant for late
posttraumatic seizures per se [105]. Other advantages of carbamazepine include efficacy for both generalized and complex partial seizures, a unique ability
to inhibit even localized kindled limbic seizure discharges [106], and beneficial
effects on mood and behavior in some patients [107]. A reduction of the white
blood cell count is not uncommon with carbamazepine, and serial blood monitoring should be performed, but the occurrence of true aplastic anemia is rare
(estimated at less than 1/50,000) [108].
In practice, decisions to institute anticonvulsant prophylaxis for patients
should be made on an individualized basis. The benefit of prophylaxis remains
unproven and the risk of adverse effects from anticonvulsants is significant.
However, it still appears reasonable to use prophylactic anticonvulsant drugs
for scIected patients in particularly high-risk categories.
The recommended duration of prophylactic treatment has ranged from six
months to two years [93, 99]. If the shorter duration of therapy is selected, it
should be with recognition that a significant seizure risk can still be present
despite multiple seizure-free months [94, 95].
3.4 Ictal Events

The seizure ictus can take many clinical forms and must be distinguished from
other episodic behaviors. The key to seizure diagnosis is an accurate account
(history) from an eyewitness. This should include descriptions of the preceding
activity, mode of onset, head or eye deviation, motor movements, incontinence,
duration, and postictal confusion. History from the patient should also include
a search for alteration in emotion, such as ictal fear, and for psychosensory
phenomena such as hallucinations or deja vu. Accuracy of diagnosis is important even for patients on prophylactic anticonvulsants because an eventual
decision about stopping these medications depends on whether seizures have
occurred. Alternative diagnoses for seizure-like events that frequently occur in

54

I. Theory and Intervention

brain-injured patients include clonus, palatal myoclonus, brainstem reflexes in


low-level patients, tremors, variable attention, syncope, conscious aggression
and acting out, and pseudoseizures.
Epilepsy is a clinical diagnosis; undue reliance should not be placed on the
EEG or other tests. Reenactment of triggering stimuli during EEG [109], prolonged EEG recording, or combined video-EEG recording in patients with
frequent episodes can sometimes clarify difficult cases.
3.5 Behavioral Alteration in Epilepsy

Discussions of epilepsy have focused on four main categories of behaviorpersonality, aggression (see Section 8), psychosis, and depression.
A specific personality alteration described in some patients with temporal lobe
epilepsy (TLE) has recently been called the Geschwind syndrome. Characteristic
manifestations of this personality change include circumstantiality (providing
excessive detail), "stickiness" (a difficulty in disengaging from the examiner),
hypo sexuality (decreased sexual interest/activity), hypergraphia (writing voluminously), and hyperreligiosity (change to new and extreme religious fervor)
[52, 110, 111]. There is a deepening of emotional intensity, in contrast to the
superficiality seen as a result of frontal lesions [112]. The temporal lobes are
considered a meeting place for sensory information and limbic emotional drives
[1]. It has been hypothesized that, even in the absence of frank seizure activity,
increased excitability of these circuits could lead to a state of sensorilimbic
hyperconnection and observed personality change [112].
At least three major controversies have surrounded the conceptualization of
the TLE personality. First, there are questions about validity of the symptom
profile, as distinct from other psychiatric disturbances [113, 114]. Supporters
have contended that a cluster of unrelated behaviors together with seizures is
consistent with the concept of a syndrome. The fact that not all patients with
TLE have the syndrome and that certain behaviors can occur in other clinical
contexts does not invalidate the concept [111].
Second, the specificity of the syndrome for TLE, as opposed to other forms
of epilepsy, has been questioned. Comparisons of behavior trait profiles have
revealed that some patients with generalized seizures have patterns similar to
those of patients with TLE. However, some patients with clinically generalized
seizures may have temporal lobe foci [111].
Finally, the basis for the behaviors has been questioned. Social factors, destructive brain lesions, and reactions to having epilepsy have all been cited as
potential factors in personality alteration. However, the nature of TLE behaviors is the opposite of the kind of behaviors that have been described in
patients who have sustained the destructive anterior temporal lesions that cause
the Kliiver-Bucy syndrome; this fact has been used in support of the limbic
hyperconnection hypothesis [111].
One of the difficulties in analyzing the "TLE personality" is that more than
one syndrome may exist; laterality may play an important role in the type of

3. Issues in Behavioral Neurology

55

syndrome expressed. For example, interictal hypergraphia has been associated


specifically with right temporal involvement [115]. It has been suggested that
ideational traits may be more related to left temporal foci and that hyperemotionality may be more related to right temporal foci [112].
The association of psychosis with epilepsy has been well-documented, even
though psychosis often occurs many years after seizure activity. At times, the
symptomatology may be indistinguishable from that seen in schizophrenia;
however, a better preservation of affect and social interaction in the schizophreniform psychosis ofTLE has been described [57, 116]. Here, too, laterality
may playa part. For example, left temporal lesions or foci have been said to be
more common than right temporal ones in patients with interictal psychosis
[117]. Differences in laterality may be more apparent on positron emission
tomography (PET) scans than on CT scans [118, 119]. However, a specific
syndrome of delayed psychosis following right temporal infarction has also
been described and is frequently associated with a seizure disorder [120].
Alternatively, it has been suggested that it is the presence of basal temporal
spike discharges detected with sphenoidal electrodes rather than the laterality
of discharge that is associated with psychosis [121].
Seizure frequency may not parallel the occurrence of psychosis. In fact,
an interesting phenomenon of "forced normalization" has been described
[117, 122-124]. In some patients, normalization of EEG patterns with the
disappearance of seizure discharges has been associated with development of
psychosis. Because of the inverse relationship with seizure control, this has
sometimes been termed an "alternative psychosis." The phenomenon has been
observed both in patients with primary generalized epilepsy and in those with
focal epilepsy (such as that seen after brain injury).
The mechanism of forced normalization is not known. However, restricted
subcortical seizure discharges have been recorded by depth electrodes in psychotic epileptic patients with normal surface EEGs. Thus, it may be that "alternative psychosis" results from different patterns of discharges, rather than
from the abolition of discharges [122]. Alternatively, it has been suggested
that controlling seizures with anticonvulsants may augment the mesolimbic
dopaminergic system, leading to psychosis [123, 124].
Depression in patients with epilepsy is common and is probably underdiagnosed [125]. In some patients, a reaction to diagnosis or dependency may be
paramount, whereas in others there appears to be an endogenous component.
Right-sided foci are sometimes said to predominate in such patients, although
there is less evidence for this than for the association of psychosis with leftsided foci [117]. Medication effects, such as the potential depressive properties
of phenobarbital or folate deficiency secondary to phenytoin, should also be
considered. Although antidepressants should be administered cautiously because of their potential risk of exacerbating seizures [126], benefits often outweigh risks in depressed patients, and antidepressants are frequently tolerated
without any worsening of seizure control [127].

56

1. Theory and Intervention

4 DISORDERS OF AROUSAL, ATTENTION, AND MENTAL CONTROL

4.1 Emerging Coma

Coma has been defined as "a state of unarousable psychologic unresponsiveness


in which the subjects lie with eyes closed" [64]. Of patients who remain unresponsive for over two weeks, three fourths exhibit spontaneous eye-opening
within one month postinjury, even if other signs of conscious awareness have
not returned [128]. These patients are no longer technically comatose, and
terminology has been a problem in describing the subsequent clinical conditions encountered in these patients [128-132]. Some of the most commonly
used terms include persistent vegetative state [131], apallic syndrome [130], and
akinetic mutism [64, 129, 130].
Although the term vegetative state was originally coined to describe patients
in whom vegetative or autonomic functions were preserved, it has been unable
to escape its pejorative connotations. Moreover, many emerging coma patients
are able to perform at above true vegetative-state criteria at least on an inconsistent basis. It is now generally agreed that the term persistent vegetative state
(PVS) should not be used earlier than one year postinjury [132].
The term apallic syndrome derives from the Latin pallium, meaning cloak or
mantle. It was originally used to describe a loss of cerebral cortical function
[133]. Some have restricted use of the term to cases with actual destruction of
the cerebral mantle, such as in severe anoxia [48]. On the other hand, the term
has found favor with German authors for describing a clinical syndrome following prolonged posttraumatic coma, even when the brainstem is affected and
the cortex is not destroyed [130]. In the "full stage" of the apallic syndrome,
patients show no sign of conscious awareness despite open eyes. Associated
behaviors and examination findings have been described in detail [134].
The term akinetic mutism has been used to describe a state of wakeful inactivity. Preservation of visual tracking, a general appearance of alertness, and
motor inactivity not attributable to paralysis are characteristic of patients with
bilateral basal-medial frontal lesions or hydrocephalus [64]. A second somnolent form of akinetic mutism with impaired extraocular movement has been
associated with meso diencephalic involvement [135]. One limitation of the
term akinetic mutism is that some noncommunicative patients are restless rather
than akinetic.
Unfortunately, the lack of consistent terminology has often led to confusion
about prognostic implicatons. Analysis of the patients' capabilities in specific
areas is probably more useful than an ambiguous descriptive term. The Bryn
Mawr Coma Emergence Scale [136] has been proposed as a means of documenting the emerging coma patients' functioning with regard to basic communication aspects (see Table 3-2).
The lengths of time until patients' eyes open and until the patients can follow
commands have both been highly significant in predicting outcome from prolonged unresponsiveness. Failure to exhibit eye-opening by one month postinjury is a poor prognostic sign. In contrast, as many as one third of patients

3. Issues in Behavioral Neurology

57

Table 3-2. Coma Emergence Scale


Team 1

Date

Team 2

Team 3

Goal

Revised goal

1. Trackilll
1 = frequent or always
o = occasional or less

2. COll1l11allds
3 = consistent daily
2 = frequently-good day
1 = occasional
o = rare or never
3. Yes/No
2 = reliable
1 = sometimes
o = unreliable or none

4. Speech
3 = conversational
2 = frequent word/phrase
1 = occasional words
o = sounds, none
5. Main COlllllllll1icatioll
2 = speaks, writes, device/board
1 = movements-yes/no
o = no consistent means

Totals:
Bryn Mawr Rehabilitation Hospital Brain Injury Program
Coma Emergence Scale

who do not follow commands at three months postinjury are reported to have
at least a moderate outcome at one year [128]. Some studies on outcome from
the "vegetative state" have been much less optimistic, but the use of prevalence
rather than incidence and the failure to separate outcomes for different causes
have limited the applicability of these studies [137, 138]. Age and motor pattern do appear to interact with duration of unresponsiveness in determining
outcome [139]. For instance, prognostic expectations would be much more
modest for older patients or for those with posturing, pupillary abnormalities,
or oculomotor impairments.
The quality of outcome in patients who emerge from prolonged unresponsiveness is dependent on the type of lesion present. Of patients who were admitted to a rehabilitation hospital when they were still at a Level II or Level III
on the Rancho Los Amigos scale, those with predominantly diffuse injury were
statistically more likely to attain a discharge Rancho level of VI or VII, a discharge Barthel score over 65 (a measure of ADL independence), and at least a
moderate outcome at one year postinjury than were patients with major focal
damage [140]. Although comparable numbers of patients in both groups pro-

58

I. Theory and Intervention

gressed beyond Rancho III, patients with focal damage tended to remain at
Rancho V. The worse cognitive outcome in the focal group presumably is due
to the superimposed cortical damage [140].
The "locked-in" syndrome with preserved conscious awareness but motor
paralysis sparing only eyeblinks and vertical eye movement is well-known.
This locked-in syndrome was originally described in patients with pontine
infarction; a similar syndrome has also been described in cases of midbrain involvement [64, 141]. Current computer technology can offer sophisticated
communication capabilities to some of these patients. Nonetheless, demands
on effort, motivation, and patience remain considerable.
A different but similar group of patients is also encountered after brain injury. These patients typically are unable to follow commands consistently for
at least several months. When they do begin to follow commands consistently,
their responses may be restricted to eyeblinks or limited to movement of one
limb. Gradual motor improvement can be seen, but patients remain dependent
for daily care. Communication switches and devices are sometimes helpful for
these patients, but residual cognitive deficits are often a limiting factor. These
patients, then, exhibit a "partially attentive, partially locked-in" syndrome.
Pathophysiologically, patients in these cases have frequently demonstrated
either subcortical intracerebral hemorrhage or extracerebral hemorrhage with
herniation (D. Long, unpublished data). The recent description of a locked-in
syndrome as a result of ipsilateral capsular involvement and contralateral compression of the cerebral peduncle may be relevant to the motor deficits in this
patient group [142].
4.2 The Confusional State and the Mental Control Continuum

After emerging from unresponsiveness, patients with severe traumatic brain


injury characteristically enter an acute confusional state. The primary disorder
in the acute confusional state is a disturbance of attention. In its full-blown
form, there is an inability to maintain a coherent line of thought. Related problems often include distractibility, impulsivity, perseveration, confabulation,
denial of illness, and altered sleep/wake cycles [17]. When decreased attention
is combined with increased arousal, the term delirium is sometimes used.
Hallucinations, motor restlessness, and aggressive behavior can occur.
Maintaining a supportive structured environment without extraneous stimuli is the cornerstone of the management approach. Efforts should be made
to treat patients' pain, address their fears, and restore appropriate sleep habits.
As patients improve, these abnormalities become less marked, and patients
progress along a continuum of improving mental control functions.
4.3 Unilateral Neglect

Unilateral neglect is a disturbance of directed attention in which one side of


space tends to be ignored [143]. Unlike the acute confusional state, which can
occur with diffuse injury, neglect usually reflects focal damage. Because a

3. Issues in Behavioral Neurology

59

neuronal network for directed attention exists, causative lesions sometimes can
involve deeper structures such as the thalamus or the mesencephalic reticular
formation rather than the parietal cortex [143, 144]. A predominantly motor
form of neglect has been described, in which patients tend not to use one side
of their bodies despite relatively good strength [144, 145]. Neglect usually decreases over time, but it can cause lasting problems for some people. One report
has suggested that bromocriptine may be helpful for patients with persistent
left neglect [146].
4.4 Perseveration

Perseveration means the persistence or recurrence of an experience or activity


without the appropriate stimulus [147]. Although perseveration can be an
extremely dramatic symptom (see Figure 3-4), it can also be one of the most
sensitive early indicators of brain pathology. It frequently occurs after head
injury but is not always recognized.
Several types of perseveration have been identified [147 -149]. In clonic (continuous or efferent motor) perseveration, a motor activity-once initiatedremains persistent. For example, a patient who is asked to draw a single circle
may continue to draw without stopping. By contrast, in cases of intentional
perseveration (inertia of the motor program), a behavioral response occurs that
is more appropriate to a preceding stimulus than to the one that elicits it. For

Figure 3-4. An attempt to draw and number a clock to verbal command rcsults in extcnsive
perseveration. This includes both perseveration of the task (numbering) and perseveration of
individual numbers within the task.

60

I. Theory and Intervention

instance, following correct completion of a command to draw a circle, a patient


who is subsequently asked to draw a square might again draw a circle. Contrary
to what the name might be construed to mean, intentional perseveration does
not represent a willful error, but rather describes an involuntary disturbance of
motor activity. Intentional perseveration can occur on immediately sequential
tasks or after a delay with intervening stimuli [150]. Although both clonic and
intentional perseveration can occur with frontal lobe pathology, clonic perseveration has been associated with a greater degree of subcortical or right
hemisphere involvement [148]. Intentional perseveration can be further subdivided into a "stuck-in-set" variety, seen with frontal system abnormalities,
and a "recurrent" form, seen in patients with aphasia [149].
In some patients, perseveration is pervasive across all activities. In others,
perseveration is limited to an area of major clinical deficit [147]. For instance,
in aphasics, perseveration is often limited to language tasks. A double dissocia-

Figure 3-5. The patient was sequentially asked to draw a clock and then a house. The circular
outline of the clock is perseverated onto the honse, despite inclusion of other house features. Such
combinations of elements of sequential tasks are not well explained by an "attentional shutter"
perseveration model, but rather suggest a higher integrative mechanism. It is noteworthy in this
example that the patient rotated the paper while drawing the clock, and it was specifically this
circular or rotatory quality that was perseverated.

3. Issues in Behavioral Neurology

61

tion between perseveration on memory tasks versus the Wisconsin card sort
has also been described [151].
The mechanism of perseveration has been the subject of considerable discussion. The frequent occurrence of perseveration in patients who have disorders of attention suggests that in some cases there may be a failure to process
or attend to the stimulus evoking the perseverative response. However, this
"closing of the attentional shutter" fails to explain other frequently observed
phenomena. For example, perseverative responses can simultaneously include
features relevant to both present and preceding stimuli, implying a higher level
integrative deficit (see Figure 3-5) [152]. A failure to inhibit preceding
memory traces has been hypothesized. A related phenomenon-in which a
response appropriate to an earlier stimulus is first produced to a later (inappropriate) stimulus-is consistent with this model. Physiologically, the concept
of recurrent firing patterns in a given neuronal network may be relevant [153].
In some patients who are aware of the incorrectness of their responses, but
who are unable to inhibit them, a more explicit motor mechanism may be
involved.
At a practical level, multiple factors impact upon the likelihood of obtaining
a perseverative response. Perseveration is believed to increase with fatigue and
with task difficulty [154, 155]. Perseveration can be decreased by working
with patients when they are fresh and are given tasks that they are able to
accomplish. Leaving ample time between stimuli and minimizing relatedness
of the stimuli can also be beneficial. Patients with initially severe perseveration
can often show dramatically improved overall performance when these techniques are used, or after perseveration clears. Specific therapy of aphasic perseveration has recently been utilized as a means of improving language functioning

[156].
4.5 Complex Aspects of Attention and Mental Control

When patients are in a florid confusional state, their attentional impairments


are obvious. However, more complex attention-related deficits are often persistent at later stages of recovery and are more likely to be overlooked. Thus,
many problems attributed to memory or impaired problem-solving may be
related to impaired concentration, failure to pay attention, or disorders of
intention (the active state of readiness for a particular action) [55].
Imitation behavior [157], utilization behavior [158], and the environmental
dependency syndrome [159] are a group of disorders associated with pathological changes of the frontal lobe in which patients are unable to independently
direct a course of action. In utilization behavior, patients characteristically feel
compelled to use items presented to them, even though they have not received
any verbal instruction to do so, and even when they have received explicit
verbal instruction not to do so. Item use includes such activities as pouring
water from a pitcher, lighting a cigarette, or peeling an apple. Imitative behavior
is a milder form, limited to copying the examiner's gestures. In the environ-

62

I. Theory and Intervention

mental dependency syndrome, certain complex behaviors can be induced by


placing the patient in unique environmental situations and providing predominantly nonverbal cues.
These three deficits are interpreted as obligatory responses to multimodal
sensory information mediated by the parietal lobe in the absence of appropriate
frontal lobe control. The importance of a structured environment for patients
with frontal-system dysfunction is well known, but these syndromes emphasize the overriding role of nonverbal and inherent environmental information
over verbal instruction in this patient group.
4.6 Akinesia, Apathy, and Drive

A paucity of spontaneous activity is common after brain injury. More severe


disorders of drive form a continuum with akinetic mutism (see Section 4.1),
in which responses may be obtained only with maximal provocation. This
group of disorders may be related to mesial frontal dopamine pathways and is
sometimes responsive to dopaminergic agonists such as bromocriptine [160].
Apathy and other impairments of drive can be considered fundamental
features of the frontal syndrome [53]. Unfortunately, they are difficult to treat.
Dissociation of verbal statements and actions can be particularly frustrating for
families or staff. However, when patients enter a phase of decreased participation in their therapy programs, aspects of their adjustment to disability or
depression can be superimposed on underlying anatomically determined
apathy. Even though a true diagnosis of depression may be difficult to establish in these patients, anecdotal experience suggests that low doses of tricyclic
medication sometimes may be helpful.
4.7 Speed and Processing

Many patients with severe diffuse injury exhibit a marked slowness in performing all tasks, which is often a major limitation in their daily activities. In patients with milder diffuse injuries, subtle deficits in speed and processing are
noteworthy. Impairments in complex reaction time [161] and in the paced
auditory serial addition test (PASAT) [162] have been documented in patients
even after concussion; such information can be useful in vocational planning.
Because the P ASA T task is often misunderstood, a further clarification
seems appropriate. In this test, a series of numbers is presented at fixed-time
intervals. When a new number is presented, it must be added to only the preceding one. Note that this is distinctly different from making a total sum of
all numbers presented and requires a repeated shifting to the last preceding
number to which the new number is added. The nature of the errors often
reflects this heavy emphasis on task shifting, in which the patient must inhibit
the tendency to make a sum. Thus the task cannot be considered a pure test of
speed alone, although speed is an important parameter. Given unlimited time,
patients with milder injuries can perform this task, but not with steady, rapid
rates of presentation.

3. Issues in Behavioral Neurology

63

The physiological mechanism that leads to disorders of speed and processing


is not known. However, diffuse axonal damage can be considered analogous
to closed lanes on many city roads. Although cars can get through, the flow of
traffic is slowed-and one hesitates to contemplate rush hour.
5 DISORDERS OF VERBAL EXPRESSION

Disorders of verbal expression can be divided into disturbances of speech


production, language, and thought [163].
5.1 Speech Disorders

Mutism, a condition in which speech is not produced, is commonly seen after


brain injury. Mutism can be part of a global dementia or a widespread failure
to initiate motor activity, as in akinetic mutism (see Section 4.1). As a specific
focal disturbance, it is typically seen in certain clinical contexts.
After prolonged unresponsiveness, patients generally regain the ability to
follow commands before they regain the ability to speak [128]. This period of
mutism often evolves to a severe, slow dysarthria. A gradual improvement in
clarity so that speech is discernible enough for daily functional needs occasionally can take two to three years. Devices that augment communication can be
helpful for these patients, as long as the problem is truly mutism/dysarthria,
rather than aphasia.
Two types of pathological conditions have identified in patients with posttraumatic mutism [61]. One group has had basal ganglia hemorrhage. The
transient mutism, persistent slow dysarthria, and subcortical aphasic syndromes that have been described in infarctions in the region of the anterior
limb of the internal capsule and in the basal ganglia may be relevant to this
patient group.
A second group of patients has exhibited the combination of severe diffuse
injury and focal left hemisphere injury. In nontraumatic conditions, transient
mutism can result from isolated left hemisphere involvement in such regions
as the inferior portions of the motor strip, Broca's area, or the supplementary
motor area [164]. Lasting mutism generally requires bilateral involvement.
Bilateral pyramidal tract involvement with pseudobulbar palsy is common
after brain injury and can also cause mutism. In addition, brainstem injury is a
major consideration in severely injured patients.
In assessing the mute patient, it is useful to note not only whether any speech
is produced but also whether any sounds are produced. Patients with focal
cortical or pyramidal tract involvement may exhibit preservation of emotionally induced sounds such as laughing or crying. It has been suggested that the
absence of these may indicate involvement of the limbic structures or connections [164]. Failure to cough or produce any sound may be due to local
laryngeal abnormality.
Many types of dysarthria are seen after brain injury. Severe slow dysarthria
with poor articulation is common and is often difficult to treat. However, per-

64 I. Theory and Intervention

sistent therapy efforts may lead to gradual gains. In addition, prosthodontic


evaluation can reveal whether a palatal prosthesis might help the patient.
Hypophonia (decreased voice loudness) and rapid festinating (hastening)
speech are characteristically associated with basal ganglia involvement.
Patients with rapid festinating speech may respond to the use of a pacing board
or delayed auditory feedback. The association of hoarseness or diplophonia
(simultaneous production of two pitches) with hypophonia usually implies
vocal cord impairment.
Stuttering and palilalia are frequently encountered. Stuttering includes prolongations, blocks, and repetition of phonemes; palilalia entails the patient's
repetition of his or her own syllables, words, or phrases. History from the
patient and family and observation of the stuttering pattern itself can help the
examiner determine whether the stuttering has been acquired or is congenital.
Congenital stuttering tends to occur on the initial phoneme of substantive
words and is often associated with grimacing, posturing, and frustration. In
contrast, acquired stuttering occurs on any phoneme of any word, without
excessive effort or associated motor patterns [165]. There is no generally effective medication known to alleviate stuttering, although anticonvulsant treatment for an associated seizure disorder in one patient resulted in the clearing of
the stuttering as well [166]. Transient stuttering is often seen with unilateral
left hemisphere lesions. Although persistent stuttering usually implies bilateral
involvement [165], lasting stuttering with unilateral involvement has been
reported after penetrating brain injuries [167]. Palilalia also occurs with bilateral lesions, most commonly those in the basal ganglia region. A trial of
dopaminergic agonists can be considered for the palilalic patient.
5.2 Aphasia
Aphasia is a loss of previously acquired language function due to brain damage
[168, 169]. The prognosis for recovery from aphasia is generally better for
patients whose condition resulted from trauma than for those whose aphasia
occurred after a stroke. This may be because brain-injury patients tend to be
younger than stroke victims, and their lesions are often hemorrhagic. Although
hemorrhages and surrounding swelling may cause a severe acute presentation,
residual parenchymal damage is often less than for an ischemic lesion that
causes similar acute findings. Of greatest importance is that the parts of the
brain most often affected differ markedly in trauma and stroke. Anomic aphasia
is the most common form of aphasia seen after trauma, and posttraumatic
transcortical aphasias are also frequently encountered. These aphasias are disturbances in which repetition is preserved and in which the pathology is outside
the immediate perisylvian region. In contrast, aphasic syndromes with repetition disturbances are commonly found in stroke patients. Nonfluent aphasia is
common after stroke, whereas fluent aphasias predominate after head injury;
these include anomic aphasia, Wernicke's aphasia, transcortical sensory
aphasia, and so~e varieties of subcortical aphasia. Table 3-3 reviews the clini-

V
V

Hypophonic dysarthria, hemiparesis,


spontaneous paraphasias, attentional
impairment
Characteristics vary with specifics oflesion
location

Not common
Common
Common
Most common

Echolalia
Decreased speech production with poor initiation
Echolalia
N onlocalizing

P
G
P
G

NF
NF
F
F

G
G
G
G

Common
Not common
Common
Not common

Frequency after
head injury

May preserve midline commands


Agrammatism
Semantic & neologistic paraphasias
Literal paraphasias

Special features

P
G
P
G

Comprehension

NF
NF
F
F

Fluency

P
P
P
P

Repetition

presentation and nature of injury often compatible, but complexity of injury often complicates localization.
poor; G = good; V = variable; NF = nonlluent; F = lIuent.

* Clinical

Basal ganglia-internal capsuleperi ventricular white matter

Cortical with repetition disturbance


Global
Broca's
Wernicke's
Conduction
Cortical without repetition disturbance
Isolation of the speech area
Transcortical motor
Transcortical sensory
Anomie
Subcortical
Thalamic

Aphasic type

Diagnostic criteria

Table 3-3. Clinical features of aphasias in head injury

8:

'<

IJQ

0"

...go

15
...
eo.

<:

O>

g-

tD

'"~

!-"

66

l. Theory and Intervention

cal features of the different aphasic syndromes and their occurrence after head
Injury.
Posttraumatic aphasias may fail to be recognized because of patients' preserved fluency and repetition. Incorrect responses to orientation questions or
to tests about recently presented information may be incorrectly interpreted as
indicating a memory disturbance rather than a more basic impairment of
naming or comprehension skills. Conversely, the combination of attentional
impairment and a speech output disturbance such as mutism or dysarthria may
be misconstrued as aphasia.
Strategies that can help ensure correct diagnosis include careful artention to
the .quality of spontaneous speech, formal testing of auditory comprehension
(including linguistic aspects), confrontation naming in specific categories and
for low-frequency items, and assessment of reading and writing. Probably the
most common error in aphasia examination is a failure to obtain a full enough
sample of spontaneous speech (H. Goodglass, personal communication). Paraphasic substitutions, word-finding pauses, paragrammatism, or an empty
circumlocutory style can often provide diagnostic clues in patients with fluent
speech. Recounting information that relies on specific vocabulary, such as the
steps involved in changing a tire, can often dramatically demonstrate spontaneous word-finding difficulty.
Comprehension should be tested by both yeslno questions and serial commands, since perseveration may contaminate the former and apraxia the latter.
Because reading comprehension and auditory comprehension can be dissociated, both should always be tested. Distinguishing transcortical sensory
aphasia from anomic aphasia with concomitant attentional impairment can be
problematic. Echolalia is particularly characteristic in true transcortical sensory
aphasia.
Prominent paraphasias may highlight a specific component of aphasia,
whereas disordered behavior beyond language function generally is indicative
of a more global confusion state. A specific disturbance of grammatical comprehension tasks that rely on prepositions or verb tenses is characteristically seen
in some aphasic patients. In contrast to patients with predominantly attentional
impairments, these patients may successfully perform multistep tasks or respond to questions by pointing to objects by complex description. For example,
some aphasic patients have more trouble with "Touch the pencil with the
pen," than with "Show me an electrical device used to communicate with
someone over a great distance" (N. Geschwind, personal communication).
Global aphasia often occurs after mass lesions have been surgically debrided
or drained, after hemispheric swelling, and after secondary infarction of extensive areas of the cortex. Midline commands such as "close your eyes" are
usually spared (preserved) in even profound global aphasias after unilateral
infarction. This is less often the case after head injury because some degree of
bilateral involvement is generally present.
Transcortical motor aphasia is characterized by sparse spontaneous speech,

3. Issues in Behavioral Neurology

67

but repetition is well preserved. The output disturbance is usually one of decreased speech initiation and production, rather than articulatory difficulty.
Characteristic frontal lesions involve either the supplemental motor area or its
connections with Broca's area [170].
Subcortical aphasia has recently received increased recognition [169]. The
symptoms have varied considerably, depending on the exact location of the
lesion.
A series of patients with putaminal or thalamic hemorrhage had a characteristic presentation [171]. After the patients recovered from their initial
mutism, their speech was usually fluent but was hypophonic with underarticulated mumbling. Paraphasias were prominent in spontaneous speech, but
repetition was generally preserved. Comprehension was variable, and ideomotor praxis was spared. Attention and memory impairments were frequent.
Hemiparesis was present, and hemisensory loss was variable. Subsequent
analysis of patients with nonhemorrhagic thalamic infarction revealed a similar
pattern, including sparse but fluent speech with paraphasias, hypophonic dysarthria, preserved repetition, and impaired auditory comprehension, reading
comprehension, and writing [172].
Infarcts in the region of the anterior limb of the internal capsule and basal
ganglia have also produced lasting aphasias [173-175]. Dysprosodic dysarthric
speech production following initial mutism was characteristic in the patients
studied. Repetition was also impaired. Comprehension impairment was more
severe when there was posterior extension across the temporal isthmus. Small
differences in the location of peri ventricular white matter involvement on CT
scans correlated well with specific clinical components of the resultant aphasic
syndromes [175].
Anomie aphasia is the most common posttraumatic aphasia. In this syndrome, anomia is prominent, but fluency, repetition, and auditory comprehension are preserved. Although angular gyrus involvement is often implicated in
anomic aphasia (sometimes along with the Gerstmann syndrome of agraphia,
acalculia, finger agnosia, right-left disorientation), inferior temporal lobe involvement is probably more frequently relevant after head injury, in view of
the contusional patterns described earlier (M. Alexander, personal communication).
Anomia is common to all aphasic disorders and is generally considered a
nonlocalizing finding [168]. However, different types of anomia have been
identified; the differences are not only anatomical but also have significant
implications for management approaches. These types include word-production anomia, word-selection anomia, semantic anomia, category-specific
anomia, and modality-specific anomia [169, 176].
In word-production anomia, the patient appears to have some knowledge
of the word but is unable to produce it. For example, in the form of wordproduction anomia seen in Broca's or transcortical motor aphasia, patients
characteristically have difficulty initiating the target words and receive prompt

68

I. Theory and Intervention

benefit from phonemic cuing. A second form of word-production anomia is


frequently seen in patients with conduction aphasia, in which literal paraphasia
prevents production of the target word.
Word-selection anomia is a pure word-finding problem. Patients can offer a
description and can select the appropriate item from a group. In such cases,
cuing is usually oflittle benefit. Lesions in the posterior inferior temporal lobe
or temp oro-occipital junction have been identified in patients for whom this
deficit occurs as an isolated severe disturbance.
Semantic anomia is characteristically a two-way defect, in which symbolic
meaning of the word is lost. Patients have difficulty not only in producing the
name but also in pointing to named objects. This disorder can occur as part of
transcortical-sensory aphasia or Wernicke's aphasia. Occasionally, patients
will have more difficulty with name recognition after a paraphasic error has
occurred for that same name; this phenomenon has been called "word intoxication." For patients with this type of anomia, it may be therapeutically helpful to inhibit paraphasic vocalizations and to work on comprehension skills
prior to working on naming skills.
Category-specific and modality-specific anomias are usually attributed to
either involvement of primary sensory regions or to disconnection of these
regions from key language areas. Color-naming impairment associated with
left medial temporo-occipital lesions is probably the best known categoryspecific anomia [11]; other anomias have been described in which categories
such as body parts, numbers, letters, or geographic names are either specifically affected ("involved") or preserved ("spared"). An example of modalityspecific anomia would be a patient's inability to name an object when it is
presented visually, even though the patient can name the object when it is
verbally described or placed in his or her hand. The use of careful testing
techniques is crucial to the identification of these disorders. Because of the
specificity and limited nature of these deficits, rehabilitation efforts for these
patients can be particularly helpful.
Not all naming impairment is necessarily aphasic in nature. Nonaphasic
misnaming has been described in patients with diffuse brain involvement and
associated confusional states [177-179]. These patients not only may provide
an incorrect name, but also may propagate a whole series of related incorrect
labels. For instance, the hospital might be called a "prison," the nurse a
"guard," and the doctor the "warden." Sometimes, the incorrect name
suggests at least partial knowledge of the real name (for example, "Mt.
Cyanide Rest Home" for "Mt. Sinai Hospital") [177]. A tendency to make
many errors particuiarly in respect to hospital- and disease-related topics has
been described.
5.3 Agraphia and Alexia

Agraphia can be one of the most sensitive indicators of the acute confusional
state [180]. Disordered spelling and written omissions or perseveration are

3. Issues in Behavioral Neurology

69

sometimes considerably more severe than verbal output disturbance in patients


with primarily attentional impairment. Agraphia also frequently accompanies
aphasic disturbances. As an isolated occurrence, however, agraphia is not
common after brain injury.
By contrast, alexia without agraphia is more common, due to the frequency
ofleft PCA territory infarction secondary to trans tentorial herniation [67]. In
the early stages of rehabilitation, it may be beneficial to assist the patients in
tracing letter outlines with their fingers. Oral spelling is typically preserved.
Considerable recovery can occur, but both performance speed and the reading
of longer words are often problematic. The tendency to impulsiveness that is
so commonly seen in patients after brain injury can compound the alexic
person's problem of assuming the completion of a word or phrase rather than
actually reading it.
Care should be taken to distinguish true alexia from two varieties of pseudoalexia sometimes encountered after brain injury. Patients with left neglect
may ignore or misread the left side of words or sentences. Presentation of the
letters of the word or sentence vertically rather than horizontally can alleviate
the error. Other patients may have difficulty changing lines when reading.
This can result from excessive convergence drives following the pretectal
lesions that sometimes are seen after diffuse injury [181].
5.4 Disorders of Thought

Disorders relating to thought, but not explicitly involving basic language


function, can also be seen after brain injury. Problems with circumstantiality,
sequencing, and abstract thinking are particularly common. It is also important
to recognize that delusions and even hallucinations can have an organic basis
[182, 183]. Complex delusions are probably more closely related to subcortical
and limbic involvement, whereas simple delusions, including paranoia, may
reflect cortical involvement. Neuroleptic dopaminergic blockers can be helpful, especially for paranoia. Paranoia is particularly likely to develop in patients
who have fluent posterior aphasia, a condition that involves difficulty with
comprehension and unawareness of disability [60]. Hallucinations are not
always spontaneously reported by patients, and an increased yield will be
obtained by specifically asking about them. Both the characteristics of the
hallucinations (modality, field laterality, formed or unformed, movement,
duration, relation to prior sensory inputs, etc.) and the patients' reaction to
them vary significantly, depending on the location of related brain lesions
[183].
6 AMNESIA AND RELATED DISORDERS

6.1 Posttraumatic Amnesia

Posttraumatic amnesia (PTA) refers to the period of time after injury when the
patient has not yet resumed consistent day-to-day memory for ongoing events
[29]. It was originally conceptualized as a period during which full conscious-

70

I. Theory and Intervention

ness had not returned. Its duration is a useful indicator of both injury severity
and the patient's prognosis. For example, aPT A of a month indicates a much
more severe injury than a PTA of a day or a week. Originally, measurement
was retrospective and included the period of coma. More recently, techniques
for prospectively measuring the duration of PTA have been advocated [184,
185], and some investigators have not included the period of unresponsiveness
in PTA [28]. Retrospectively assessed PT As may be longer than prospectively
measured ones because patients may later fail to recall periods of time when
they appeared oriented and generally appropriate.
In some respects, PTA is a misnomer, because patients are frequently in a
confusional state (with impairments of arousal and attention) rather than
exhibiting a true specific syndrome of amnesia [21]. Islands of preserved
memory are characteristic of patients during the period of PTA and are
consistent with a state of variable attention.
Although PTA duration (including coma) averages three to four times coma
duration, considerable variability in this ratio has been described [29]. The
ratio appears to be much more consistent for diffuse injuries than for focal
injuries. This finding supports the use ofPT A duration as a measure of severity
of concussion or immediate impact damage. Patients with anoxia may pose an
exception because anoxia may cause prolonged memory impairment even
after only brief unconsciousness.
Patients with focal injury frequently show marked dissociation of PTA,
coma duration, and injury severity. Some patients with penetrating injuries
show no significant loss of consciousness (LOC) or PTA despite significant
focal damage [186]. Left hemisphere involvement has been associated with
LOC duration in these injuries.
In a study of 34 patients with nonmissile injuries, who were divided into
focal, diffuse, and mixed groups, patients in the "focal frontal group" were
found to have had prolonged PTA (over one month), despite LOC ofless than
one hour [22]. In another study, all but one of the patients who had PTA (not
counting coma period) of over 15 days, but coma of 10 days or less, had focal
mass lesions [28]. The trend to non correlation of PTA and LOC was reported
as particularly notable with left hemisphere lesions [28]. Penetrating lesions,
including those involving basal brain regions, have also produced lasting
amnesia without significant LOC [187, 188].
Older patients have been reported to exhibit relatively longer PTA for the
same duration ofLOC [189]. It is not known whether the increased frequency
of focal hemorrhages in these older patients contributed to this finding.
6.2 Retrograde Amnesia

Retrograde amnesia (RA) refers to a condition in which the patient does not remember the period of time immediately preceding the brain injury. In most
cases, RA is much shorter than PTA. The length ofRA often shrinks dramatically as the patient emerges from PTA [190]. The length of RA is often not

3. Issues in Behavioral Neurology

71

absolute, which has been demonstrated by extensive studies of the temporal


gradient of recall in patients whose RA was due to other conditions besides
head injury [191]. Better recall of remote events, rather than recent events, has
often been described. Short RAs in patients with anterior temporal/hippocampal lesions and marked retrograde deficits in patients with diencephalic,
midbrain, or posterior temporal lesions have been described [191-193]. However, the exact pathological basis for RA is unknown and is the subject of
ongoing research.
6.3 Anterograde Amnesia

Anterograde amnesia is the term for a specific syndrome in which an impairment


of memory occurs in the absence of more generalized attention or language
impairments [194]. Because brain-injury patients frequently exhibit these more
basic abnormalities, they may not qualify for the specific designation of
"amnesia. "
Other cognitive impairments may often give the superficial appearance of
amnesia, but improved performance with structuring of the request may indicate that the primary disturbance is not that of classic amnesia. For instance, a
patient may initially fail to recall the date but may correctly provide this information when specifically asked the month, day, and date individually. Other
patients may exhibit the phenomenon of "forgetting to remember." When
these patients come home from the store without the milk, they may be cued
to their error by family or environment. This is in contrast to amnesic patients
who reply, "What milk?"
A dissociation of verbal and visuospatial memory can be seen particularly
with unilateral temporal involvement. This has provided an impetus for such
techniques as imagery mnemonics. Although these approaches can be useful
for some patients, problems with generalization and the degree of mental
effort required have often been limiting in the head-injury population [195].
The importance of proactive interference and perseveration in amnesia has
been demonstrated. Although whether these deficits are integral to the amnesia
or an associated finding has been debated [196, 197], there is no doubt of the
importance of the clinical phenomena in brain-injury patients. Thus, context,
sequence of information, and cuing should be carefully considered when
planning rehabilitation approaches.
Perhaps the most exciting implications for rehabilitation come from the fact
that some aspects of learning in amnesic patients seem to be relatively preserved [198]. These patients can learn motor skills or even perform perceptual
tasks, despite profound memory impairment. In fact, patients may retain skills
or information despite not remembering ever having been taught them. This
phenomenon, sometimes called source amnesia, suggests that patients in a
rehabilitation program can benefit from therapy techniques and training even
when they are unable to remember the therapy sessions.

72

I. Theory and Intervention

"Priming" techniques that have led to word recognition with decreasing


letter cues have recently been described [198]; this type of approach may be
useful for learning information about an isolated domain. In this model, there
is no attempt to effect a generalized improvement in memory function, but
rather to teach skills relevant to use of memory aids (such as a computer) or to
specific vocational goals. If a similar dissociation between emotional learning and recall of specific events could be demonstrated, it would lead to a
heightened awareness of the importance of the emotional environment and the
effectiveness of behavioral strategies for even severely amnesic patients.
6.4 Confabulation

Confabulation represents the recounting and fabrication of false information.


Many patients conf.abulate only to fill information gaps in response to specific
questions, but others may exhibit a phenomenon of spontaneous, and often
fantastic, confabulation [199-200]. Such patients may vividly recount imagined
stories of journeys abroad or sexual exploits. Impulsivity, misuse of environmental cues, failure to self-monitor or inhibit incorrect responses, and unconcern regarding behavior are characteristics of this patient group. This type
of confabulation appears closely related to perseveration and "set rigidity"
[199], which are associated with frontal lobe injury, and in fact frontal lobe
pathology has been demonstrated by CT scan in patients who confabulate
spontaneously [199].
6.5 Reduplicative Para-Amnesia

In reduplicative para-amnesia, a distortion of memory most commonly results in


a reduplication of place. Patients who exhibit this phenomenon may correctly
name the hospital but mislocate it at a different geographic site that was significant earlier in their lives [201]. Thus they might claim to be in a hospital with
the same name, but one that is located in their home town. Such beliefs are
persistent despite evidence to the contrary, although the condition is usually
temporary with a gradual adjustment to the actual location. The problem
appears related to an inability to synthesize conflicting information. Combined
frontal and right hemisphere pathology has been described in such patients
[201 ].
Reduplicative phenomena including place, time, person, and body parts
have also been noted in association with denial of illness and confusional features
[202]. One additional specific reduplicative phenomenon deserving mention is
the Capgras syndrome [203]. Patients who exhibit this syndrome believe that
someone close to them (most commonly a family member) has been replaced
by an impostor. For instance, patients may claim that a second spouse or other
family member-who resembles, but is different from, the first-has been
substituted. Originally reported in schizophrenia, the syndrome has now been
seen after head injury [203].

3. Issues in Behavioral Neurology

73

7 EMOTIONAL DISTURBANCES

The use of "mood" to refer to internal subjective feeling state and "affect" to
refer to outward motor emotional expression can be a useful distinction when
one is dealing with brain-injured patients [53], some of whom exhibit a marked
dissociation of these aspects.
7.1 Pathological Laughing and Crying
Two varieties of pathological laughing and crying have been described [204,
205]; in some of these patients, the motor expression of laughing or crying is
totally dissociated from underlying mood. These patients may be unable to
control laughter even when they are not feeling amused; they may state that
the laughter actually hurts. Similarly, crying may occur spontaneously or in
response to attempted facial movements or nonspecific stimuli. These responses
tend to occur in a stereotypic all-or-none manner rather than as a graded response. Bilateral involvement of descending motor traits is usually seen, but
the syndrome can occur with unilateral lesions that involve either the anterior
limb or genu of the internal capsule and adjacent subcortical structures. The
disconnection of cortical or subcortical structures from the motor nuclei of
the pons and medulla has been the hypothesized explanation. Lesions within
the brain stem itself can also cause this syndrome [206]. Interestingly, there
have been no reports of this phenomenon having been produced by a unilateral
cortical lesion.
A second type of pathological laughing or crying has sometimes been
termed "emotional incontinence." These patients exhibit a lack of control over
the intensity of affect but do have some underlying emotional mood. For instance, patients may report feeling only slightly sad or unhappy even when
they outwardly exhibit uncontrollable sobbing. Alternatively, patients may be
unable to limit the severity of laughter even when they feel only slightly
amused.
Two medications have recently been reported to help pathologic laughing
or crying. Patients have shown a prompt response to amitriptyline at dosages
below that usually used for depression [207]; some have also been helped by
levodopa [208].
7.2 Aprosodia

The intonational pattern or melody of speech has been termed prosody [209]. In
English and other European languages, one of the most important roles of prosody is to convey emotion. In contrast to the well-recognized linguistic dominance of the left hemisphere, the right hemisphere appears to be dominant for
emotional prosody.
A group of disorders of prosody have been recognized and termed the aprosodias [210]. In some instances, patients' dramatic inability to impart emotion
to speech can have important functional consequences. For example, after she

74

I. Theory and Intervention

had sustained a right hemisphere lesion, a schoolteacher was unable to control


her class; this was attributed to a disturbance of prosodic production [211].
Alternatively, other patients show a prominent inability to recognize the
emotional tone of what is said to them [212, 213]. For instance, the enthusiastic
or angry tone of voice accompanying the statement "It cost ten dollars" might
ordinarily indicate either a bargain or an extravagant price. "He scored a goal"
could be saiCl jubilantly or in despair. Patients with prosodic comprehension
impairment would have difficulty recognizing these implicit emotional aspects.
In contrast, aprosodic patients generally do understand emotional statements
conveyed explicitly by the language rather than by the intonation (e.g., "I am
angry"). In these situations, of course, "I am very angry" would be understood
better than "I am angry" said with a very angry tone.
Within the right hemisphere, anterior lesions have been correlated with impaired prosodic production and posterior lesions with impaired prosodic comprehension [210-215]. It has even been suggested that the aprosodias may be
organized in the right hemisphere in a manner analogous to that for the aphasias
in the left hemisphere [210, 216].
However, difficulties with demonstrating interobserver reliability for portions of the bedside prosody examination [217], and mental control deficits
secondary to right hemisphere lesions [218], have complicated precise localization of the aprosodias. Furthermore, in view of the importance of the right
hemisphere in visuospatial function, the existence of two neuronal systems of
visual processing within the right hemisphere may also be central to the organization of emotional communication [219]. Facial and gestural communication
of emotion, the ability to extract situational connotations not explicitly stated,
and the ability to make inferences from written material can all be impaired as
a result of right hemisphere lesions [220]. All are relevant to effective emotional
communication, though not dependent on auditory intonational patterns.
7.3 Depression

Diagnosing depression after brain injury is difficult, because the presence of an


organic lesion excludes the classic diagnosis of depression [221]; component
subjective and objective factors may be dissociated; and criteria need to be altered
in a rehabilitation setting [222]. Nevertheless, components of depression are
frequently seen and need to be addressed.
Anatomical factors and situational adjustment factors warrant special consideration [223, 224]. Left hemisphere lesions have frequently been associated
with catastrophic reactions, particularly in patients who have nonfluent aphasia
[154,225]. A severe behavioral reaction in this type of patient sometimes may
be prevented by building sufficient success experience into the patient's program and by not placing demands beyond his or her reach. Recent studies have
shown a greater frequency of depression following left hemisphere damage
than following right hemisphere damage [226, 227]. Depression has been correlated with more anteriorly placed cortical or subcortical lesions [226-228].
In the subcortical cases, the importance of lesions of the caudate and anterior

3. Issues in Behavioral Neurology

75

limb of the internal capsule has been emphasized [228]. The possible interruption of ascending noradrenergic neuronal systems has been postulated, and
treatment with tricyclic antidepressants has been advocated [229-231].
A definite diagnosis of depression cannot be reached in some patients, particularly if their ability to communicate is limited. Apparent withdrawal from
therapy activities can be a clue in some patients who may respond to treatment.
The recognition of the appropriateness of depressed mood at some stages of
recovery and facilitation of the process of adjustment to disability are particularly crucial for successful rehabilitation.
The possibility that the brain-injured patient may commit suicide is of particular concern. Not only is suicidal ideation common after head injury, but death
from suicide has been reported to occur in 1 % of patients with war-time brain
injuries. Furthermore, the risk of suicide appears to be greatest late after injury,
reaching a peak more than 15 years postinjury. Risk factors have included
a change of character, problems in interpersonal relationships, difficulties in
attempting to return to work, excessive drinking, and depressive psychosis
[232].
8 AGGRESSIVE BEHAVIOR

8.1 "Aggression" Is Not a Diagnosis

One of the biggest clinical problems encountered after brain injury is aggression.
Although some patients are never aggressive, those who are can cause major
problems and risks for themselves, their families, the treating staff, and society.
The difficulty is compounded by the fact that aggression is a symptom, not a
specific diagnosis. The clinical context, underlying diagnosis, and situational
factors each playa major role [233, 234].
Factors that must be considered when one is trying to determine the underlying cause of a patient's aggression include premorbid personality, psychiatric
diagnosis, substance abuse, developmental experiences or abuse, epilepsy,
brain damage, and the patient's adjustment to disability and dependency. Brain
injury is particularly likely to cause or exacerbate aggressive behavior for at
least three reasons:
1. There is frequent injury to limbic, temporal, and frontal structures, causing
direct alteration of emotional behavior.
2. Deficits in linguistic and emotional communication, memory, and mental
control lead to misunderstandings, frustration, and rigidity of approach.
3. Brain injury frequently imposes significant dependency. This dependency
is often a source of anger.
8.2 Epilepsy and Violence

The role of epilepsy in violent behavior has been a matter of considerable controversy. Studies of violent behavior in prisoners initially showed an increased
frequency of epilepsy. However, subsequent studies that took into account the
underprivileged, lower socioeconomic class from which such prisoners usually

76

l. Theory and Intervention

come failed to confirm this finding. Carefully performed EEG studies of


patients referred forevaluation because of violent behavior have not shown a
significant incidence of epileptiform abnormalities. It has also been questioned
whether the violent behavior that is seen in some epileptics is related to the
epilepsy [235] or to the underlying brain damage or dysfunction [236].
The question of violent behavior during a seizure itself has been a matter of
medico-legal importance as well as diagnostic importance. Violent behavior
rarely occurs during a seizure, but when it does, it is typically "stereotyped,
simple, unsustained, and never supported by consecutive series of purposeful
movements" [237]. Although depth electrodes may show discharges that are
not evident on cortical electrodes [238] and behavioral contexts may affect the
nature of behavior associated with electrical discharge [239], the consensus
appears to be that sustained, directed violent behavior is incompatible with
seizure activity [235, 237, 240].
Postictal aggression has been described, but such aggression is usually nondirected and more frequently is a response to external stimulation rather than a
spontaneous behavior. Interictal aggression in patients with epilepsy is more
commonly encountered, and can even be directed aggression, but here again
the etiologic relationship has been questioned [236, 240].
8.3 Behavioral Context and Management

Practical management of the aggressive patient depends on recognition of both


the underlying etiological factors and the immediate precipitants. Management must be individualized; for such patients there are no "cookbook" approaches or "magic bullets." Behavior can remain problematic despite the best
of management.
Many patients pass through an acute agitated phase that may include aggression as part of a more global confusional state (see Section 4.2). Although
minimal use of medication is desirable for patients in the confusional state,
pharmacologic treatment may be needed when aggression occurs. The patients'
fear and pain should be particularly addressed. In addition, the environment
and amount of stimulation should be carefully structured.
The episodic dyscontrol syndrome occurs in patients who do not have general
impulsivity and yet lose control when they are angry-and who subsequently
feel remorseful about their aggressive behavior [241]. However, brain-injury
patients often exhibit a more widespread impulsivity, and this can also contribute to their aggressiveness. Anger and difficulties adjusting to dependency are
a major etiologic consideration in aggressive behavior. This is especially true
for patients for whom a longer time has passed since their injury, and who are
therefore past the initial acute confusional state.
8.4 Psychopharmacology and Aggression

There is no specific or consistently effective medication for controlling aggression. Individual patient characteristics, coexisting medical conditions or

3. Issues in Behavioral Neurology

77

allergies, other medications, or the clinical context may all influence drug management. Furthermore, medication is best used in conjunction with other behavioral techniques. A number of recent pharmacologic approaches will briefly
be reviewed below; this list is not exhaustive, and other agents may also be
considered. It should be recalled and considered that significant risks are attendant to all of the possible medications described.
Neuroleptic agents such as haloperidol, chlorpromazine, or thioridazine are
frequently used, and offer the distinct advantage of relatively rapid action [242,
243]. Patients who exhibit paranoia or psychosis may be particularly likely
to respond to these drugs. Side effects include sedation, hypotension, extrapyramidal effects, neuroleptic malignant syndrome [244], and tardive dyskinesia. Increased aggression secondary to drug-induced akathisia (motor
restlessness with inability to remain sitting) has also been reported [245, 246].
Additional concerns specifically for brain-injury survivors include potentially
adverse effects on motor recovery [247], memory [248], and posttraumatic
amnesia [249].
Benzodiazepines are frequently used and can be particularly helpful for states
of increased anxiety [250]. However, concerns have been raised regarding
"paradoxical rage reactions" with these drugs, and increased disinhibition of
behavior can be produced. Oxazepam may offer advantages over diazepam or
chlordiazepoxide in this regard [251, 252].
Lithium has successfully decreased aggression in prison populations [253,
254] and in mentally subnormal patients [255, 256]. Its use in brain-injury
cases also has been reported [257-259]. However, risks of toxicity are multiple.
Blood levels must be monitored closely [260], and specific neurotoxicity in
combination with haloperidol or carbamazepine has been reported [261, 262].
Beta-adrenergic blockers [263-266] such as propranolol and metoprolol
[267] have been advocated for the management of aggression. Propranolol was
initially advocated at low dosage [263], but subsequent studies have generally
shown that very high dosages are necessary to achieve the desired result [264266]. At these dosages, cardiovascular side effects have often been limiting
[268]. Pindolol, a beta blocker with partial agonist properties, has recently been
reported as effective in reducing aggression, with many fewer cardiovascular
complications [269].
The serotonergic neurotransmitter system may have an important inhibitory
role in preventing aggression [270, 271]. Certain antidepressant drugs such as
trazodone and amitriptyline block reuptake of serotonin and hence may augment serotonergic activity [272, 273]. Both have been reported to help reduce
agitation in some patients with brain abnormalities [273-275]. Patients who
exhibit characteristics of depression may be particulary good candidates for
such treatment. However, occasional paradoxical increased agitation has also
been reported [276].
One particularly exciting development has been the finding that carbamazepine may have beneficial psychotropic effects in some patients. Although its

78

I. Theory and Intervention

effects on limbic kindling [106] make it particularly attractive for use with epileptic patients, carbamazepine has beneficial effects for some nonepileptic patients
with mania [277], depression, and psychosis [278]. An increase in plasma tryptophan (a serotonin precursor) concentration with carbamazepine use has been
reported and may relate to its psychotropic but not to its anticonvulsant actions
[279]. Successful specific usage in patients with frontal pathology [280] and
rage outbursts [281] has been described.
In addition, methylphenidate has been tried in patients with preexisting attention deficit disorder [282], and hormonal agents have been tried for sexual
offenders [283].
Each of the drugs listed here has sometimes been effective in the treatment
of aggressive behavior but on other occasions either has been ineffective or has
caused complications. Thus, drug management continues to remain based on
individual patient needs.
9 DENIAL, INSIGHT, AND ADJUSTMENT

9.1 The Anatomy of Denial (Anosognosia)

Denial is a phenomenon that certainly is not restricted to patients and is often


observed in people dealing with catastrophic circumstances [284, 285]. The
sudden onset, severity, and threatening nature of damage to the brain may
well be significant in the frequent occurrence of denial after brain injury.
However, denial has also been found to occur most frequently in patients who
have specific lesions of the brain. Anosognosia literally means "lack of knowledge of the existence of disease" [202]. Although the term was originally used
to describe denial ofleft hemiparesis by patients with right hemisphere lesions,
it also refers to the denial of other di~aotlities. Denial of blindness in Anton's
syndrome is probably best known, but denial of such unrelated phenomena as
an amputated limb or enucleated eye has been reported.
Five types of explicit verbal denial have been delineated [202]: 1) complete
denial of any disability; 2) denial of major disability; 3) minimization, or attribution to some benign cause; 4) projection of the disability outside the self; and
5) temporal displacement of the disability. Different types can be exhibited
successively, and an absence of anxiety is characteristic.
In patients with lesions of the right hemisphere, denial has classically been
attributed to right parietal lobe involvement. However, Weinstein [202] has
emphasized instead the importance of deep involvement in the regions surrounding the third and lateral ventricles, diencephalon, and midbrain. Disruption
of the centrencephalic system, including the reticular formation, thalamus, and
thalamocortical connections has been hypothesized. It is noteworthy that these
regions are often affected by closed-head injuries, including diffuse ones. Although fluent aphasia may limit recognition of anosognosia with posterior left
hemisphere lesions, a count of cases reveals marked right hemisphere predominance in the two major reported series [202, 286]. The decreased insight that is
so central to the frontal syndrome [287] is a closely related phenomenon-and,
of course, is also frequent in brain-injured patients.

3. Issues in Behavioral Neurology

79

Management of denial is difficult. Patients are characteristically not swayed


by logical argument. Some patients appear to gain insight after a restricted failure
experience, but others do not. As patients improve and their denial decreases,
anxiety and depression may increase. Supportive counseling is particularly important for these patients.
9.2 Adjustment

This chapter has emphasized the correlation between the anatomic aspects of
cerebral dysfunction and the clinical manifestations frequently exhibited by
brain-injured patients. However, it is evident that the patients' personal reactions have a tremendous impact on their overall adjustment. This is probably
the most crucial area for intervention and is one that is addressed at length in
other chapters of this book. Adjustment aspects are generally superimposed
upon and intertwined with the anatomic ones. There is no doubt that a more
precise understanding of clinical strengths and weaknesses can facilitate addressing these key adjustment issues and optimize personal functioning.
REFERENCES

1. Geschwind N, (1965). Disconnexion syndromes in animals and man. Brain 88, 237-294,
585-644.
2. Damasio, A. (1984). Behavioral neurology: Research and practice. Semin. Neurol. 4,117119.
3. Geschwind, N. and Kaplan, E. (1962). A human cerebral disconnection syndrome. Neurology 12, 675-685.
4. Kaplan, E. (1983). Process and achievement revisited. In Toward a Holistic Developmelltal
Psychology, Wapner S., and Kaplan B., eds. Erlbaum Assoc., Hillsdale, Nj, pp. 143-156.
5. Naeser, M.A. and Hayward, R.W. (1978). Lesion localization in aphasia with cranial computed tomography and the Boston Diagnostic Aphasia Exam. Neurology 28, 545-551.
6. Gado, M., Hanaway,)' and Frank, R. (1979). Functional anatomy of the cerebral cortex by
computed tomography.). Comput. Assist. Tomogr. 3, 1-19.
7. Naeser, M.A. and Helm-Estabrooks, N. (1985). CT scan lesion localization and response to
Melodic Intonation Therapy with nonfluent aphasia cases. Cortex 21, 203-223.
8. Lezak, M.D. (1978). Living with the characterologically altered brain injured patient.). Clin.
Psychiatry 39, 592-598.
9. Brooks, D.N. and Aughton, M.E. (1979). Psychological consequences of blunt head injury.
Int. Rehabil. Med. 1, 160-165.
10. Thompson, I.V. (1984). Late outcome of very severe blunt head trauma: A 10-15 year
second follow up. j. Neurol. Neurosurg. Psychiatry 47, 260-268.
11. Damasio, A. R. (1985). Disorders of complex visual processing. In Principles ~f behavioral
neurology. Mesulam, M.-M., ed., F.A. Davis Company, Philadelphia, pp. 259-288.
12. Auerbach, S.H., Allard, T., Naeser, M., Alexander, M.P. and Albert, M.L. (1982). Pure
word deafness: Analysis of a case with bilateral lesions and a defect at the prephonemic level.
Brain 105, 271-300.
13. Benson, D.F. and Blumer, D. (1982). Psychiatric Aspects of Neurologic Disease, Vol. II. Grune
& Stratton, New York, p. xvi.
14. Frankowski, R.F. (1986). The demography of head injury in the United States. In Neurotrauma, Miner, M.E. and Wagner, K.A., eds., Butterworths, pp. 1-17.
15. Teasdale, G. and jennett, B. (1974). Assessment of coma and impaired consciousness.
Lancet, 2, 81-84.
16. jennett, B. and Bond, M. (1975). Assessment of outcome after severe brain damage. Lancet
1, 480-484.
17. Alexander, M.P. (1982). Traumatic brain injury. In Psychiatric Aspects of Neurologic Disease;
vol. II. Benson, D.F. and Blumer, D. eds., Grune & Stratton, New York, pp. 219-249.

80

\. Theory and Intervention

18. Hagen, C, Malkmus, D. and Durham, P. (1977). Levels of cognitive functioning. Presented
at Head Trauma Rehabilitation Seminar, Rancho Los Amigos Hospital, Los Angeles, CA.
19. Bond, M.R. (1979). The stages of recovery from severe head injury with special reference to
late outcome. Int. Rehabil. Med. 1, 155-159.
20. Langfitt, T.W. and Gennarelli, T.A. (1982). A holistic view of head injury including a new
clinical classification. In Head If/jury: Basic alld elillical Aspects. Grossman, R.G., and
Gildenberg, P.L., eds., Raven Press, New York, pp. 1-15.
21. Auerbach, S.H. (1983). Cognitive rehabilitation in the head injured: A neurobehavioral
approach. Semin Neurol 3, 152-163.
22. Long, D.F. and Alexander, M.P. (1984). Outcome characteristics of subgroups of closed
head injury. Presented at 5th Annual Traumatic Head Injury Conference, Braintree, MA.
23. Uzzell, B.P., Zimmerman, R.A., Dolinskas, CA. and Obrist, W.D. (1979). Lateralized
psychological impairment associated with CT lesions in head injured patients. Cortex 15,
391-401.
24. Lobato, RD., Cordobes, F., Rivas,j.j., de la Fuente, M., Montero, A., Barcena, A., Perez,
C, Cabrera, A. and Lamas, E. (1983). Outcome from severe head injury related to the type
of intracranial lesion. j. Neurosurg. 59, 762-774.
25. Adams, j.H., Graham, D.\., Murray, L.S. and Scott, G. (1982). Diffuse axonal injury due to
nonmissile head injury in humans: An analysis of 45 cases. Ann. Neurol. 12, 557-563.
26. Gennarelli, T.A., Thibault, L.E., Adams,j.H., Graham, D.\., Thompson, Cj. and Marcincin, R.P. (1982). Diffuse axonal injury and traumatic coma in the primate. Ann. Neurol. 12,
564-574.
27. Povlishock, T. T., Becker, D.P., Cheng, CL. Y. and Vaughan, G. W. (1983). Axonal change
in minor head injury. j. Neuropathol. Exp. Neurol. 42, 225-242.
28. Levin, H. S. and Eisenberg, H. M. (1986). The relative durations of coma and posttranmatic
amnesia after severe non missile head injury: Findings from the pilot phase of the National
Traumatic Coma Data Bank. In Nellrotrawna, Miner, M.E. and Wagner, K.A., eds.,
Butterworths, Boston, pp. 89-97.
29. Russell, W.R. (1971). The Traulllatic Alllllfsias. Oxford University Press, New York,
pp. 1-78.
30. Povlishock,j.T., Becker, D.P., Miller,j.D., jenkins, L.W. and Dietrich, W.D. (1979). The
morpho pathologic substrates of concussion? Acta Neuropathol. 47, 1-11.
31. Zimmerman, RA., Bilaniuk, L.T. and Gennarelli, T. (1978). Computed tomography of
shearing injuries of the cerebral white matter. Radiology 127, 393-396.
32. Zimmerman, RA. and Bilaniuk, L.T. (1979). Computed tomography in diffuse traumatic
cerebral injury. In Neurological Trauma, Popp, A.J., et aI., eds. Raven Press, New York,
pp. 253-262.
33. Snoek, j., Jennett, B., Adams, j.H., Graham, D.1. and Doyle, D. (1979). Computerized
tomography after recent severe head injury in patients without acute intracranial hematoma.
j. Neurol. Neurosurg. Psychiatry 42, 215-225.
34. Cooper, P.R, Maravilla, K., Kirkpatrick, j., Moody, S.F., Sklor, F.H., Diehl, j. and Clark,
W.K. (1979). Traumatically induced brain stem hemorrhage and the computerized tomographic scan: Clinical, pathological and experimental observations. Neurosurgery 4, 115124.
35. Zuccarello, M., Fiore, D.L., Trincia, G., DeCaro, R., Pardatscher, K. and Andrioli, G.C
(1983). Traumatic brain stem hemorrhage. Acta Neurochir. 67, 103-113.
36. Auerbach, S., Friedman, j., Shape, R., Weinberg, R. and Moore, S. (1984). Disorders of
attention in closed head injuries: Correlations with a failure to visually suppress the vestibular-ocular reflex in diffuse axonal injury. Neurology 34 (Sup pI. 1), 189-190.
37. Sharpe, j.A. (1986). Supranuclear disorders of horizontal eye motion. Semin. Neurol. 6,
155-166.
38. Levin, H.S., Amparo, E., Eisenberg, H.M., Williams, D.H., High W.M., McArdle, CB.
and Weiner, R.L. (1987). Magnetic resonance imaging and computerized tomography in
relation to the neurobehavioral sequelae of mild and moderate head injuries. j. Neurosurg.
66, 706-713.
39. Wilberger, JE., Deeb, Z. and Rothflls, W. (1987). Magnetic resonance imaging in cases of
severe head injury. Neurosurgery 20, 571-576.
40. Levin, H.S., Kalisky, Z., Handel, S.F., Goldman, A.M., Eisenberg, H.M., Morrison, D.

3. Issues in Behavioral Neurology

41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.

81

and Von Laufen, A. (1985). Magnetic resonance imaging in relation to the sequelae and rehabilitation of diffuse closed head injury: Preliminary findings. Semin. Neurol. 5, 221-232.
Zimmerman, R.A., Bilaniuk, L.T., Bruce, D., Dolinskas, C, Obrist, W. and Kuhl, D.
(1978). Computed tomography of pediatric head trauma: Acute general cerebral swelling.
Radiology 128, 403-408.
Uzzell, B.P. (1986). Pathophysiology and behavioral recovery. In Clinical Neuropsychology of
Intfl"vention, Uzzell, B.P. and Gross, Y. eds., Martinus Nijhoff, Boston, pp. 3-18.
Teasdale, E., Cardoso, E., Galbraith, S. and Teasdale, G. (1984). CT scan in severe diffuse
head injury: Physiological and clinical correlations. J. Neurol. Neurosurg. Psychiatry 47,
600-603.
vanDongen, K.J., Braakman, R, and Gelpke, G.J. (1983). The prognostic value of computerized tomography in comatose head injured patients. J. Neurosurg. 59, 951-957.
Esperson, J.O. and Peterson, O.F. (1982). Computerized tomography (CT) in patients with
head injuries: Assessment of outcome based upon initial clinical findings and initial CT scans.
Acta Neurochir. 65, 81-91.
Graham, D.I., Adams, J.H. and Doyle, D. (1978). Ischemic brain damage in fatal nonmissile head injuries. J. Neurol. Sci. 39, 213-234.
Levy, D.E., Caronna, J.J., Singer, B.H., Lpinski, R.H., Frydman, H. and Plum, F. (1985).
Predicting outcome from hypoxic-ischemic coma. JAMA 253, 1420-1426.
Ingvar, D.H., Brun, A., Johansson, L. and Samuels son, S.M. (1978). Survival after severe
cerebral anoxia with destruction of the cerebral cortex: The apallic syndrome. Ann. N.Y.
Acad. Sci. 315, 184-214.
Brierley, J.B. and Graham, D.1. (1984). Hypoxia and vascular disorders of the central nervous
system. In Greenfield's Neuropathology, Adams, J.H., Corsellis, J.A.N. and Duchen, L.W.
eds., John Wiley & Sons, New York, pp. 125-207.
Gurdjian, E. S. and Gurdjian, E. S. (1976). Cerebral contusions: Re-evaluation of the
mechanism of their development. J. Trauma 16, 35-51.
Courville, C B. (1937). Pathology of the Cmtral Nervous System, part 4, Pacific, Mountain
View, CA.
Blumer, D. and Benson, D.F. (1975). Personality changes with frontal and temporal lobe
lesions. In Psychiatric Aspects of Neurologic Disease, Benson, D. F. and Blumer, D. eds., Grune
& Stratton, New York, pp. 151-170.
Stuss, D.T. and Benson, D.F. (1986). The Frontal Lobes. Raven Press, New York,
pp. 1-249.
Damasio, A. (1979). The frontal lobes. In Clinical Neuropsychology, Heilman, K.M. and
Valenstein E. eds., Oxford University Press, New York, pp. 360-412.
Luria, A.R (1980). Highfl" Cortical FunctiollS ill Mall. Basic Books, New York, pp. 246-365.
Fuster, J.M. (1980). The Prefrontal Cortex. Raven Press, New York, pp. 125-142.
Cummings, J. L.. (1985). Clinical Neuropsychiatry. Grune & Stratton, New York, pp. 57-67.
Adamovich, B.B., Henderson, J.A. and Auerbach, S. (1985). Cogllitive Rehabilitation of
Closed Head Iniured Patients: A Dynamic Approach. College Hill Press, San Diego, pp. 20-23.
Minderhoud, T.M., Huizenga, J. and vanWoerkom, T.CA.M. (1982). The pattern ofrecovery after severe head injury. Clin. Neurol. Neurosurg. 84, 15-27.
Benson, D.F. 1973. Psychiatric aspects of aphasia. Br. J. Psychiatry 123, 555-566.
Levin, H.S., Madison, CF., Bailey, CB., Meyers, CA., Eisenberg, H.M. and Guinto,
F.G. (1983). Mutism after closed head il~ury. Arch. Neurol. 40, 601-606.
Maki, Y., Akimoto, H. and Enomoto, T. (1980). Injuries of basal ganglia following head
trauma in children. Child's Brain 7, 113-123.
MacPherson, P., Teasdale, E., Dhaker, S., Allerdyce, G. and Galbraith, S. (1986). The
significance of traumatic hematoma in the region of the basal ganglia. J. Neurol. Neurosurg.
Psychiatry 49, 29-34.
Plum, F. and Posner, J.B. (1980). The Diagnosis qf Stupor and Coma. F.A. Davis Company,
Philadelphia, pp. 1-373.
Sunderland, S. (1958). The tentorial notch and complications produced by herniations of the
brain through that aperture. Br. J. Surg. 45, 422-438.
Romanul, F.C.A. (1970). Examination of the brain and spinal cord. In Neuropathology,
Tedeschi CG., ed., Little, Brown and Co., Boston, pp. 131-214.
Damasio, A.R and Damasio, H. (1983). The anatomic basis of pure alexia. Neurology 33,

82

I. Theory and Intervention

1573-1583.
68. Benson, D.F., Marsden, C.D. and Meadows, J.L. (1974). The amnesic syndrome of posterior cerebral artery occlusion. Acta Neurol. Scand. 50, 133-145.
69. Decroix, J. P., Cambier, J. and Masson, M. (1985). The anterior choroidal artery syndrome.
Presse Med. 14, 1085-1087.
70. Helgason, C, Caplan, L.R. Goodwin, J. and Hedges, T. (1986). Anterior choroidal arteryterritory infarction: Report of cases and review. Arch. Neurol. 43, 681-686.
71. Tomlinson, B.E. (1970). Brain stem lesions after head injury. J. Clin. Pathol. 23(Suppl. 4),
154-165.
72. Rosenblum, W.I., Greenberg, R.P., Seelig,J.M. and Becker, D.P. (1981). Midbrain lesions:
Frequent and significant prognostic feature in closed head injury. Neurosurgery 9, 613-620.
73. Born, J.D., Albert, A. Hans, M.D. and Bonnal, J. (1985). Relative prognostic value of best
motor response and brain stem reflexes in patients with severe head injury. Neurosurgery 16,
595-601.
74. Jellinger, K. (1977). Pathology and pathogenesis of apallic syndromes following closed head
injuries. In The Apallic Syndrome. Daile are, G. Gerstenbrand, F., Lucking, CH., Peters G.
and Peters, V.H., eds., Springer-Verlag, New York, pp. 88-103.
75. Jellinger, K. (1986). Neuropathology and clinical signs of brainstem disorders. In Clillical
Problems of BraiflStem Disorders. Kunze, K. Zangemeister, W.H. and Arlt A. eds., Thieme
Medical Pubs., Inc., New York, pp. 17-29.
76. Fisher, CM. (1977). The clinical picture in occult hydrocephalus. Clin. Neurosurg. 24, 270284.
77. Black, P. McL., Ojemann, R.G. and Tzouras, A. (1985). CSF shunts for dementia, incontinence, and gait disturbance. Clin. Neurosurg. 32, 632-651.
78. Lanksch, W., Grumme, T. and Kazner, E. (1979). Computed Tomography ill Head Injuries.
Springer-Verlag, New York, p. 125.
79. Gerard, G. and Weisberg, L.A. (1986). Magnetic resonance imaging in adult white matter
disorders and hydrocephalus. Semin. Neurol. 6, 17-23.
80. Bradley, W.G., Jr. (1987). Pathophysiologic correlates of signal alterations. In Magl/etic Reso/wllce Imagillg of the Celltral Nervous System. Brant-Zawadzki, M. and Norman, D. eds.
Raven Press, New York, p. 32.
81. Wikkels6, C, Andersson, H., Blomstrand, C, Lindqvist, G. and Svendsen, P. (1986).
Normal pressure hydrocephalus: Predictive value of the cerebrospinal fluid tap-test. Acta
Neurol. Scand. 73, 566-573.
82. B0rgesen, S.E. and Gjerris, F. (1982). The predictive value of conductance to outflow of
CSF in normal pressure hydrocephalns. Brain lOS, 65-86.
83. Thomsen, A. M., B0rgesen, S.E., Bruhn, P. and Gjerris, F. (1986). Prognosis of dementia in
normal-pressure hydrocephalus after a shunt operation. Ann. Neurol. 20, 304-310.
84. Meyers, CA., Levin, H.S., Eisenberg, H.M. and Guinto, F.C (1983). Early versus late
lateral ventricular enlargement following closed head injury. J. Neurol. Neurosurg. Psychiatry 46, 1092-1097.
85. Levin, H.S., Meyers, CA., Grossman, R.G. and Sarwar, M. (1981). Ventricular enlargement after closed head injury. Arch. Neurol. 38, 623-629.
86. Kishore, P.R.S., Lipper, M.H., Miller, J.D., Girevendulis, A.K., Becker, D.P. and Vines,
F. S. (1978). Post-traumatic hydrocephalus in patients with severe head injury. Neuroradiology 16, 261-265.
87. Bradley, W.G. (1987). Magnetic resonance appearance of flowing blood and cerebrospinal
fluid. In Magnetic ReSOllallce Imagillg of the Celltral Nervous System, Brant-Zawadzki, M. and
Norman, D., eds., Raven Press, New York, pp. 93-96.
88. Jennett, B. (1975). Epilepsy After NOll-Missile Head Injuries. Year Book Medical Pubs.; Inc.,
Chicago; pp. 1-179.
89. Annegers, J.F., Grabow, J.D., Groover, R. V., Laws, E. Elveback, L.R and Kurland, L. T.
(1980). Seizures after head trauma: A population study. Neurology 30, 683-689.
90. D'Alessandro, R, Ferrara, R, Benassi, G., Lenzi, P.L. and Sabattini, L. (1988). Computed
tomographic scans in posttraumatic epilepsy. Arch. Neurol. 45, 42-43.
91. Guidice, M.A. and Berchov, R.C (1987). Post-traumatic epilepsy following head injury.
Brain Injury I, 61-64.
92. Caveness, W.F., Meirowsky, A.M., Rish, B.L., Mohr, J.P., Kistler, J.p., Dillon, J.D. and

3. Issues in Behavioral Neurology

83

Weiss, G.H. (1979). The nature of posttraumatic epilepsy. J. Neurosurg. 50, 545-553.
93. Jennett, B. and Teasdale, G. (1981). Mallagemellt of Head IIl;uries. F.A. Davis Company,
Philadelphia, pp. 281-288.
94. Feeney, D.M. and Walker, A.E. (1979). The predicition of posttraumatic epilepsy. Arch.
Neurol. 36, 8-12.
95. Jennett, B., Teather, D., and Bennie, S. (1973). Epilepsy after head injury-residual risk
after varying fit-free intervals since injury. Lancet 2, 652-653.
96. Willmore, L.J (1986). Mechanisms and management of posttraumatic epilepsy. In Neuyotrauma, Miner M.E., and Wagner, K.A. eds., Butterworths, Boston, pp. 99-106.
97. Racine, R.J and Mcintyre, D. (1986). Mechanisms of kindling: A current view. In The
Limbic System: FunctiOlwl OrganizatiOlI d1Id Clinical Disorders, Doane, B.K. and Livingston,
K.E. cds. Raven Press, New York, pp. 109-121.
98. Goddard, G.V., Dragunow, M., Marv, E. and Macleod, E.K. (1986). Kindling and the
forces that oppose it. In The Limbic System: FUllctional OrgallizatiOlI alld Clillical Disorders,
Doane, B.K. and Livingston, K.E., eds., Raven Press, New York, pp. 95-108.
99. Oles, K. S. and Penry, J K. (1985). Pharmacological prophylaxis of post-traumatic seizures.
In CUrYCllt Therapy ill NeuroloRic Disease, 1985-1986, Johnson, R. T. ed., B.C. Decker, Philadelphia, pp. 46- 51.
100. Penry, JK., White, B.G. and Bracket, C.E. (1979). A controlled prospective study of the
pharmacologic prophylaxis of posttraumatic epilepsy. Neurology 29, 600-601.
101. Young, B., Rapp, RP., Norton,JA., Haack, D., Tibbs, P.A. and Bean,JR. (1983). Failure
of prophylactically administered phenytoin to prevent late posttraumatic seizures. J. Neurosurg. 58, 236-241.
102. Mcqueen, J.K., Blackwood, D.H.R., Harris, P., Kalbag, RM. and Johnson, A. (1983).
Low risk oflate post-traumatic seizures following severe head injury: implications for clinical
trials of prophylaxis. J Neurol. Neurosurg. Psychiatry 46, 899-904.
103. Reynolds, E.H. (1983). Mental effects of antiepileptic medication: A review. Epilepsia
24(Suppl. 2), 585-596.
104. Shaw, M.D.M., Foy, P., and Chadwick, D. (1983). The effectiveness of prophylactic anticonvulsants following neurosurgery. Acta Neurochur. 69, 253-258.
105. Gli:itzner, F.L., Haubitz, /., Miltner, F., Kapp, G. and Pflughaupt, K.-W. (1983). Anfallsprophylaxe mit Carbamazepin nach schweren Schadelhirnverletzungen. Neurochirurgia 26,
66-79.
106. Albright, P.S., Burnham, W.M. and Livingston, K.E. (1986). Seizure patterns and pharmacological responses in the kindling model. In The Limbic System: Ftlflctiollal OrRallizatioll alld
Clinical Disorders, Doane B.K. and Livingston, K.E. eds. Raven Press, New York, pp. 147157.
107. Evans, R. W. and Gualtieri, C. T. (1985). Carbamazepine: A neuropsychological and psychiatric profile. Clin. Neuropharmacol. 8, 221-241.
108. Hart, R.G. and Easton, JD. (1982). Carbamazepine and hematological monitoring. Ann.
Neurol. 11, 309-312.
109. Fariello, RG., Booker, H.E., Chun, R.W.M. and Orrison, W.W. (1983). Reenactment of
the triggering situation for the diagnosis of epilepsy. Neurology 33, 878-887.
110. Waxman, S.G. and Geschwind, N. (1974). Hypergraphia in temporal lobe epilepsy. Neurology 24, 629-636.
111. Bear, D. M. (1986). Behavioral changes in temporal lobe epilepsy: Conflict, confusion,
challenge. In Aspects of Epilepsy alld Psychiatry, Trimble M.R., and Bolwig, T.G. eds., John
Wiley & Sons, New York, pp. 19-30.
112. Bear, D.M. (1979). Temporal lobe epilepsy-a syndrome of sensory limbic hyperconnection.
Cortex 15, 357-384.
113. Dam, M. and Dam, A.M. (1986). Is there an epileptic personality? In Aspects of Epilepsy mid
Psychiatry, Trimble; M.R and Bolwig, T.G. eds., John Wiley & Sons, New York, pp. 9-18.
114. Stark-Adamec, C. and Adamec, RE. (1986). Psychological methodology versus clinical
impressions: different perspectives on psychopathology and seizures. In The Limbic System:
Functional OrRanizatioll mid Clinical Disorders, Doane, B.K. and Livingston, K.E. eds. Raven
Press, New York, pp. 217-227.
115. Trimble, M.R. (1986). Hypergraphia. In Aspects of Epilepsy and Psychiatry, Trimble, M.R.
and Bolwig, T.G. eds. John Wiley & Sons, New York, pp. 75-88.

84

I. Theory and Intervention

116. Trimble, M.R. (1982). The interictal psychoses of epilepsy. In Psychiatric Aspects of New'ological Disease, Vol. II. Benson, D.F. and Blumer, D. eds. Grune & Stratton, New York,
pp. 75-91.
117. Flor-Henry, P. (1983). Determinants of psychosis in epilepsy: Laterality and forced normalization. Bio. Psychiatry 18, 1045-1057.
118. Trimble, M.R. (1986). PET scanning in epilepsy. In Aspects of Epilepsy al/d Psychiatry, Trimble M.R., and Bolwig, T.G. eds., john Wiley & Sons, New York, pp. 147-162.
119. Trimble, M.R. (1986). Radiological studies in epileptic psychosis. In The Lilllbic Systelll:
FUI/ctiol/al Orgal/izatiol/ alld Clillical Disorders, Doane, B.K. and Livingston, K.E. eds., Raven
Press, New York, pp. 195-199.
120. Levine, D. N. and Finklestein, S. (1982). Delayed psychosis after right temporoparietal
stroke or trauma: Relation to epilepsy. Neurology 32, 267-273.
121. Sindrup, E. (1986). Epilepsy and psychosis: Electrophysiological aspects. In Aspects of
Epilepsy al/d Psychiatry, Trimble, M.R. and Bolwig, T.G. eds., john Wiley & Sons, New
York, pp. 163-176.
122. Wolf, P. (1986). Forced normalization. In Aspects ~f Epilepsy and Psychiatry, Trimble, M.R.
and Bolwig, T.G. eds. john Wiley & Sons, New York, pp. 101-116.
123. Pakalnis, A., Drake, M.E., jr., Kuruvilla, j. and Kellum, j.B. (1987). Forced normalization:
Acute psychosis after seizure control in seven patients. Arch. Neurol. 44, 289-292.
124. Griffith, J.L. (1988). Comments on "Forced normalization: Acute psychosis after seizure
control in seven patients." In 1988 Year Book o{Neurology mld Neurosw:~ery, Dejong, R.N.
Currier, R.D., and Corwell, R.M., eds. Year Book Medical Pubs., Inc., Chicago, pp. 105106.
125. Robertson, M. M. (1986). Ictal and interictal depression in patients with epilepsy. In Aspects
~f Epilepsy and Psychiatry, Trimble and M.R. Bolwig, T.G. eds., john Wiley & Sons, New
York, pp. 213-234.
126. Edwards, J. G. (1985). Antidepressants and seizures: Epidemiological and clinical aspects. In
The Psychopharmacology of Epilepsy, Trimble, M.R., ed., john Wiley & Sons, New York,
pp. 119-139.
127. Ojemann, L.M., Baugh-Bookman, C. and Dudley, D.L. (1987). Effect of psychotropic
medications on seizure control in patients with epilepsy. Neurology 37, 1525-1527.
128. Bricolo, A., Turazzi, S. and Feriotti, G. (1980). Prolonged posttraumatic unconsciousness.
J. Ncurosurg. 52, 625-634.
129. Bricolo, A. 1976. Prolonged post-traumatic coma. In Hmldbook o{Clil/ical Neurology, Vinken,
P.J. and Bruyn, G.W. eds., North Holland, Amsterdam, pp. 699-755.
130. Peters, V.H. and Gerstenbrand, F. (1977). Clinical picture and problems in terminology. In
The Apallic SYI/drollle, Daile Ore, G., Gerstenbrand, F., Lucking, C.H., Peters, G. and
Peters, V.H. cds., Springer-Verlag, New York, pp. 8-13.
131. jennett, B. and Plum, F. (1972). Persistent vegetative state after brain damage. Lancet 1,
734-737.
132. Berrol, S. (1986). Evolution and the persistent vegetative state. J. Head Trauma Rehabil. 1,
7-13.
133. Kretschmer, E. (1940). Der apallische Syndrome. Zbl. ges Neurol. Psychiat. 169,576-579.
134. Gerstenbrand, F. 1977. The symptomatology of the apallic syndrome. In The Apallic SYI/drome, Daile Ore, G., Gcrstenbrand, F., Lucking, C.H., Peters, G. and Peters, V.H. eds.,
Springer-Verlag, New York, pp. 14-21.
135. Segarra, J.M. (1970). Cerebral vascular disease and behavior I: The syndrome of the mesencephalic artery. Arch. Neurol. 22, 408-418.
136. Long, D. and Peters, S. (1987). The Bryn Mawr Coma Emergence Scale and low level team
graph. Presented at the Pennsylvania Head Injury Foundation Conference, April.
137. Higashi, K., Sakata, Y., Hatano, M., Abiko, S., Ihara, K., Katayama, S., Wakuta, Y.,
Okamura, T., Veda, H., Zenke, M. and Aoki, H. (1977). Epidemiological studies on patients
with a persistent vegetative state. J. Neurol. Neurosurg. Psychiatry 40, 876-885.
138. Higashi, K., Hatano, M., Abiko, S., Ihara, K., Katayama, S., Wakata, Y., Okamura, T. and
Yamashita, T. (1981). Five year follow-up study of patients with persistent vegetative state.
J. Neurol. Neurosurg. Psychiatry 44, 552-554.
139. Roberts, A.H. (1979). Severe Aaidel/tal Head II/jury. MacMillan Press, Ltd., New York, pp.
1-226.

3. Issues in Behavioral Neurology 85

140. Long, D. and McNichol, B.H. (1986, June). Rehabilitation of low level head trauma patients-predictors of outcome. In Abstracts ~f the Proceedings, 10th Annual Postgraduate
Course on Rehabilitation of the Brain Injured Adult and Child, Williamsburg, VA.
141. Uematsu, D., Suematsu, M., Fukuuchi, Y., Ebihara, S. and Gotoh, F. (1985). Midbrain
locked-in state with oculomotor nucleus lesion. J. Neurol. Neurosurg. Psychiatry 48, 952956.
142. Keane, J.R. and Habashi, H.H. (1985). Locked-in syndrome due to tentorial herniation.
Neurology 35, 1647-1649.
143. Mesulam, M.-M. (1981). A cortical network for directed attention and unilateral neglect.
Ann. Neurol. 10,309-325.
144. Heilman, K.M., Valenstein, E. and Watson, R.T. (1984). Neglect and related disorders.
Semin. Neurol. 4, 209-219.
145. Laplane, D. and Degos, J.D. (1983). Motor neglect. J. Neurol. Neurosurg. Psychiatry 46,
152-158.
146. Fleet, W.S., Valenstein, E., Watson, R.T. and Heilman, K.M. (1987). Dopamine agonist
therapy for neglect in humans. Neurology 37, 1765-1770.
147. Allison, R.S. (1966). Perseveration as a sign of diffuse and focal brain damage. Br. Med. J. 2,
1027-1032, 1095-1101.
148. Luria, A.R. (1965). Two kinds of motor perseveration in massive injury of the frontal Lobes.
Brain 88, 1-10.
149. Sands on, J. and Albert, M.L. (1987). Perseveration in behavioral neurology. Neurology 37,
1736-1741.
150. Yamadori, A. (1981). Verbal perseveration in aphasia. Neuropsychologia 19, 591-594.
151. Kapur, N. (1985). Double dissociation between perseveration in memory and problem
solv.iog tasks. Cortex 21, 461-465.
152. Hudson, A.J. (1968). Perseveration. Brain 91, 571-582.
153. Jasper, H.H. (1931). Is perseveration a functional unit participating in all behavior processes?
J. Soc. Psychol. 2, 28-51.
154. Goldstcin, K. (1948). LanJiuaJie alld LallJiuage Disturbances, Grune & Stratton, New York,
pp. 16-18.
155. Halpern, H. (1965). Effect of stimulus variables on verbal perseveration of dysphasic subjects. Percept. Mot. Skills 20, 421-429.
156. Helm-Estabrooks, N., Emery, P. and Albert, M.L. (1987). Treatment of aphasic perseveration (TAP) program. Arch. Neurol. 44, 1253-1255.
157. Lhermitte, F., Pillon, B. and Serdaru, M. (1986). Human autonomy and the frontal lobes.
Part I: Imitation and utilization behaviors: A neuropsychological study of75 patients. Ann.
Neurol. 19,326-334.
158. Lhermitte, F. (1983). "Utilization behavior" and its relation to lesions of the frontal lobes.
Brain 106, 237-255.
159. Lhermitte, F. (1986). Human autonomy and the frontal lobes. Part II: Patient behavior in
complex and social situations: the "environmental dependency syndrome." Ann. Neurol.
19,335-343.
160. Ross, E.D. and Stewart, R.M. (1981). Akinetic mutism from hypothalamic damage: Successful treatment with dopamine agonists. Neurology 31, 1435-1439.
161. vanZomeren, A.H., Brouwer, W.H. and Deelman, B.G. (1984). Attentional deficits: The
riddles of selectivity, speed, and alertness. In Closed Head Injury, Brooks N., ed., Oxford
University Press, New York, pp. 74-107.
162. Gronwall, D.M.A. (1977). Paced auditory serial-addition task: A measure of recovery from
concusion. Percept. Mot. Skills 44, 367-373.
163. Benson, D.F. (1975). Disorders of verbal expression. In Psychiatric Aspects of Neurological
Disease, Benson D.F. and Blumer D., eds., Grune and Stratton, New York, pp. 121-136.
164. Cummings, J.L., Benson, D.F., Houlihan, J.P. and Gosenfeld, L.F. (1983). Mutism: Loss of
neocortical and limbic vocalization. J. Nerv. Ment. Dis. 171,255-259.
165. Helm, N.A., Butler, R.B. and Benson, D.F. (1978). Acquired stuttering. Neurology 28,
1159-1165.
166. Baratz, R. and Mesulam, M.-M. (1981). Adult onset stuttering treated with anticonvulsants.
Arch. Neurol. 38, 132.
167. Ludlow, C.L. Rosenberg, J., Salazar, A., Grafman, J. and Smulak, M. (1987). Site of pene-

86

I. Theory and Intervention

trating brain lesions causing chronic acquired stuttering. Ann. Neurol. 22, 60-66.
168. Benson, D.F. (1979). Aphasia, Alexia, alld Agraphia. Churchill Livingstone, New York,
pp. 1-213.
169. Benson, D. F. and Geschwind, N. (1985). Aphasia and related disorders: A clinical approach.
In Principles oj Behavioral NeuroloJIY, Mesulam, M.-M., ed., F.A. Davis Company, Philadelphia, pp. 193-258.
170. Freedman, M., Alexander, M.P. and Naeser, M.A. (1984). Anatomic basis of transcortical
motor aphasia. Neurology 34, 409-417.
171. Alexander, M.P. and Loverme, S.R. (1980). Aphasia after left hemispheric intracerebral
hemorrhage. Neurology 30, 1193-1202.
172. Graff-Radford, N.R. and Damasio, A.R. (1984). Disturbances of speech and language associated with thalamic dysfunction. Semin. Neurol. 4, 162-168.
173. Naeser, M.A., Alexander, M.P., Helm-Estabrooks, N., Levine, H.L., Laughlin, S.A. and
Geschwind, N. (1982). Aphasia with predominantly subcortical lesion sites. Arch. Neurol.
39, 2-14.
174. Damasio, A.R., Damasio, H., Rizzo, M., Varney, N. and Gersh, F. (1982). Aphasia with
nonhemorrhagic lesions in the basal ganglia and internal capsule. Arch. Neurol. 39, 15-20.
175. Alexander, M.P., Naeser, M.A. and Palumbo, c.L. (1987). Correlations of subcortical CT
lesion sites and aphasia profiles. Brain 110, 961-991.
176. Benson, D.F. (1978). Neurological correlates of aphasia and apraxia. In Recent Advallces ill
Clillical Neurology, Matthews, W.B. and Glaser, G.H. eds., Churchill Livingstone, New
York, pp. 163-175.
177. Weinstein, E.A. and Keller, N.J.A. (1963). Linguistic patterns of misnaming in brain injury.
Neuropsychologia 1, 79-90.
178. Geschwind, N. (1964). Non-aphasic disorders of speech. Int. J. Neurol. 4, 207-214.
179. Geschwind, N. (1967). The varieties of naming errors. Cortex 3,97-112.
180. Chedru, F. and Geschwind, N. (1972). Writing disturbances in acute confusional states.
Neuropsychologia 10, 343-353.
181. Leigh, R.J. and Zee, D.S. (1983). The Neurology of Eye Movements. F.A. Davis Company,
Philadelphia, p. 136.
182. Cummings, J.L. (1985). Organic delusions: phenomenology, anatomical correlations and
review. Br. J. Psychiatry 146, 184-197.
183. Brown, J. W. (1985). Hallucinations. In Handbook oj Clillical Neurology, Vol. 1(45); Clillical
Neuropsychology, Fredericks, J.A.M., ed., Elsevier Science Pub., Amsterdam, pp. 351-372.
184. Levin, H.S., O'Donnell, V.M. and Grossman, R.G. (1979). The Galveston Orientation and
Amnesia Test. J. Nerv. Ment. Dis. 167, 675-684.
185. Artiola, I., Fortuny, 1.., Briggs, M., Newcombe, F., Ratcliff, G. and Thomas, C. (1980).
Measuring the duration of posttraumatic amnesia. J. Neurol. Neurosurg. Psychiatry 43,
377-379.
186. Salazar, A.M., Grafman, J.H., Vance, S.c., Weingartner, H., Dillon, J.D. and Ludlow, C.
(1986). Consciousness and amnesia after penetrating head injury: Neurology and anatomy.
Neurology 36, 178-187.
187. Hillbom, E. and Jarbo, I.. (1969). Posttraumatic Korsakoff syndrome. In The Late Effects of
Head Injury, Walker, A.E., Caveness, W.F., and Critchley, M. eds., Charles C Thomas,
Springfield, IL, p. 98.
188. Salazar, A.M., Grafman,J., Schlesselman, S., Vance, S., Mohr,J.P., Carpenter, M., Pevsner,
P., Ludlow, C. and Weingartner, H. (1986). Penetrating war injuries of the basal forebrain.
Neurology 36, 459-465.
189. Von Wowern, F. (1966). Posttraumatic amnesia and confusion as an index of severity in head
injury. Acta Neurol. Scand. 42, 373-378.
190. Benson, D.F. and Geschwind, N. (1967). Shrinking retrograde amnesia. J. Neurol. Neurosurg. Psychiatry. 30, 539-544.
191. Butters, N. (1979). Amnesic disorders. In Clinical Neuropsychology, Heilman, K. and Valenstein, E., eds., Oxford University Press, New York, pp. 439-474.
192. Teasdale, G. and Brooks, D.N. (1985). Traumatic amnesia. In Handbook ojClillical Neumlogy. Vol. I (45); Clinical Neuropsycholo,~y, Fredericks, J.A.M. ed., Elsevier Science Pub.,
Amsterdam, pp. 185-191.
193. Parkin, A.J. (1984). Amnesic syndrome: A lesion-specific disorder. Cortex 20, 479-508.

3. Issues in Behavioral Neurology

87

194. Benson, D.F. and Blumer, D. (1982). Amnesia: A clinical approach to memory. Psychiatric
Aspects of Neurological Disease, Vol. II, Benson, D.F. and Blumer, D., eds., Grune & Stratton,
New York, pp. 251-278.
195. Crovitz, H.F., Harvey, M.T. and Horn, RW. (1979). Problems in the acquisition of
imagery mnemonics: Three brain-damaged cases. Cortex 15, 225-234.
196. Mayes, A.R, Meudell, P.R. and Pickering, A. (1985). Is organic amnesia caused by a selective deficit in remembering contextual information? Cortex 21, 167-202.
197. Freedman, M. and Cermak, L.D. (1986). Semantic encoding deficits in frontal lobe disease
and amnesia. Brain Cognit. 5, 108-114.
198. Schacter, D.L. and Glisky, E.L. (1986). Memory remediation: restoration, alleviation, and
the acquisition of domain-specific knowledge. In Clinical Neuropsychology ~f illtervelltiotl,
Uzzell, B.P., and Gross, V., eds., Martinus NijhoffPublishing, Boston, pp. 257-282.
199. Stuss, D. T., Alexander, M. P., Lieberman, A. and Levine, H. (1978). An extraordinary form
of confabulation. Neurology 28, 1166-1172.
200. Shapiro, B.E., Alexander, M.P., Gardner, H. and Mercer, B. (1981). Mechanisms of confabulation. Neurology 31, 1070-1076.
201. Benson, D.F., Gardner, H. and Meadows, J.c. (1976). Reduplicative paramnesia. Neurology 26,147-151.
202. Weinstein, A. and Kahn, R.L. (1955). Dellial of Il/lless. Charles C. Thomas, Springfield, IL,
pp.I-166.
203. Alexander, M.P., Stuss, D.T. and Benson, D.F. (1979). Capgras syndrome: A reduplicative
phenomenon. Neurology 29, 334-339.
204. Poeck, K. (1985). Pathological laughter and crying. In Halldbook ofClillical Neurology. Vol. I
(45): Clinical Neuropsychology, Fredericks, J.A.M. ed., Elsevier Science Pub., Amsterdam;
pp. 219-225.
205. Wilson, S.A.K. (1924). Some problems in neurology II: Pathological laughing and crying.
J. Neurol. Psychopathol. 16, 299-333.
206. Tatemichi, T.K., Nichols, F.T. and Mohr, J.P. (1987). Pathological crying: A pontine
pseudobulbar syndrome. Ann. Neurol. 22, 133.
207. Schiffer, RB., Herndon, R.M. and Rudick, R.A. (1985). Treatment of pathologic laughing
and weeping with amitryptyline. N. Engl. J. Med. 312, 1480-1482.
208. Udaka, F., Yamao, S., Nagata, H., Nakamura, S. and Kameyama, M. (1984). Pathologic
laughing and crying treated with levodopa. Arch. Neurol. 41, 1095-1096.
209. Monrad-Krohn, G.H. (1947). Dysprosody or altered "melody of language." Brain 70,
405-415.
210. Ross, E.D. (1981). The aprosodias. Arch. Neurol. 38, 561-569.
211. Ross, E.D. and Mesulam, M.-M. (1979). Dominant language functions of the right hemisphere? Arch. Neurol. 36, 144-148.
212. Heilman, K. M., Scholes, R and Watson, R T. (1975). Auditory affective agnosia. J. Neurol.
Neurosurg. Psychiatry 38, 69-72.
213. Tucker, D.M., Watson, R.T. and Heilman, K.M. (1977). Discrimination and evocation of
affectively intoned speech in patients with right parietal disease. Neurology 27, 947-950.
214. Ross, E.D., Harney,J.H., deLacoste-Utamsing, C. and Purdy, P.D. (1981). How the brain
integrates affective and propositional language into a unified behavioral function. Arch.
Neurol. 38, 745-748.
215. Weintraub, S., Mesulam, M.-M. and Kramer, L. (1981). Disturbances in prosody. Arch.
Neurol. 38, 742-744.
216. Gorelick, P.B. and Ross, E.D. (1987). The aprosodias: Further functional-anatomical
evidence for the organization of affective language in the right hemisphere. J. Neurol.
Neurosurg. Psychiatry 50, 553-560.
217. Bachman, D.L., Bauer, M., Dagge, M. and Long, D. (1985). Validation ofa bedside test of
prosody in right cerebral hemisphere stroke patients. Neurology 35(Suppl. 1), 178.
218. Bowers, D., Coslett, H.B., Bauer, R.M., Speedie, L.J. and Heilman, K.M. (1987). Comprehension of emotional prosody following unilateral hemispheric lesions: Processing defect
versus distracting defect. Neuropsychologia 25, 317-328.
219. Bear, D.M. (1983). Hemispheric specialization and the neurology of emotion. Arch. Neurol.
40, 195-202.
220. Brownell, H.H., Potter, H.H. and Bihrle, A.M. (1986). Inference deficits in right brain

88

I. Theory and Intervention

damaged patients. Brain Lang. 27, 310-321.


221. The American Psychiatric Association. (1980). Diagllostic alld Statistical Manual of Me11tal
Disorders, 3rd ed. APA, Washington, D.C., p. 214.
222. Gans, J S. (1981). Depression diagnosis in a rehabilitation hospital. Arch. Phys. Med. Rehabil. 62, 386-389.
223. Finklestein, S., Benowitz, L.I., Baldessarini, R.J., Arana, G.W., Levine, D., Woo, E., Bear,
D., Moya, K. and Stoll, A.L. (1982). Mood, vegetative disturbance, and dexamethasone
suppression test after stroke. Ann. Neurol. 12, 463-468.
224. Robinson, R.G., Starr, L.B., Kubos, K.L. and Price, T.R. (1983). A two-year longitudinal
study of post-stroke mood disorders: Findings during the initial evaluation. Stroke 14,
736-741.
225. Gainotti, G. (1972). Emotional behavior and hemispheric side of the lesion. Cortex 8, 41-55.
226. Robinson, R.G., Kubos, K.L., Starr, L.B., Rao, K. and Price, T.R (1984). Mood disorders
in stroke patients. Brain 107, 81-93.
227. Sinyor, D.,Jacques, P., Kaloupek, D.G., Becker, R., Goldenberg, M. and Coopersmith, H.
(1986). Poststroke depression and lesion location. Brain 109, 537-546.
228. Starkstein, S.E., Robinson, R.G. and Price, T.R. 1987. Comparison of cortical and subcortical lesions in the production of poststroke mood disorders. Brain 110, 1045-1059.
229. Lipsey, JR. and Robinson, R. G. (1984). Nortriptyline for poststroke depression. Lancet
(April 7), 803.
230. Reding, M.J, Orta, L.A., Winter, S.W., Fortuna, I.M., DiPonte, P. and McDowell, F.H.
(1986). Antidepressant therapy after stroke. Arch. Neurol. 43, 763-765.
231. Finklestein, S.P., Weintraub, R.I., Karmouz, N., Askinazi, c., Davar, G. and Baldessarini,
R.J (1987). Antidepressant drug treatment for poststroke depression: Retrospective study.
Arch. Phys. Med. Rehabil. 68, 772-776.
232. Lishman, W.A. (1978). Or}?allic Psychiatry. Blackwell Scientific Pub., Boston, pp. 234, 361.
233. Moyer, K.E. (1971). The physiology of aggression and the implications for aggression
control. In The Control of A.R.Rressiotl alld Violence: CO}?l1itive and Physiological Factors, Singer,
JL., ed., Academic Press, New York, pp. 61-92.
234. Conn, L.M. and Lion, JR. (1984). Pharmacologic approaches to violence. Psychiatr. Clin.
North Am. 7, 879-886.
235. Devinsky, O. and Bear, D. (1984). Varieties of aggressive behavior in temporal lobe epilepsy. Am. J Psychiatry 141, 651-656.
236. Stevens, JR. and Hermann, B.P. (1981). Temporal lobe epilepsy, psychopathology, and
violence: The state of the evidence. Neurology 31, 1127-1132.
237. Delgado-Escueta, A.V., Mattson, RH., King, L., Goldensohn, E.S., Spiegel, H., Madsen,
J., Crandall, P., Dreifuss, F. and Porter, R.J (1981). The nature of aggression during epileptic
seizures. N. Engl. J Med. 305, 711-716.
238. Monroe, R.R. (1986). Episodic behavioral disorders and limbic ictus. In The Limbic System:
FUllaiOlwl Organizatioll alld Clillical Disorders, Doane, B.K. and Livingston, K.E., eds., Raven
Press, New York, pp. 251-266.
239. Post, RM. (1986). Does limbic system dysfunction playa role in affective illness? In The
Limbic System: FIII/aio/wl Orgallizatioll alld Clil1ical Disorders, Doane B.K. and Livingston,
K.E., eds., Raven Press, New York, pp. 229-249.
240. Treiman, D.M. (1986). Epilepsy and violence: medical and legal issues. Epilepsia 27
(Suppl. 2), S77-104.
241. Rickler, K.C. (1982). Episodic dyscontrol. In Psychiatric Aspects of Neurological disease, Vol. II.
Benson D.F. and Blumer, D., eds., Grune & Stratton, New York, pp. 49-73.
242. Itil, T. and Wadid, A. (1975). Treatment of human aggression with major tranquilizers, antidepressants, and newer psychotropic drugs. J. Nerv. Ment. Dis. 160, 83-99.
243. Jacobs, D. (1983). Evaluation and management of the violent patient in emergency settings.
Psychiatr. Clin. North Am. 6, 259-269.
244. Kellam, A.M.P. (1987). The neuroleptic malignant syndrome, so-called. Br. J Psychiatr.
150, 752-759.
245. Keckish, W. A. (1978). Neuroleptics: Violence as a manifestation of akathisia. JA.M.A. 240,
2185.
246. Siris, S.G. (1985). Three cases of akathisia and "acting out." J Clin. Psychiatr. 46, 395-397.
247. Feeney, D.M., Gonzalez, A. and Law, W.A. (1982). Amphetamine, haloperidol and experience interact to affect rate of recovery after motor cortex injury. Science 217, 855-857.

3. Issues in Behavioral Neurology

89

248. Diamond, B.!., Burrows, E., Hitri, A. and Borison, R. L. (1985). Neuroleptic impairment of
memory as determined by their anticholinergic properties. Neurology 35(Suppl. 1), 179.
249. Rao, N., Jellinek, H.M. and Woolston, D.C. (1985). Agitation in closed head injury:
Haloperidol effects on rehabilitation outcome. Arch. Phys. Med. Rehabil. 66, 30-34.
250. Bond, A. & Lader, M. 1979. Benzodiazepines and aggression. In Psychopharmacology of
Aggression, Sandler, M., ed., Raven Press, New York, pp. 173-182.
251. Gardos, G., DoMascio, A., Salzman, C. and Shader, R.I. (1968). Differential actions of
chlordiazepoxide and oxazepam on hostility. Arch. Gen. Psychiatry 18, 757-760.
252. Lion, JR., Azcarate, c.L. and Koepke, H.H. (1975). "Paradoxical rage reactions" during
psychotropic medication. Dis. Nerv. Sys. 36, 557-558.
253. Sheard, M.H. (1975). Lithium in the treatment of aggression. J Nerv. Ment. Dis. 60,
108-118.
254. Sheard, M.H., Marini, JL., Bridges, c.I. and Wagner, E. (1976). The effect of lithium on
impulsive aggressive behavior in man. Am. J Psychiatry 133, 1409-1413.
255. Dale, P.G. (1980). Lithium therapy in aggressive mentally subnormal patients. Br. J. Psychiatry 137, 469-474.
256. Tyrer, S.P., Walsh, A., Edwards, D.E., Berney, T.R. and Stephens, D. (1984). Factors
associated with a good response to lithium in aggressive mentally handicapped subjects.
Prog. Neuropsychopharmacol. BioI. Psychiatry 8, 751-755.
257. Oyewumi, L.K. and Lapierre, Y.D. (1981). Efficacy of lithium in treating mood disorder
occurring after brain stem injury. Am. J Psychiatry 138, 110-112.
258. Haas, J.F. and Cope, N. (1985). Neuropharmacologic management of behavior sequelae in
head injury: A case report. Arch. Phys. Med. Rehabil. 66, 472-474.
259. Glenn, M.B. and Joseph, A.B. (1987). The use of lithium for behavioral and affective disorders after traumatic brain injury. J Head Trauma Rehabil 2, 68-76.
260. DePaulo, JR. (1984). Lithium. Psychiatr. Clin. North Am. 7, 587-599.
261. Sansone, M.E. G. and Ziegler, D.K. (1985). Lithium toxicity: A review of neurologic complications. Clin. Neuropharmacol. 8, 242-248.
262. Shukla, S., Godwin, L.D., Long, L.E.B. and Miler, M.G. (1984). Lithium-carbamazepine
neurotoxicity and risk factors. Am. J Psychiatry 141, 1604-1606.
263. Elliott, F. A. (1977). Propranolol for the control of belligerent behavior following acute brain
damage. Ann. Neurol. 1, 489-491.
264. Yudofsky, S., Williams, D. and Gorman, J (1981). Propranolol in the treatment of rage and
violent behavior in patients with chronic brain syndromes. Am. J Psychiatry 138, 218-220.
265. Ratey, JJ, Morrill, R. and Oxenkrug, G. (1983). Use of propranolol for provoked and
unprovoked episodes of rage. Am. J Psychiatry 140, 1356-1357.
266. Greendyke, R.M., Schuster, D.B. and Wooten, JA. (1984). Propranolol in the treatment of
assaultive patients with organic brain disease. J Clin. Psychopharmacol. 4, 282-285.
267. Mattes, JA. (1985). Metoprolol for intermittent explosive disorder. Am. J Psychiatry 142,
1108-1109.
268. Mattes,JA., Rosenberg, M.S. and Mays, D. (1984). Carbamazepine versus propranolol in
patients with uncontrolled rage outbursts: A random assignment study. Psychopharmacol.
Bull. 20, 98-100.
269. Greendyke, R.M. and Kanter, D.R. (1986). Therapeutic effects of pindolol on behavioral
disturbances associated with organic brain disease: A double blind study. J. Clin. Psychiatry
47, 423-426.
270. Valzelli, L. (1984). Reflections on experimental and human pathology of aggression. Prog.
Neuropsychopharmacol. BioI. Psychiatry 8, 311-325.
271. Brown, G.L., Ebert, M.H. Goyer, P.F., Jimerson, D.C., Klein, W.J, Bunney, W.E. and
Goodwin, F.K. (1982). Aggression, suicide, and serotonin: Relationships to CSF amine
metabolite. Am. J Psychiatry 139, 741-746.
272. Rudorfer, M. V., Golden, R.N. and Potter, W.Z. (1984). Second-generation antidepressants.
Psychiatr. Clin. North Am. 7, 519-534.
273. Mysiw, W.J and Jackson, R.D. (1987). Tricyclic antidepressant therapy after traumatic
brain injury. J. Head Trauma Rehabil. 2, 34-42.
274. Jackson, R.D., Corrigan, JD. and Arnett, JA. (1985). Amitriptyline for agitation in head
injury. Arch. Phys. Med. Rehabil. 66, 180-181.
275. Simpson, D.M. and Foster, D. (1986). Improvement in organically disturbed behavior with
trazodone treatment. J Clin. Psychiatry 47, 191-193.

90

I. Theory and Intervention

276. Rampling, D. (1978). Aggression: A paradoxical response to tricyclic antidepressants. Am.J.


Psychiatry 135, 117-118.
277. Lercr, B., Moore, N., Meyendorff, E., Cho, S.-R. and Gershon, S. (1985). Carbamazepine
and lithium: Different profiles in affective disorder? Psychopharmacol. Bull 21, 18-22.
278. Post, R. M. and Uhde, T. W. (1986). Anticonvulsants in non-epileptic psychosis. In Aspects of
Epilepsy al1d Psychiatry. Trimble, M.R., and Bolwig, T.G., eds., John Wiley and Sons, New
York, pp. 177-212.
279. Pratt, J.A., Jenner, P., Johnson, A., Shorvon, S.D. and Reynolds, E.H. (1984). Anticonvulsant drugs alter plasma tryptophan concentrations in epileptic patients: Implications
for antiepileptic action and mental function. J. Neurol. Neurosurg. Psychiatry 47, 11311133.
280. McAllister, T. W. (1985). Carbamazepine in mixed frontal lobe and psychiatric disorders.
J. Clin. Psychiatry 46, 393-394.
281. Mattes, J.A. (1984). Carbamazepine for uncontrolled rage outbursts. Lancet 2, 1164-1165.
282. Stringer, A.Y. and Josef, N.C. (1983). Methylphenidate in the treatment of aggression in
two patients with antisocial personality disorder. Am. J. Psychiatry 140, 1365-1366.
283. Blumer, D. Migeon, C. (1975). Hormone and hormonal agents in the treatment of aggression. J. Nerv. Ment. Dis. 160, 127-137.
284. Romano, M.D. (1974). Family response to traumatic head injury. Scand. J. Rehab. Med. 6,
1-4.
285. Caplan, B. and Schechter, J. (1987). Denial and depression in disabling illness. In Rehabilitatioll Psychology Desk R~ferCl/ce, Caplan, B. ed., Aspen Pub., Inc., Rockville, MD, pp. 133170.
286. Cutting, J. (1978). Study of anosognosia. J. Neurol. Neurosurg. Psychiatry 41, 548-555.
287. Bond, M. (1984). The psychiatry of closed head injury. In Closed Head IIl;ury, N. Brooks,
ed., Oxford University Press, New York, p. 167.

4. PSYCHOPHARMACOLOGICAL AGENTS IN
THE TREATMENT OF BRAIN INJURY

GREGORY J. O'SHANICK AND DEAN X. PARMELEE

1 INTRODUCTION

Behavioral disturbances in patients who have sustained brain injury are multiply dctcrmined cvents that rclate to the patients' ncurological status and their
intrinsic adaptational rcsponses, as well as to extrinsic environmcntal factors.
Onc kcy intcrvcntion with thesc individuals has becn thc usc of psychotropic
agcnts, which cxert their therapeutic effects through alteration of ncurotransmitters in the central nervous system (eNS). At thc samc timc, studies
of paticnts who have not sustained head injuries arc providing evidence that
links alterations in neurotransmitters to affective disorders, psychosis, aggression, irritability, and memory functioning. Although thesc behavioral changes
may reprcsent a final common pathway in terms of hcterogcneous causes,
psychotropic agents may help correct these changes.
By understanding the pathophysiology of ncurobehavioral abnormalitics
secondary to brain injury and the ncuropsychopharmacology of psychotropics,
thc clinician can choose the appropriate psychotropic medications to ameliorate
the patients' disabling symptoms. In this chapter, each of the major classes of
psychotropic agcnts will be reviewed in alphabetical order, with particular
reference to thcir use for brain-injurcd patients. Because children and adolesccnts incur such a great numbcr of the serious brain injuries, we have noted
any applications that are useful for the treatment of this population.
2 ANTICONVULSANTS

Anticonvulsants are typically used following traumatic brain injury to prcvent


the occurrence of posttraumatic seizures. Howevcr, prophylaxis of posttrau91

92

I. Theory and Intervention

matic seiZures is not the focus of this section; information on this subject can be
found in a review by Deutschman and Haines [1]. Rather, it is the behavioral
side effects of these drugs that we will be discussing.
The behavioral complications of anticonvulsants are well known [2]. For
example, phenobarbital is associated with drowsiness, dizziness, ataxia, dysarthria, excitation, and increase in activity. There is evidence that long-term
administration of phenobarbital can reduce cognitive functioning [3]. Benzodiazepines have also been reported to have significant behavioral side effects.
Aggression, anorexia, depression, sedation, paradoxical excitation, and irritability have been reported; these will be described in more detail in the section
on benzodiazepines. Phenytoin has also been implicated in behavioral symptoms such as anorexia, dementia, hyperactivity, aggressive behavior, and restlessness. As with the use of phenobarbital, recent evidence has confirmed
the development of cognitive impairment in children following long-term
phenytoin administration [3]. Carbamazepine has been shown to cause
anorexia, dysphoria, sedation, and psychosis. However, carbamazepine does
not appear to have the cognitive-impairing potential that has been associated
with other anticonvulsants [4].
Carbamazepine has been used widely in the treatment of patients who have
organic mental disorders, including mixed frontal lobe syndromes [5], KliiverBucy syndrome [6A], and aggressive behavior following traumatic brain
injury [6B]. Often, these patients do not have abnormal electroencephalograms
(EEGs) [7]. Carbamazepine is sometimes used to treat behavioral disorders of
children, when nothing else has proved effective [8]. However, controlled
studies have not yet been conducted on its use with either behaviorally dysfunctional brain-injured children or nonorganically impaired, behaviorally
dysfunctional children.
Hematological difficulties may be encountered in the patient who is being
treated with carbamazepine; however, much of the earlier concern has receded
in the face of more recent data [9, 10]. Careful monitoring of the patients'
hematological indices is essential to prevent the evolution of full-blown aplastic
anemia; this condition may be averted if bone-marrow suppression is detected
early.
3 ANTIDEPRESSANTS

Antidepressants have shown considerable promise in the treatment of behavioral problems following brain injury. Although antidepressants have
classically been used in treating postconcussive syndromes following mild
head injury [11, 12], recent evidence suggests that these agents may also be
useful in treating agitation [13], aggressive behavior [14], posttraumatic sleep
disturbance [15], and posttraumatic stress disorders [16]. With children? antidepressants have been found useful in treating attention deficit syndromes [17]
as well as depression, and they should be considered safe, effective treatment
options for some of the behavior and affect problems that occur postinjury

4. Psychopharmacologic

A!n~s

93

[18]. Dosages for children often have to be higher than adult dosages because
of children's increased metabolism. Careful monitoring of blood levels can
prevent toxicity that can lead to seizures [19].
Antidepressants most likely exert their therapeutic action by inhibiting the
reuptake of neurotransmitters at the presynaptic level, thus increasing their
availability to the postsynaptic membrane [20, 21]. The major neurotransmitters involved in this process are norepinephrine and serotonin, although
recent evidence suggests that the anticholinergic function of antidepressants
may be equally responsible for their effect [22]. Antidepressants exert multiple
influences on other neurotransmitter systems, including alpha- and betaadrenergic systems, histaminic systems, and dopaminergic systems. These
interactions result in side effects that can limit their efficacy [21].
Antidepressants that amplify serotonergic mechanisms in the CNS are most
frequently cited as being effective for treating brain-injured patients. Amitriptyline has been shown to be useful for decreasing agitated behavior following frontallobeinjury [13]. Trazodone has recently been reported to substantially
improve agitated behavior and hypersexuality in geriatric patients who have
severe organic brain syndromes [14]. Posttraumatic night terrors have been
shown to be responsive to imipramine, an agent with both serotonergic and
noradrenergic augmenting properties [15, 16].
As mentioned, significant side effects may be encountered during the use of
antidepressants. Some of these are especially important to take into account
in the treatment of head-injured patients. For example, the ability that these
agents have to interact with muscarinic cholinergic receptors may exacerbate
cognitive impairment. Recent studies note the lack of cognitive impairment in
geriatric patients who have been treated with antidepressants that have lower
anticholinergic profiles (such as trazodone), rather than with ones that have
higher anticholinergic profiles (such as amitriptyline) [22, 23]. In addition,
antidepressants may interfere with anticonvulsant levels because of competitive protein binding; this may induce either toxicity or breakthrough seizures
[24]. The potential that such side effects have for inducing problems for the
brain-injured patient is readily apparent.
4 BENZODIAZEPINES

The use of benzodiazepines in the treatment of behavior problems following traumatic brain injury relies on the sedative and anxiolytic effects that
are generated by these substances' ability to enhance the activity of gammaaminobutyric acid (GAB A) [25]. (Although benzodiazepines have been used to
control spasticity, seizure activity, and tremor, such usage will not be considered in this section.) There is some evidence that benzodiazepines are useful
for treating mentally retarded psychotic patients [26], b\1t no evaluations of the
efficacy of benzodiazepines for the behavioral management of head-injured
patients have been conducted.
Evidence has been accumulating about a possible association between lora-

94

I. Theory and Intervention

zepam and triazolam (two short-acting benzodiazepines) and the development


of anterograde amnesia [27, 28]. Diazepam has been found to disrupt aspects
of attention and episodic memory without disrupting retrieval of previous
learning and long-term memory [29]. Benzodiazepines with longer half-lives,
such as clorazepate, have not demonstrated significant amnesic effects in either
young or geriatric patients [28, 30]. Clearly, disruption of memory processes
by these agents in the recovering head-injury patient could contribute to the
deterioration of cognitive skills.
Reports in the anecdotal literature support the existence of an additional
adverse effect-the development of hostile behavior in patients-following
benzodiazepine administration [31-35]. Still more complications include
patients' tolerance for these agents and their withdrawal symptoms when the
medications are discontinued [36]. It has been estimated that withdrawal may
emerge in 15% to 44% oflong-term benzodiazepine users [37]. This abstinence
phenomenon may be more acute and dramatic in patients who are given benzodiazepines that have extremely short half-lives, such as triazolam and lorazepam; in these patients, abstinence phenomena may, in fact, be mistaken for
worsening of anxiety and agitation. Use ofbenzodiazepines with brain-injured
patients iqnot advisable until more is known about the adverse effects of these
agents.
5 BETA-ADRENERGIC ANTAGONISTS

In 1977, the so-called "beta blockers" were first reported to be efficacious in


the treatment of episodic aggressive behavior in patients who had acute brain
damage. Elliot found that rage, irritability, and belligerence diminished significantly after patients had been treated with propranolol [38]. Since that time,
numerous studies and anecdotal reports have indicated that beta-adrenergic
blockade is effective in the treatment of aggression related to severe mental
retardation and brain injury from trauma, seizure disorder, hepatolenticular
degeneration [39], and infection [40]. In the adolescent psychiatric population,
propranolol has been used successfully to treat difficult-to-manage patients
who have had brain damage from many causes, including head injury [41].
Researchers have also found positive results following propranolol use for
patients who have dementia [42] and Korsakoff's syndrome [43].
In both animal studies [44] and human studies [45], it has been noted that
elevations in noradrenergic activity coincide with some types of aggressive
behavior, although the precise mechanism of action is unproved. When the
elevation of noradrenergic activity has been inhibited, the aggressive behavior
diminished.
Patients who have pulmonary disease, type I diabetes mellitus, congestive
heart failure, and hypothyroidism should not be treated with beta blockers
[46]. Careful monitoring of pulse rate is required to detect profound bradycardia (fewer than 50 beats per minute) and reduce dosage. Beta blockers may
elevate blood levels of other psychotropic agents. They have also been reported

4. Psychopharmacologic Agents

95

to cause impotence and secondary depressive syndromes [47]. Despite these


side effects and contraindications, beta blockade does appear to be one alternative for combating agitated, aggressive behavior among brain-injured patients.
6 CALCIUM CHANNEL BLOCKERS

The use of agents that antagonize calcium's role in neurotransmitter release


has recently been discussed as a method of treatment in post-head-injury impulsivity and self-abuse [48]. Calcium's role in neurotransmitter release has
been well-described [49], and its use in head-injured patients is only now being
explored. Further work is needed on this potentially beneficial strategy to
diminish aggressive, impulsive behaviors following head injury.
7 CHOLINERGIC AGENTS

Agents that influence cholinergic pathways of the CNS have been the subject
of two lines of investigation in brain-injured patients. In the acute setting,
evidence implicates excess cholinergic activity as an etiological factor in coma
[50, 51]. Studies in which animals have been treated with anticholinergic agents
before head injury find shortened duration of coma [51]. No human data exist
on this use at this time. However, there is evidence from one small-scale study
that scopolamine, an anticholinergic agent, is useful in treating the syndrome of
tactile defensiveness following head injury [52]. Further investigation continues
in these areas.
Cholinergic involvement in memory function has been well-demonstrated
in patients with senile dementia of the Alzheimer's type [53]. Researchers feel
that such patients' memory defects in retrieval and registration are the result of
deficiencies in the CNS's cholinergic system, which occur in this syndrome.
Treatment strategies have suggested the use of choline, lecithin, or physostigmine to enhance central cholinergic levels and thereby improve cognition [54,
55]. Results have been disappointing so far, although ol,le report cited dramatic
improvement in performance IQ after physostigmine was administered to an
individual with brain injury [56].
A more practical aspect of this area of research is the ability ofmedication
to exert. profound effects on muscarinic receptors in the CNS [57]. Varying
affinities to muscarinic receptors have been demonstrated for neuroleptics,
antidepressants, and anticonvulsants [58-60]. Anticholinergic side effects can
then be predicted, based upon these binding affinities. For example, studies
have shown evidence of cognitive impairment in psychiatric patients who have
been exposed to anticholinergic agents for extended periods of time [57,59].
Furthermore, combinations of medications that exert .anticholinergic side
effects can induce significant delirium [61]. A general strategy has been proposed that minimizes exposure to anticholinergic agents or agents with strong
anticholinergic properties to minimize cognitive problems and the potential
for delirium following head injury [62].

96

I. Theory and Intervention

8 LITHIUM

8.1 Usage

Lithium has been used successfully to treat agitation and hyperactivity in


patients with organic mental disorders. For this reason, it has been utilized in
treating agitation and aggressive behavior in individuals following traumatic
brain injury [63, 64]. For example, lithium carbonate has been reported to
be helpful for brain-injured adolescents who develop impulsivity and mood
lability during the often protracted rehabilitation process (R. Kowatch and
D.X. Parmelee, The use oflithium carbonate for behavioral sequelae of closed
head injury in adolescents, unpublished.) There is also a case report of reversal
of retrograde posttraumatic amnesia with lithium [65].
8.2 Metabolic Activity

The mechanism of action of lithium is extremely broad. Lithium has been


shown to enhance blood-brain barrier permeability when cerebral blood flow
is low [66]. There is also evidence of an increase in amino acid concentration in
the cerebrospinal fluid (CSF) with lithium treatment [67]. From a neurotransmitter standpoint, lithium induces decreased dopamine formation while blocking presynaptic and postsynaptic dopamine receptor sensitivity [68,69]. Lithium
also increases norepinephrine neuronal uptake while increasing urinary excretion of norepinephrine metabolites [70, 71]. Evidence also exists that lithium
has a role in decreasing beta-receptor binding and in decreasing the density of
beta receptors centrally [72, 73].
Of interest relative to agitation and aggression is lithium's ability to increase
brain concentrations of tryptophan, stimulate 5-hydroxytryptophan (5HT)
synthesis, and increase levels of 5-hydroxy-indoleacetic acid (5HIAA) in the
CSF [68, 74-76]. Aggression and agitated behavior have been linked with
lowered serotonergic activity in experimental studies with humans and animals
[77]. For that reason, one could speculate that lithium's ability to decrease
agitated, aggressive behavior is secondary to its neurotransmitter effects.
8.3 Side Effects and Contraindications

Lithium use has toxic complications [79], including gastrointestinal symptoms


such as anorexia, nausea, and vomiting, along with excessive thirst. Polyurea
may be seen as a response to the solute load. Toxicity is often associated with
reversible EEG alteration, and seizures have been reported in patients who
have been in severe intoxicated states. Lithium may also interfere with thyroid
function, resulting in decrease of free thyroxine. Some nephrotoxicity has
been associated with chronic lithium administration [78]. In addition, lithium
has multiple interactions with many drugs [79]. Adverse interactions between
lithium and anticonvulsants have been reported. Specifically, carbamazepine
and lithium in combination have been reported to induce a neurotoxic syn-

4. Psychopharmacologic Agents

97

drome when administered to adolescents following head injury [6B]. Further


investigation into this combination is certainly indicated.
Treatment guidelines for using lithium in brain-injured patients include
careful monitoring of electrolyte status and hematological status, monitoring
fluid intake and output to prevent dehydration and toxicity, maintaining blood
levels at approximately 0.5 m Eq/L to 1.1 m Eq/L, and monitoring blood levels
weekly or every two weeks to prevent significant fluctuations in therapeutic
levels due to drug interaction [78, 80].
9 NEUROLEPTICS

9.1 Usage

Neuroleptics have been widely used to control agitation and aggressive behavior in individuals who have organic mental disorders. Although the preponderance of experience reported in the literature centers upon either patients
with metabolic encephalopathies (chiefly delirium) [81] or elderly patients with
dementia [82], the use of neuroleptics to control behavioral outbursts following brain injury had been widely accepted until 1982. At that time, however, a
report of a study that explored the interaction of amphetamine and haloperidol
relative to rate of recovery following brain injury in rodents was published:
That study demonstrated impairment in recovery among those animals exposed
to haloperidol [83]. That observation suggested that use of dopamine-blocking
agents may have a negative effect with respect to human recovery from traumatic brain injury. A retrospective study of patients who received haloperidol
to control post-brain-injury behavioral disruption found no difference between
haloperidol-treated or nontreated patients relative to success of rehabilitation
outcome, although duration of posttraumatic amnesia was significantly longer
for patients who were treated with haloperidol [84].
9.2 Side Effects and Contraindications

Neuroleptic use is further limited by the dramatic side effects that may be
produced by these agents [85]. Extrapyramidal side effects can include acute
dystonic reactions, akathesia, pseudo parkinsonism, and tardive dyskinesia.
Tardive dyskinesia is an irreversible movement disorder that is generally oralfacial in location; however, it has also been reported as a more generalized
trunk phenomenon. No effective treatment has yet been developed for this
condition [86]. Because underlying neurological disorder is a predisposing
factor for the development of tardive dyskinesia, the risk of its occurring in the
brain-injured individual would seem to be disproportionally high. However,
no definitive study has yet been performed to test that hypothesis. Neuroleptics
can also have the effect of lowering seizure thresholds; the most commonly
implicated antipsychotic agent is chlorpromazine [87].
Another significant complication in the use of neuroleptics is neuroleptic
malignant syndrome [88]. This potentially fatal idiosyncratic reaction is charac-

98

I. Theory and Intervention

terized by muscular rigidity, fever, autonomic dysfunction, and altered consciousness; leukocytosis and markedly elevated serum creatinine phosphokinase
levels are also commonly seen. (This syndrome may be easily overlooked in
the multiply traumatized patient with head injury, since those signs and symptoms could be attributed to other causes.) Current treatment strategies involve
discontinuing all neuroleptics and providing general supportive measures,
including hydration, adequate nutrition, and reduction of fever.
10 NEUROPEPTIDES

Attempts to enhance memory deficits by utilizing neuropeptides have included


the use of vasopressin [89] and adrenocorticotropic hormone (ACTH) [90].
The preponderance of work has been with geriatric populations, and little
effect has been noted upon memory and new learning overall. Geriatric volunteers with memory impairment showed no consistent improvement in their
performance on various memory tests that could be attributed to ACTH [90].
Similarly, in a population of brain-injured patients, those who used vasopressin
nasal spray showed no significant difference in memory and learning performance over those who used a placebo. Anecdotal reports of improvement
may relate to these agents' actions as reinforcement mechanisms or attention/
arousal augmentation.
11 NEUROTRANSMITTERS

Direct introduction of neurotransmitters in the treatment of patients with


brain injuries has centered upon studies that add exogenous L-dopa. Anecdotal
clinical reports of improvement in children with progressive neurological disorders after the introduction of L-dopa [91], as well as animal studies indicating
that small doses of L-dopa reversed certain types of brain damage of an acute
nature [92], spurred investigation of the use of L-dopa in the treatment of
patients with severe brain injury. The most complete study was of inpatients
who were in a persistent vegetative state. In one series, clinical improvement
was noted in 29 of 45 patients who received L-dopa treatment [93]. Further
investigation is under way.
Other neurotransmitters such as tryptophan have not been studied in the
acute head-trauma phase, primarily because of their inability to cross the bloodbrain barrier [94]. Uptake of tryptophan (an amino acid) is shared with other
branched-chain amino acids that are preferentially taken up in acute brain injury
[95]. For this reason, methods of augmenting neurotransmitter levels through
the administration of precursor amino acids may not be as beneficial a treatment modality as originally hoped.
12 STIMULANTS

Agents that augment release or increase the availability of catecholamines


(norepinephrine and dopamine) in the CNS have long been used in the treatment of attention deficit disorder and minimal brain dysfunction in children
and adults. In these syndromes, significant improvement has been achieved in

4. Psychopharmacologic Agents

99

areas such as sustained attention and memory, and decreases have been noted
in hyperactivity and interpersonal intrusiveness [96-98]. Stimulants have also
been used to treat secondary depression in medically ill adult and geriatric
patients [99, 100]. There is controversy about the long-term efficacy of these
agents in patients with affective disorders, but short-term administration has
been shown to be beneficial [101, 102].
In our discussion of neuroleptics, we cited a study that also demonstrated
that methylphenidate administration significantly improved relearning of a
specific paradigm by laboratory animals following traumatic brain injury
[83]. Anecdotal evidence exists of improvement in patients with posttraumatic
confusion, paranoia, and short-term memory deficits as a result of stimulant
treatment [103, 104].
In children, common sequelae of serious brain injuries include inattention,
distractibility, impulsivity, and hyperactivity [105], all quite problematic,
espeeially in the school environment. Treatment of these behaviors with stimulants is well-documented for the child psychiatric population, and similar
drug choices and dosages are indicated for the head-injured child before other
psycho tropics are tried. Appetite suppression with some weight loss is the
most common side effect in children.
Significant toxic manifestations can occur with the use of psycho stimulants
[106]. For example, paranoid misinterpretations and inappropriate social behavior may be seen early in treatment; these may progress to frank delusional
states that include hallucinations and delirium. This may be seen in either acute
toxic or acute abstinence (withdrawal) states.
The use of any psychostimulants should be carefully monitored, since these
agents have high abuse potential. Treatment should be initiated in inpatient
settings, with controlled observations to assure that any positive therapeutic
response is not due to placebo effects [107]. Patients should be periodically
reevaluated when they are no longer receiving psychostimulants to determine
whether their medication should be continued.
13 CONCLUSION

As psychiatry has returned to its biological underpinnings, especially behavioral neurology and psychopharmacology, progress has been made in helping those who have survived serious brain injury. Effective rehabilitation of
the neuropsychiatric sequelae of brain injury includes psychopharmacological
interventions that are well-coordinated with other services [108]. Considerable
research is still needed in this field so that advances in rehabilitation medicine
can parallel those in acute care medicine and surgery.
REFERENCES
1. Deutschman, C. S. and Haines, S.]. (1985). Anticonvulsant prophylaxis in neurological
surgery. Neurosurgery 17, 510-516.
2. Rivinus, T.M. (1982). Psychiatric effects of anticonvulsant regimens. ]. Clin. Psychopharmaco!. 2, 165-192.
3. Pellock, ].M. (1986). The role of anti-seizure medications on cognitive function. In Ahstracts

100

4.
5.
6A.
6B.

7.
8.

9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.

I. Theory and Intervention

of the Proceedings, 10th Annual Medical College of Virginia Conference on the Rehabilitation
of the Head-Injured Child and Adult, Williamsburg, VA.
Thompson, P.J. and Trimble, M.R. (1982). Anticonvulsant drugs and cognitive functions.
Epilepsia 23, 531-544.
McAllister, T.W. (1985). Carbamazepine in mixed frontal lobe and psychiatric disorders. J.
Clin. Psychiatry 46, 393-394.
Stewart, J.T. (1985). Carbamazepine treatment of a patient with Kluver-Bucy syndrome. J.
Clin. Psychiatry 46, 496-497.
Parmelee, D.X. and O'Shanick, G.J. (1988). Carbamazepine-lithium toxicity in braindamaged adolescents. Brain Injury 2 (4), 305-308.
Dalby, M.A. (1975). Behavioral effects ofcarbamazepine. Adv: Neurol. 11,331-344.
Remschmidt, H. (1976). The psychotropic effect of carbamazepine in nonepileptic patients,
with particular reference to problems posed by clinical studies in children with behavioral
disorders. In Epileptic Seizures-Behavior-Paill, Birkmayer, W. ed., Hans Huber, Berne,
pp. 235-258.
Schain, R.J., Ward, J.W. and Guthrie, D. (1977). Carbamazepine as an anticonvulsant in
children. Neurology 27, 476-480.
Schmidt, D. (1982). Adverse E,[fects of Alltiepileptics, Raven Press, New York.
Thompson, G.N. (1977). Post-traumatic psychoneurosis: Evaluation of drug therapy. Dis.
Nerv. Sys. 38, 617-619.
Tyler, G.S., McNeely, H.E. and Dick, M.L. (1980). Treatment of post-traumatic headache
with amitriptyline. Headache 20, 213-216.
Jackson, RD., Corrigan, J.D. and Arnett, J.A. (1985). Amitriptyline for agitation in head
injury. Arch. Phys. Med. Rehabii. 66, 180-181.
Simpson, D.M. and Foster, D. (1986). Improvement in organically disturbed behavior with
trazodone treatment. J. Clin. Psychiatry 47, 191-193.
Marshall, ]. R (1975). The treatment of night terrors associated with the posttraumatic
syndrome. Am.]. Psychiatry 132, 293-295.
Falcon, S., Ryan, c., Chamberlain, K. and Curtis, G. (1985). Tricyclics: Possible treatment
for posttraumatic stress disorder. J. Clin. Psychiatry 46, 385-389.
Gastfriend, D.R, Biederman, J. and Jellinek, M. S. (1984). Desipramine in the treatment of
adolescents with attention deficit disorder. Am. J. Psychiatry 141, 906-908.
Parmelee, D.X. and O'Shanick, G.]. (1987). Neuropsychiatric interventions with head
injured children and adolescents. Brain Injury 1, 41-47.
Popper, C.W. (1985). Child and adolescent psychopharmacology. In Psychiatry, Cavenar,
J.O., ed., ].B. Lippincott and Co., Philadelphia, pp. 1-23.
Axelrod, J. (1978). Central catecholamine neurotransmitters and psychoactive drugs. In
Neurotransmissio/l a/ld Disturbed Behavior, Praag, H.M. van and Bruinvels,]., eds., SP Medical
and Scientific Books, New York, pp. 6-18.
Richelson, E. and Nelson, A. (1984). Antagonism by antidepressants of l1eurotransmitter
receptors of normal human brain in vitro. J. Pharmacol. Exp. Ther. 230, 94-102.
Moskowitz, H. and Burns, M.M. (1986). Cognitive performance in geriatric subjects after
acute treatment with antidepressants. Neuropsychobiology 15, 38-43.
Burns, M., Moskowitz, H. and Jaffe, J. (1986). A comparison of the effects of trazodone and
amitriptyline on skills performance by geriatric subjects. J. Clin. Psychiatry 47, 252-254.
Dorn, J.M. (1986). A case of phenytoin toxicity possibly precipitated by trazodone.]. Clin.
Psychiatry 47, 89-90.
Ameer, B. and Greenblatt, D.]. (1981). Lorazepam: A review of its clinical pharmacological
properties and therapeutic uses. Drugs 21, 161-200.
Saklad, S.R., Ereshefsky, L., Jann, M.W., Ward, M.E., Richards, A.L. and Birkhimer, L.J.
(1985). Us~ful/less ~f //liectable a/ld Oral Lorazepam ill Psychotic and Developmelltally Disabled
Patie/lts, Psychiatric Pharmacy Program, College of Pharmacy, University of Texas, Austin.
Mac, D.S., Kumar, R. and Goodwin, D.W. (1985). Anferograde amnesia with oral lorazepam. In New Research: Scielltific Proceedings ~fthe 138tll AtHmal Meetillg of the Americau Psychiatric Associatio/l. APA, Washington, DC, p. 49.
Scharf, M.B., Hirschowitz, J., Woods, M. and Scharf, S. (1985). Lack of amnesic effects of
c10razepate on geriatric recall. ]. Clin. Psychiatry 46, 518-520.
Wolkowitz, O.M., Weingartner, H. and Hommer, D.W. (1985). Specificity ofbenzodiaze-

4. Psychopharmacologic Agents

101

pine disruption of memory. In New Research: Sciel1tific Proceedil1gs of the 138th Alll1ual Meetillg
American Psychiatric Association. APA, Washington, DC, p. 48.
30. Healey, M., Pickens, R, Meisch, R. and McKenna, T. (1983). Effects of clorazepate, diazepam, lorazepam, and placebo on human memory.]. Clin. Psychiatry 44, 436-439.
31. Rickels, K. and Downing, R W. (1974). Chlordiazepoxide and hostility in anxious outpatients. Am]. Psychiatry 131, 442-444.
32. Barron, ]. and Sandman, C.A. (1985). Paradoxical excitement to sedative-hypnotics in
mentally retarded clients. Am. ]. Ment. Defic. 90, 124-129.
33. Kochansky, G.E., Salzman, C., Shader, R.I., Harmatz, ].S. and Ogletree, A.M. (1977).
Effects of chlordiazepoxide and oxazepam administration on verbal hostility. Arch. Gen.
Psychiatry 34, 1457-1459.
34. Hall, RC.W. and Zisook, S. (1981). Paradoxical reactions to benzodiazepines. Br.]. Clin.
Pharmacol. 11, 99S-104S.
35. Rosenbaum, ].F., Woods, S.W., Groves, ].E. and Klerman, G.L. (1984). Emergence of
hostility during alprazolam treatment. Am. ]. Psychiatry 141, 792-793.
36. Zisook, S. and DeVaul, R.A. (1977). Adverse behavioral effects ofbenzodiazepines. J. Fam.
Pract. 5, 963-966.
37. Lader, M. and Higgitt, A. (1986). Management of withdrawal from benzodiazepines. lnt.
Drug Ther. Newslett. 21, 21-23.
38. Elliott, F. A. (1977). Propranolol for the control of belligerent behavior following acute brain
damage. Ann. Neurol. 1, 489-491.
39. Schreier, H.A. (1979). Use of propranolol in the treatment of postencephalitic psychosis.
Am.]. Psychiatry 136, 840-841.
40. Yudofsky, S.c., Williams, D. and Gorman,]. (1981). Propranolol in the treatment of rage
and violent behavior in patients with chronic brain syndrome. Am. ]. Psychiatry 138,
218-220.
41. Williams, D.T., Mehl, R., Yudofsky, S., Adams, D. and Roseman, B. (1982). The effect of
propranolol on uncontrolled rage outbursts in children and adolescents with organic brain
dysfunction.]. Am. Acad. Child Psychiatry 21,129-135.
42. Petrie, W.M. and Ban, T.A. (1981). Propranolol in organic agitation. Lancet 1, 324.
43. Yudofsky, S.c., Stevens, L., Silver,]., Barasa,]. and Williams, D. (1984). Propranolol in
the treatment of rage and violent behavior associated with Korsakoff's psychosis. Am. ].
Psychiatry 141, 114-115.
44. Elliott, F.A. (1976). The neurology of explosive rage. Practitioner 217,51-59.
45. Brown, G.L., Ebert, M.H., Goyer, P.F., Jimerson, D.C., Klein, W.]., Bunney, W.E. and
Goodwin, F.K. (1982). Aggression, suicide, and serotonin relationships to CSF amine
metabolites. Am. J. Psychiatry 139, 741-746.
46. Silver, ]. M. and Yudofsky, S. 1985. Propranolol for aggression: Literature review and
clinical guidelines. Int. Drug Ther. Newslett. 20, 9-12.
47. Griffin, S.]. and Friedman, M.]. (1986). Depressive symptoms in propranolol users.]. Clin.
Psychiatry 47, 453-457.
48. Childs, A. (1986). Calcium channel blockers in psychiatry. In Physicialls Executive Network,
Health Care International Press, Austin, TX, pp. 10-12.
49. Cooper, ]., Bloom, F.E. and Roth, R.H. (1982). Cellular foundations of neuropharmacology: Transmitter secretion. In The Biochemical Basis of Neuropharmacology, Cooper,
].R, Bloom, F.E. and Roth, RH., eds., Oxford Press, New York, pp. 33-35.
50. Hayes, R.L., Pechura, C.M., Katayama, Y., Povlishock, ].T., Giebel, M.L. and Becker,
D.P. (1984). Activation of pontine cholinergic site implicated in unconsciousness following
cerebral concussion in the cat. Science 223, 301-303.
51. Hayes, R.L., Stonnington, H.H., Lyeth, B.G., Dixon, C.E. and Yamamoto, T. (1986).
Metabolic and neurophysiologic sequelae of brain injury: A cholinergic hypothesis. Cent.
Nerv. Sys. Trauma 3,163-173.
52. Childs, A. (1986). Scopolamine effects in vestibular defensiveness. Arch. Phys. Med.
Rehabil. 67, 554-556.
53. Coyle, J.T., Price, D.L. and DeLong, M.R (1983). Alzheimer's disease: A disorder of
cortical cholinergic innervation. Science 219, 1184-1190.
54. Brinkman, S.D., Pomara, N., Goodnick, P.]., Barnett, N. and Domino, E.F. (1982). A
dose-ranging study oflecithin in the treatment of primary degenerative dementia (Alzheimer
~fthe

102

I. Theory and Intervention

disease).]. Clin. Psychopharmacol. '2, 281-285.


55. Jenike, M.A., Albert, M.S., Heller, H., LoCastro, S. and Gunther,]. (1986). Combination
therapy with lecithin and ergoloid mesylates for Alzheimer's disease.]. Clin. Psychiatry 47,
249-251.
56. Goldberg, E., Gerstman, ].L., Mattis, S., Hughes,].E., Bilder, R.M. and Sirio, C.A. (1982).
Effects of cholinergic treatment on posttraumatic anterograde amnesia. Arch. Neurol. 39,
581.
57. Katz, I.R., Greenberg, W.H., Barr, G.A., Garbarino, c., Buckley, P. and Smith, D. (1985).
Screening for cognitive toxicity of anticholinergic drugs.]. Clin. Psychiatry 46, 323-326.
58. Snyder, S.H., Greenberg, D. and Yamamura, H.1. (1974). Antischizophrenic drugs and
brain cholinergic receptors. Arch. Gen. Psychiatry 31, 58-61.
59. Tune, L.E., Strauss, M.E., Lew, M.F., Breitlinger, E. and Coyle,]. T. (1982). Serum levels
of anticholinergic drugs and impaired recent memory in chronic schizophrenic patients. Am.
]. Psychiatry 139, 1460-1462.
60. Peroutka, S.]. and Snyder, S.H. (1980). Relationship of neuroleptic drug effects and brain
dopamine, serotonin, adrenergic, and histamine receptors to clinical potency. Am. J. Psychiatry 137, 1518-1522.
61. Tune, L.L. Damlouji, N.F., Holland, A., Gardner, T.]., Folstein, M.F. and Coyle, ].T.
(1981). Association of post-operative delirium with raised serum level of anticholinergic
drugs. Lancet 2, 651-653.
62. O'Shanick, G.j. (1988). Psychotropic management of behavioral disorders after head
trauma. Psychiatr. Med. 6, 67-82.
63. Haas, ].F. and Cope, N. (1985). Neuropharmacologic management of behavior sequelae in
head injury: A case report. Arch. Phys. Med. Rehabil. 66, 472-474.
64. Gergersen, A.B. and Hagen, S. (1983). Sekundaer mani efter cerebralt traume. Ugeskr
Laeger 145, 3361.
65. Kline, N.A. (1979). Reversal of post-traumatic amnesia with lithium. Psychosomatics 20,
363-364.
66. Preskorn, S.H., Irwin, G.H., Simpson, S., Friesen, D., Rinne,]. and Jerkovich, G. (1981).
Medical therapies for mood disorders alter the blood-brain barrier. Science 213, 469-471.
67. Goodnick, P.]., Fieve, R.R., and Dunner, D .. (1981). Factors in time-course of lithium
response: Pharmacokinetics and CSF parameter changes. Abstract, Third World Congress of
Biological Psychiatry, Stockholm.
68. Engel,]. and Berggren, V. (1980). Effects oflithium on behavior and central monoamines.
Acta Psychiatr. Scand. 280(Suppl.), 133-143.
69. Flekmenbaum, A. (1977). Lithium inhibition of norepinephrine and dopamine receptors.
BioI. Psychiatry 12, 563-572.
70. Schildkraut, ].]. (1973). Pharmacology-the effects of lithium on biogenic amines. In
Lithium: Its Role ill Psychiatric Research alld Treatmellt, Gershon, S. and Shopsin, B. eds.,
Plenum Press, New York, pp. 51-73.
71. Greenspan, K., Schildkraut, ].]., Gordon, E.K., Baer, L., Aronoff, M.S. and Durell, ].
(1970). Catecholamine metabolism in affective disorders. 3. MHPG and other catecholamine
metabolites patients treated with lithium carbonate. ]. Psychiatr. Res. 7, 171-183.
72. Rosenblatt, ].E., Pert, C.B., Tallman, J.R., Pert, A. and Bunney, W.E. (1979). The effect
of imipramine and lithium on alpha- and beta-receptor binding in rat brain, Brain Res. 160,
186-191.
73. Rosenblatt, ].E. (1981). Neurobiologic mechanisms of lithium action. Sciet1tijic Proceedillgs
of the 134th All/lual Meetillg ~f the Americall Psychiatric Associatioll. APA, Washington, DC,
p. 160.
74. Tagliamonte, A., Tagliamonte, P., Perez-Cruet,]., Stern, S. and Gessa, G.L. (1979). Effect
of psychotropic drugs on tryptophan concentration in the rat brain. ]. Pharmacol. Exp.
Ther. 177, 475-480.
75. Perez-Cruet,]., Tagliamonte, A. and Tagliamonte, P. (1981). Stimulation of serotonin synthesis by lithium. ].Pharmacol. Exp. Ther. 178, 325-330.
76. Ho, A.K., Loh, H.H. and Craves, F. (1970). The effect of prolonged lithium treatment on
the synthesis rate and turnover of monoamines in brain regions of rats. Eur.]. Pharmacol.
10, 72-78.
77. Greenberg, A.S. and Coleman, M. (1976). Depressed 5-hydroxyindole levels associated

4. Psychopharmacologic Agents

103

with hyperactive and aggressive behavior. Arch. Gen. Psychiatry 33, 331-336.
78. Schou, M., Baastrup, P.e. and Grof, P. (1970). Pharmacological and clinical problems of
lithium prophylaxis. Br. J. Psychiatry 116, 616-619.
79. Jefferson, J.W., Greist, J.H. and Baudhuin, M. (19tll). Lithium: Interactions with other
drugs. J. Clin. Psychopharmacol. 1, 124: 134.
80. Jefferson, J.W. and Greist, J.H. (1978). Lithium intoxication. Psychiatr. Ann. 8, 47-56.
81. Adams, F. (1984). Intravenous haJoperidol-lorazepam therapy for delirium. Int. Drug Ther.
N ewslett. 19, 33- 35.
82. Steele, E., Lucas, M.J. and Tune, L. (1986). Haloperidol versus thioridazine in the treatment
of behavioral symptoms in senile dementia of the Alzheimer's type: Preliminary findings. J.
Clin. Psychiatry 46, 410-312.
83. Feeney, D.M., Gonzales, A. and Law, W.A. (1982). Amphetamine, haloperidol, and
experience interact to affect rate of recovery after motor cortex injury. Science 217, 855-857.
84. Rao, N., Jellinek, H.M. and Woolston, D.e. (1985). Agitation in closed head injury:
Haloperidol effects on rehabilitation outcome. Arch. Phys. Med. Rehabil. 66, 30-34.
85. Richclson, E. (1985). Pharmacology of neuroleptics in use in the United States. J. Clin.
Psychiatry 46, 8-14.
86. Nasrallah, H.A., Dunner, F.J., McCalley-Whitters, M. and Smith, R.E. (1986). Pharmacologic probes of neurotransmitter systems in tardive dyskinesia: implications for clinical
management. J. Clin. Psychiatry 47, 56-59.
87. Remick, R.A. and Fine, S.H. (1979). Antipsychotic drugs and seizures. J. Clin. Psychiatry
40, 78-80.
88. Levenson, J.L. (1985). Neuroleptic malignant syndrome. Am. J. Psychiatry 142, 1137-1145.
89. Fewtrell, W.D., House, A.O.,Jamie, P.F., Oates, M.R. and Cooper,J.E. (1982). Effects of
vasopressin on memory and new learning in a brain-injured population. Psycho!. Med. 12,
423-425.
90. Abuzzahab, F.e., Will, J.e. and Zimmerman, R.L. (1982). Effects of single dose ACTH
4-10 versus placebo on the memory of symptomatic geriatric volunteers. Brief Reports 2,
65-69.
91. Bugiani, O. and Gatti, R. (1979). L-dopa in children with progressive neurological disorders.
Ann. Neurol. 7, 93 (letter).
92. Zis, A.P., Bibiger, H.e. and Phillips, A.G. (1974). Reversal by L-dopa of impaired learning
due to destruction of the dopaminergic nigro-neostriatal projection. Science 185, 960-962.
93. van Woerkom, T.e.A.M., Minderhoud,J.M., Gottschal, T. and Nicolai, G. (1982). Neurotransmitters in the treatment of patients with severe head injuries. Eur. Neurol. 21, 227-234.
94. Growdon, J.H., Cohen, E.L. and Wurtman, R.J. (1977). Treatment of brain disease with
dietary precursors of neurotransmitters. Ann. Int. Med. 86, 337-339.
95. Freund, H., Yoshimura, N., Lunetta, L. and Fischer, J.E. (1978). The role of the branchedchain amino acids in decreasing muscle catabolism in vivo. Surgery 83, 611-618.
96. Yellin, A.M., Hopwood, J.H. and Greenberg, L.M. (1982). Adults and adolescents with
attention deficit disorder: Clinical and behavioral responses to psychostimnlants. Brief
Reports 2, 133-136.
97. Garfinkel, B.D. (1986). Recent developments in attention deficit disorder. Psychiatr. Ann.
16, 11-15.
98. Klee, S.H., Garfinkel, B.D. and Beauchesne, H. 1986. Attention deficits in adults. Psychiatr.
Ann. 16, 52-56.
99. Woods, S.W .. Tesar, G.E., Murray, G.B. and Cassem, N.H. (1986). Psychostimulant
treatment of depressive disorders secondary to medical illness. J. Clin. Psychiatry 47, 12-15.
100. Ayd, F.J. (1985). Psychostimulant therapy for depressed medically ill patients. Psychiatr.
Ann. 15, 462-465.
101. Mattes, J.A. (1985). Methylphenidate in mild depression: A double-blind controlled trial. J.
Clin. Psychiatry 46, 255-257.
102. Stimulants for depression? (1986). MGH Bio!. Ther. Psychiatry 9, 17-18.
103. Lipper, S. and Tuchman, M.M. (1976). Treatment of chronic post-traumatic organic brain
syndrome with dextroamphetamine: First reported case. J. Nerv. Ment. Dis. 162,366-371.
104. Weinstein, G.S. and Wells, C.E. (1981). Case studies in neuropsychiatry: Post-traumatic
psychiatric dysfunction-diagnosis and treatment. J. Clin. Psychiatry 43, 120-122.
105. Brown, G., Chadwick, 0., Shaffer, D., Rutter, M. and Traub, M. (1981). A prospective

104

I. Theory and Intervention

study of children with head injuries: III. Psychiatric sequelae. Psychol. Med. 2, 63-78.
106. Angrist, B.M. (1978). Toxic manifestations of amphetamine. Psychiatr. Ann. 8, 13-18.
107. Peterson, L.G. and O'Shanick, G.]., eds. (1986). Psychiatric Aspects of Trauma. Vol. 16.
Advances ill Psychosomatic Medicjlle, Karger, Basel.
108. O'Shanick, G.]. (1986, Summer). Approaching the behaviorally disordered head injury
patient: A holistic assessment. Trends Rehabil., 10-12.

5. TRAUMATIC BRAIN INJURY AND THE REHABILITATION


PROCESS: A PSYCHIATRIC PERSPECTIVE

IRWIN W. POLLACK

1 INTRODUCTION

No illness or physical insult has a more disruptive or painful impact on the


lives of both victims and their families than does a traumatic brain injury
(TBI). Bewildered and frightened, they turn to the professionals for answers.
How will the injury affect the survivor? Is recovery possible? Will it be complete? Are there therapeutic measures that will help? How do they bring about
improvement?
A review of the results_ of neurological examinations, computerized tomographic (CT) scans, intracranial pressure readings, evoked potentials, and neuropsychological test findings contributes little to the family's understanding-and
even less to its support. A more relevant, easily understood, and reasonably
valid description of the extent and after-effects of the injury, as well as of the
prospects for successful rehabilitation, must be made available to the injured
persons and their families.
It is not easy to explain the complex phenomena associated with TBI in
simple yet valid terms, but-in every case-this is exactly what is required to
protect the developing therapeutic alliance between the therapist and the injured persons and their families. Some of the concepts that have been borrowed
or devised by members of our staff to explain the phenomena of traumatic
brain injury to our clients and their families will be described in the following
pages.
2 THE SURVIVOR: MISSED OPPORTUNITIES AND UNMET CHALLENGES

An opportunity is a situation or condition favorable for attainment of a goal


[1]. Some opportunities-such as noticing a road sign, meeting another person,
105

106

I. Theory and Intervention

or having paper and pencil available-are experienced universally, whereas


others-such as being admitted to a unviersity, hearing about an available job,
or learning that a desirabk,apartment has become available-are less common.
Each of these opportunities, when used to advantage, can positively affect the
course of a person's life.
Unfortunately, individuals who have suffered brain injuries frequently are
deprived of the usual opportunities that are appropriate for their age, educational background, occupation, and socioeconomic condition. This poverty of
opportunities can be accounted for by several factors.
First, most individuals who have suffered a traumatic brain injury go through
a prolonged "time out" from the ongoing world. There is a sudden break in
the progression of social experiences; the injured person spends months or
even years as a patient, moving from hospital to rehabilitation facility to outpatient care, all the while becoming more and more out of step with the rest
of the community. The young age (between 17 and 35) of most head-injury
survivors contributes further to the distancing, since it is during that age span
that profound changes in life-style most commonly occur. Because braininjured persons are unable to take advantage of the significant opportunities
that become available at this time, they fall behind their peers and, as a result,
they are often friendless and isolated.
Second, many traumatically brain-injured persons suffer some significant
physical impairments that limit their mobility and hence their access to the
opportunities that are available in the community outside of the home. Although
limitations in a person's ability to move about easily may have important consequences at any time oflife, the effect of those limitations is most significant
during early adulthood, when the satisfaction of occupational and social demands often requires considerable mobility. The extent of disability caused
by physical deficits is compounded by coexisting cognitive impairments.
Heightened distractibility, slowed response time, and impaired judgment-all
of which frequently result from a TBI-make it unsafe or even impossible for
the injured person to use the motorized devices that ordinarily provide physically handicapped people with the means of getting around in their community.
Finally, in most cases of severe brain injury, it is the subsequent problems
with cognition, rather than with motor function, that are most devastating
to the injured person. Inefficient information-processing, reduced ability to
recognize cause-and-effect relationships, impaired short-term memory, and
other deficits in cognitive function all contribute to the individual's difficulty
in recognizing and taking advantage of available opportunities.
2.1 Case Illustration: Joe

Joe was 19 when we first met him, three and a half years after he had been hit
by a car and had suffered a serious brain injury. He had lost all of his friends,
and he took no initiative in developing new relationships. In fact, unless he
was actively engaged by a member of his family, he appeared to be content to

5. Traumatic Brain Injury

107

sit alone, listening to his stereo. Even when he was stimulated enough to begin
an activity, he required constant cueing to follow through. His slow, dysarthric
speech was difficult for others to understand, and the possibility of his developing new friendships was compromised even further by his inability to find his
way around his neighborhood without a guide.
As a result of his brain injury, this 19-year-old could no longer meet many
of the demands that society makes on the average 10-year-old. The opportunities for further development that are open to most teenagers were completely
out of Joe's reach.
During the three and a half years that had passed since Joe's accident, his
peers had left him far behind, and there appeared to be little hope that he could
catch up. The gulf separating Joe from the rest of the world started with the
injury and widened with each passing day.
First, Joe was in a coma, completely unresponsive for one and a half months.
The total period of hospitalization was 18 months-a year and a half "out of
circulation," beginning when he was 15, an age when most adolescents are
experiencing new challenges and are undergoing significant changes.
The gulf was made even wider becauseJoe retained only a vague, nonspecific
recollection of his past life. For example, he knew that before his injury he had
had a number of friends, but he could not recall their names or any specific
experiences that they had shared. He remembered that his family frequently
spent summers at the seashore, but he could not recall the name of the town or
any descriptive details about the beaches, nor could he point to any specific
activities or events in which he had participated. Not only had his head injury
kept him out of the mainstream of life for 18 months, but it had essentially
wiped out the ties to other people that he had developed during the first 15
years of his life.
The rebuilding of broken relationships is always difficult, but for Joe it was
impossible, because he was no longer able to learn from new experiences. He
seemed unable to comprehend and internalize information, whether about
people, activities, or events, so that he retained little or no memory of an
experience that might have occurred only minutes before. The stultifying
effect that this state of affairs had upon Joe's attempts to re-engage is wellillustrated by the following vignette.
Early inJoe's cognitive rehabilitation, it was the therapist's practice to interrupt therapy after about an hour so that Joe could have a snack. The offices
were located on the third floor; to reach the cafeteria, which was on the first
floor, he had to walk down a corridor to the elevator, ride down three levels,
turn left as he came out of the elevator, turn left again at a dead end approximately 25 feet away, and finally turn right when he came to the lobby area.
Signs pointing the way to the cafeteria were located on the wall directly opposite the first-floor elevator doors, as well as at each intersection of corridors.
Joe had walked past the signs three times each week for months and still was
unable to find his way to the cafeteria without verbal directions.

108

I. Theory and Intervention

Even after the presence of a sign was brought to his attention, Joe seemed to
be unable to comprehend its significance. Only when he was asked specific
questions relating to the meaning and relevance of the words and symbols did
he appear to understand that he could use the sign as a guide to the cafeteria.
However, when he arrived at the next intersection, he once again failed to take
note of the information provided by the sign on the wall unless he was cued to
do so. In effect, Joe was able to interpret and use the information provided by
the sign ifhe was first told that something important could be gained by reading and thinking about the words and symbols printed on it. His cognitive
deficits caused him to be unaware that the signs provided him with the opportunity to find his own way to the cafeteria and to thus increase his ability to
function independently. Of all the impairments that followed his serious head
injury, perhaps the most devastating to Joe was the loss of initiative, the ability
to self-direct that is so important if one is to take advantage of opportunity.
2.2 Case Illustration: Bob

Bob, 23, had been injured several years before his admission to the Center
for Cognitive Rehabilitation. His small car had been struck from behind by
another vehicle that was traveling at a high speed. Bob was in a coma for a
period of six weeks, and although he made important gains during the next 12
months, he was left with significant physical and cognitive impairments. Bob's
insistence that he had no deficits and therefore had no need to follow advice or
to accept guidance was of particular concern to his parents and therapists.
On one occasion, Bob's parents requested a meeting to discuss his "inappropriate" behaviors. They complained that they could not longer entertain
friends or family because Bob said or did things that no one could understand
and that at times seemed almost bizarre. They wondered aloud if it was possible that Bob was crazy. Throughout the ensuing discussion, Bob sat quietly,
contributing nothing. He appeared to be uninvolved, or at least uninterested.
Suddenly he interrupted the conversation: "You've been talking about
kinesthesiology, haven't you, Doctor?" In fact, at that moment, we were discussing another topic quite far removed from anything physical in nature.
With some effort, I recalled that about five minutes earlier, I had mentioned
Bob's participation in our movement-therapy group and had described some
of the therapeutic exercises to his parents. Guessing that this was what Bob
was referring to, I acknowledged that indeed, sometime earlier, we had been
discussing movement activities, but that now the conversation had turned to
other issues. However, Bob was not so easily put off. Having found a topic of
mutual interest on which he could discourse, he launched into a monologue,
describing all that he could recall about movement and movement therapy.
When he finally ran out of material, he once again lapsed into silence.
Clearly exasperated, Bob's parents restrained themselves until the end of his
unsolicited contribution to the group discussion and then gave vent to their
annoyance, bewilderment, and concern. This, they said, was just the sort of

5. Traumatic Brain Injury

109

inappropriate behavior that they had been trying to describe to me earlier.


Understandably, they had failed to recognize that Bob had wanted to take part
in the conversation and be included in the group, but because he processed
information slowly, he could not keep up with the exchange. At one point, he
heard a familiar topic being discussed; over the next few minutes, he collected
his thoughts and then was ready to make his contribution. Unfortunately, by
that time, what he had to say was no longer relevant to the topic being discussed.
Bob had recognized that an opportunity to join the group had presented
itself, but because of his cognitive deficits he was unable to take advantage of
that fact. His belated attempts to make a contribution were seen as inappropriate, and the distance between himself and others was increased, rather than
reduced (as he had obviously hoped it would be).
2.3 Summary

Embedded in every opportunity are a number of challenges that must be


met if the opportunity is to be pursued to the individual's advantage. Without opportunities, there are no challenges, and without challenges, there
is no motivation. Opportunities can be "manufactured" by the family or
the therapist, but unless the injured person is required to meet the associated
challenges, no real growth will occur.
The challenges that must be met in order for people to make the most of
opportunities require action on their part. The required action must be more
than motion or movement behavior; it must be a true transaction, that is,
"action in which an individual is simultaneously organizing his or her internalized concepts of self into a unity while at the same time forming a systematic
relationship with things and people" (p. 3) [2].
The development of a sense of onself comes with the development of a sense of the
world. This happens through the person's actions that impact upon the world eliciting
some reaction (feedback); for example, an infant's spontaneous smile elicits a hug and
expressions of pleasure from the mother which in turn stimulate pleasurable sensations
in the baby, eliciting another smile (p. 7) [2].

Most people who have suffered a serious head injury have difficulty carrying
out transactions, even those of an elementary nature. The deficits that interfere
with "successful" transactions between brain-injured people and elements of
the surrounding world include 1) a lack of initiative (because of which, they
never get started); 2) a lack of awareness that some specific behavior is appropriate or required; 3) the failure to attend to guideposts, landmarks, or signals
that could guide their actions appropriately; and 4) the inability to differentiate
between relevant (important) and irrelevant (unimportant) information.
Because of such deficits, injured people perform in ways that further interfere
with successful transactions, such as 1) engaging in behaviors and then not
recalling that they have done so; 2) not really "getting the point" and therefore

110

I. Theory and Intervention

behaving inappropriately; 3) responding so slowly that the demand characteristics of a situation have already changed; and 4) behaving appropriately in
respect to one aspect of a situation but failing to modify those behaviors when
the demand characteristics change.
It is not easy to comprehend the degree of frustration that must be felt by a
brain-injured person who is able to recognize that opportunities exist but
who consistently fails to meet the challenges necessary to take advantage of
them. This is much more disheartening to the individual than is the failure to
recognize that any opportunities exist.
3 DEVELOPMENTAL EPOCHS: A PROBLEM OF FAILED TRANSITIONS

Family members, therapists, and others who are in frequent close contact with
brain-injured persons tend to focus on the difficulties that such individuals
have in carrying out ordinary transactions effectively. However, friends and
acquaintances who interact with the brain-injury survivors less frequently are
particularly struck by the fact that they appear to have stopped developing.
The injured people do not seem to be making transitions from one stage of
development to another, so that in many areas they appear to be very
immature when compared to other people of the same age and background.
Transitions occur as a result of a large number of transactions by means of
which a person successfully copes with a series of challenging physical, intellectual, and social demands. Successful transitions, therefore, depend not only
on the character of the environmental demands but also on the person's "readiness" for change. Periods of transition are especially stressful, because they
require some adjustment on the part of all parties involved, including the
developing person, family members, friends, colleagues, and therapists.
As a result of one unsuccessful transaction after another, impaired braininjured people fail to make friends, graduate from school, hold a job, marry,
or even live independently. In fact, their usual course is one of regressionnot of their physical condition but rather of their social situation: Established
friendships slip away, marriages dissolve, adult child-parent relationships
revert to dependent child-parent transactions. All too frequently, for the survivors, the most painful long-term effect of a serious head injury is loneliness.
3.1 Case Illustration: Dan

Dan was admitted to the Center for Cognitive Rehabilitation five years after
he suffered brain damage as a result of a viral infection. At the time of his
injury (at age 17), Dan was a bright, talented high school senior who had many
friends and an unlimited future.
The infection had relatively spared the left side of Dan's brain but had caused
significant damage to the right posterior brain areas. As a result of his injury,
Dan could no longer develop a mental image of the space that surrounded him.
Because he could not find his way around his neighborhood without a guide,
he was totally dependent on his family. His verbal skills were basically intact,
but his ability to comprehend any situation that had spatial aspects was severely

5. Traumatic Brain Injury

111

limited. For a while after his injury, he denied his difficulties, but gradually he
recognized that, in many circumstances, he could not trust his own senses.
Unable to depend on the accuracy of his interpretations of the world, he
became depressed and retreated to his room. There he spent his time alone,
reading, watching television, listening to his stereo, and, on occasion, writing
poetry.
3.2 Isolation: A Progressive Loss of Community

In similar societies (i.e., those having like customs and values), the same
sequence of transitions is experienced by most individuals at about the same
point in their lives. To a degree, a similar sequence of transitions was experienced by past generations as well. The shared aspects of the experiences that
lead to transitions foster a sense of family and community, a feeling of closeness. In a true sense, transitions arise in the present and are a link between a
person's past and the future.
A serious TBI brings about a sudden rupture in the continuity of shared
experiences, which is followed by a period of time, varying in length, during
which the survivor fails to make the "expected" and necessary transitions. As
a result, the distance between an injured person and his or her community
widens, and the comforting sense of closeness and familiarity is replaced by a
feeling of "not belonging" and isolation. It may be said that a brain injury
precipitates a true social "dis-ease"-an impairment of the individual's ability
to interact effectively with an everchanging world.
The residual deficits in cognitive and social function that result from serious
brain injuries can best be understood as by-products of unconsummated or
ineffectual transactions between injured people and their immediate physical
and social environments rather than as impairments that originate wholly
within themselves. Immediately following the injury, there is a sudden change
in the victim's state of consciousness, from being aware and in control to
having no awareness and no self-control (coma). During the period of coma,
no transactions between the injured person and his or her environment are
evident. In a sense, time stands still for a person in coma, while the rest of the
world moves on. Typically, the recovery process is prolonged; during that
period, most of the injured person's transactions occur in a very constricted
physical and social environment, often limited to family, close friends, and
hospital personnel.
Invariably, people who have suffered serious head injuries have no memory
about varying periods of time preceding and following the trauma (retrograde
and anterograde amnesia). Like prisoners of war who have been confined alone
in a cell for months, brain-injured people have been out of contact with the rest
of the world. However, unlike the released prisoners, who can reestablish their
connections with others, the brain-injury survivors have physical and cognitive deficits that interfere with the acquisition and retention of new information, so that reestablishing their connections to the community is always very
difficult-and may be impossible.

112

I. Theory and Intervention

4 THE CONTEXT: UNRECOGNIZED OR ILL-DEFINED

The combination of impairments suffered by most brain-injury survivors


interferes with their ability to recognize the context in which they are functioning [3, 4]. This ability is necessary to solve problems, to anticipate, and to
plan.
The term context is defined as "the set of interrelated circumstances or facts
that surround a particular event or situation which influences its meaning or
effect." [1] The ability to define a context appropriately requires the processing
of large amounts of information from various sources, along with the subsequent development of an abstract representation of the "whole." This ability
rarely remains unaffected following a serious brain injury.
Even everyday activities as simple as preparing for bed or mailing a letter
require the development of a sequence of events directed toward achieving that
end. Sequences cannot unfold in an appropriate manner unless people have the
ability to comprehend the interrelationships between the not-yet-achieved
goal, their past experiences in similar circumstances, and the intra personal and
extra personal factors that facilitate or impede each event. When it is recognized
that (except in the earliest stages of learning) an unimpaired person comprehends the entire sequence of events as a configuration rather than as a sum of
unrelated parts, the extraordinary complexity of even the most basic goaldirected activities becomes evident.
When some process (such as that produced by a traumatic brain injury)
dissociates one event from another, a person's goal-directed activity is disrupted. Under these circumstances, starting an activity-that is, initiating the
first of a sequence of events (behaviors)-is most difficult because the person
has little ability to predict what will follow. Furthermore, without a sense of
the interrelationship between events (the context), the person will find it all
but impossible to differentiate between new information that is relevant to the
attainment of the goal and information that is not relevant.
4.1 Case Illustration: Joe

Joe's failure to notice-let alone comprehend-the significance of the words


and symbols on the signs that he passed, day after day, can in large part be
explained by his inability to form a context. He appeared to be unable to conceive of how the various components of his journey to the cafeteria were interrelated: the purpose of the journey (to get a snack); the starting point (the
therapist's office); the end point (the cafeteria); the guiding landmarks (the
elevators, the signs that pointed the way to the cafeteria that were opposite
the elevator doors and at each intersection of the corridors), and finally the sign
on the cafeteria door.
Joe did not recognize that an elevator or a sign could be a guiding landmark
until the therapist pointed it out to him, and even then he did not comprehend
the relevance of that information to the goal of "getting a snack." However,
when the therapist called Joe's attention to the sign and indicated that he could

5. Traumatic Brain Injury

113

use it to help him find his way to the cafeteria where he would be able to get
his snack, Joe was able to read the sign and follow its directions. Unfortunately, when he arrived at the next intersection, he once again failed to respond
to the sign at that location. He seemed to have no mental image or "model" of
the journey to the cafeteria "to get a snack" against which he could evaluate
the relevance of each new bit of information. Because of this, no new learning
occurred.
5 DISCONNECTION FROM THE SENSE OF SELF:
A PROTRACTED IDENTITY CRISIS

For people to develop and maintain a sense of self, they need consistent feedback from the environment in response to their actions. Following a serious
brain injury, injured people are disconnected from their preinjury identity
in many ways, all of which contribute to a loss of the "sense of self." These
include the extended period of hospitalization (when they are out of contact
with the ongoing world), the loss of some or all memories of the past, the
impaired ability for learning new material, the loss of the ability to anticipate
future events, and the loss of a sense of community with others. This disturbing state of affairs is made even worse by the fact that any possible new
identity is unacceptable to the remnants of the former self because it (the
new self) is seen as flawed, impaired, incompetent, or even crazy. It is in these
circumstances that denial becomes evident.
There is no doubt that some denial may be the direct result of impaired
cognitive function caused by the destruction of tissue in selected areas of the
brain [5]. However, students of human behavior have recognized for many
years that denial can occur in individuals who have not suffered a brain injury.
They have observed further that, in many cases, denial serves as a "last ditch"
defense against the recognition of some unacceptable aspect of the self that, if
confronted, would result in significant damage to the integrity of that person
[6].
There is every reason to believe that denial that is the result of disturbed
brain anatomy can serve a similar protective function for the injured person
[7]. When this is the case, a direct assault on the brain-injured person's denial is
disrespectful and potentially damaging both to the rehabilitation process and
to the injured person.
The assumption that denial is always a stumbling block that interferes with
progress during rehabilitation, and therefore that it must be eliminated, is
wholly unfounded. It is likely that, in the majority of brain injuries, denial acts
as a significant defense mechanism as well as a hindrance to the rehabilitation
process. Of course, when denial interferes with rehabilitation, it must be eliminated, or at least its effects must be minimized. This can be best accomplished not by a direct assault, but rather indirectly-through the strengthening
of intact areas of cognitive and physical function. When injured people recognize that they have regained a level of competence in several areas of func-

114

I. Theory and Intervention

tion, they become less threatened by the recognition of significant deficits in


other areas. This can only be accomplished through completing a series of
successful transactions, such as the sharing of rewarding experiences with
others. In this way, the brain-injured person is gradually able to reestablish
an aceptable identity. The rehabilitation program must, therefore, be designed
to foster the active interplay between clinician and client, as well as between
client and client, and client and family.
6 THE REHABILITATION PROCESS

The goal of the rehabilitation process is the restoration of the injured person's
ability to function in society. More narrowly, the goal for the person is to
establish and maintain his or her individuality while participating in group
activities. Ultimately, the goal is to reestablish a sense of self-an acceptable
identity.
6.1 The Concept

The rehabilitation effort must make a series of opportunities available to braininjured people while providing them with the assistance that they require to
make the most of each of them. Early in the course of rehabilitation, therapists
present carefully selected opportunities in which are embedded a limited number of known challenges. Later in the process, therapists help the brain-injured
clients to recognize naturally occurring opportunities, while continuing to
provide them with assistance so that the associated challenges can be successfully met. The point in the course of rehabilitation at which the first phase ends
and the second begins is individually determined, and no clear boundary can,
or should, be drawn between them.
Successful rehabilitation requires the development of therapeutic situations
in which true transactions between therapist and client (or clients) take place.
These are situations in which both parties must adjust if the challenges are to
be successf~lly met. Ordinarily, a didactic teaching approach does not meet
this requirement. However, an approach in which the patient is required to
respond to cues provided by the therapist does meet the criteria of mutual
adjustment, because the therapist must modify his or her cues to fit the injured
person's changing behavior. Within limits, the therapist can dictate the direction of changes in the behavior of a client but not the rate of these changes or
the strategy by which the changes are brought about.
Ideally, the therapist should not propose a particular strategy for the client to
use. To do so would imply either that, in a given situation, the same strategy
that would be effective for the nonimpaired therapist would be effective for the
brain-injured client or that the therapist is so certain of both the impairing
effects of the injury and of the client's preserved abilities that he or she can
select the "best" approach for the client to pursue. Because neither of these is
usually the case, it makes more sense (and is more respectful) for the therapist
to encourage the client to tackle the problem at hand with whatever abilities

5. Traumatic Brain Injury

115

are available to him or her. Then, by presenting appropriate clues, the therapist
can help the client to modify his or her approach to the problem so that its
effectiveness is increased.
6.2 A Logical Framework for the Rehabilitation Process

The integrity of the cognitive process depends upon the effective integration
of information arising from many sources, including vision, hearing, touch,
taste, smell, movement, and position sense. In addition, it requires an intact
memory for past experiences as well as an awareness of the present. The cognitive process is a complex phenomenon that cannot be broken down into its
component parts, because one mental event does not relate to another in any
simple way [8].
Despite this, rehabilitation specialists regularly divide the process into simpler, more manageable units such as attention, memory, orientation, discrimination, sequencing, and categorizing. This is done so that the therapists
can easily recognize the specific information that they have imparted to the
clients. This, in turn, permits the clients' responses to be evaluated systematically. Client output can be assessed and can even be assigned a numerical
"score." Although such numbers permit the therapist to document changes in
particular abilities, they do not truly reflect the process itself. Practicing isolated mental activities can lead to increased competence in much the same way
that exercising an isolated muscle can increase its strength. However, it must
be recognized that the effort to strengthen a person's particular mental activity,
even when successful, will not automatically improve that person's ability to
reintegrate that activity into the ongoing process of cognition.
This breakdown in the injured person's ability to integrate mental "events"
effectively undermines an analytic and directive approach to cognitive rehabilitation, because this approach breaks down the process of cognition into component parts, thereby placing the burden of reintegrating these components on
the brain-injured person. An effective rehabilitation program should emphasize
the reintegration of mental abilities, rather than the "rebuilding" of isolated
impaired elements.
To help clients relearn old strategies or develop new, more effective strategies for processing information, the rehabilitation staff must operate from
an integrative model, rather than from a traditional team model. In such an
integrative model, each therapist is prepared to assist clients in their attempts
to overcome a wide range of problems-physical, intellectual, and social.
Each therapist contributes-and, to an extent, shares-his or her special expertise with other staff members in their common therapeutic efforts. In this
model, each therapist functions in four separate but related roles: 1) cognitive
therapist; 2) practitioner of a particular rehabilitation specialty (e.g., occupational therapy, physical therapy, speech pathology, vocational counseling,
therapeutic recreation, neuropsychology, social work); 3) consultant to the
rest of the therapeutic staff in matters that pertain to his or her area of special-

116

I. Theory and Intervention

ization; and 4) case manager. The role of the case manager in this model is that
of a coordinator of clinical activities as they relate to specific clients; it includes
the development of working relationships with the clients' families as well as
with other health professionals, school authorities, employers, and the like.
The role of a therapist shifts to keep step with the changing needs of the
clients. For example, the speech pathologist may be the appropriate case
manager for a client who has major deficits in language functions. Later,
after the client has made significant gains in language ability, the vocational
counselor may take over as case manager, because the major emphasis has
changed from improving language functions to preparing the client for a return
to the work force.
During the earliest phases of the rehabilitation process, therapists often work
to reestablish or to strengthen clients' isolated cognitive or physical abilities.
During later phases of the rehabilitation process, greater emphasis is placed
upon integrating several areas of function. Rehabilitation activities that are
structured to foster the integration of function are less appropriate for one-toone therapeutic work. Instead, a small group of therapists, representing different areas of expertise, works with a small group of brain-injured clients,
engaging them in activities that are meaningful and realistic samples of the
"real world." In this way, the full range of expertise possessed by the individual members of the therapeutic team is available to help each client at each
step in the rehabilitation process. (One-to-one rehabilitation activities continue
to be used to bolster areas of deficit in basic cognitive and physical abilities that
limit clients' effectiveness in the group situation.) In these groups, the staff
functions collectively and the role ofleader shifts from therapist to therapist to
meet changing group needs.
Every rehabilitation program should present clients with a series of graded
challenges representative of those that they will meet in the community. Rehabilitation activities based on familiar tasks and situations enhance the clients'
motivation and ease the transfer from the therapeutic environment to the "real
world. "
Cognitive impairments that follow a brain injury often include deficits in
short-term memory and in information-processing. Such deficits make it difficult for individuals to call upon past experiences as a guide for successfully
meeting the challenges that arise in the course of rehabilitation. For this reason,
therapists initially must structure rehabilitation activities so that all the necessary elements to meet the challenges are immediately available to the clients.
Later, sequences of data or activities are added, so that clients can relearn how
to use information from the immediate past to assist them in solving current
problems. The course of rehabilitation gradually moves from a here-and-now
orientation to incorporate information derived from past experiences.
Brain-injured people cannot begin to anticipate future consequences appropriately until information from both past and present sources is once again
available to them. Their inability to anticipate the future consequences of their

5. Traumatic Brain Injury

117

present actions is a major deficit; in extreme cases, it can lead to immobility.


For brain-injured people, moving ahead without being able to visualize a possible future is tantamount to stepping into a void. No one can be certain about
the future, but a clear sense of the present and a good recall of one's past provides material for the development of a concept of tomorrow. An unacceptable
or inaccessible past leaves the individual with no guideposts from which to
develop a concept of "future." This, at least in part, accounts for the braininjured person's lack of initiative and the failure to make expected transitions.
7 GROUPS: THE BASIC VEHICLE FOR REHABILITATION

This "real world" to which the survivor hopes to return some day is composed of groups-large groups, small groups, quartets, triads, and pairs.
Individuals are linked together as classmates, business partners, team players,
families, partners in marriage or business, sellers, buyers, lovers, and friends.
In the "real world," no one functions wholly in isolation; there are always
others present, if only in a person's memories, thoughts, plans, fears, desires,
and hopes. In a therapeutic environment, isolation is an unacceptable condition, except for brief periods as part of a behavior-modification program. Each
brain-injured client must interact regularly with at least one other person, or
no change in his or her situation will occur.
7.1 How Groups Function

Each type of group has its own ground rules or scripts [9] that individual
members must follow in order to ensure that the concerted action will be effective. For example, transactions between individual members of a large group
(e. g., school classes, legislative assemblies, fraternity meetings, boards of
directors) are guided by formal rules of procedure, which are enforced by an
appointed or elected leader. Individuals must be formally recognized to be
permitted to speak, and the issue under discussion is defined by group interests
rather than by individual interests. Following discussion, subgroups may be
formed to develop additional information. When the group leader believes that
sufficient information has been made available, a decision relating to group
action is made by soliciting group approval or disapproval through a vote.
Although the large group itself is not task-oriented, it assigns tasks to smaller
working groups. Smaller groups, perhaps consisting offour to eight members,
function less formally, often without a preselected leader. Individual group
members, who frequently are well-acquainted, join together in order to work
on a specific task or toward some agreed-upon goal. The personality, needs,
and desires of individual group members influence the outcome of the smallgroup endeavor to a far greater extent than is the case in large groups. Triads
and pairs are concerned with the needs or interests of one or more of the individuals involved, and the personal characteristics of the members have a m~or
impact on the nature of the transactions.

118

I. Theory and Intervention

7.2 Therapeutic Groups

In a therapeutic group activity, the needs of the injured persons take precedence
over those of any other group members (e.g., therapists, family, friends).
A serious traumatic brain injury invariably results in the loss of a person's
formerly automatic social behaviors. For this reason, group experiences of
various types should constitute the "core" experiences in every brain-injured
person's rehabilitation program.
7.2.1 The

La~!?e

Group

The orientation of a large group is towards the short-term future rather than
to the present. The task of a large therapeutic group is to define the context
and to determine the agenda that, when followed, will delimit the "world" in
which the next phase of the rehabilitation process will take place. The actions
of the large group are basically verbal in character. The group is structured in
ways that enable each member to be informed; it provides each member with
an opportunity to express his or her opinion; and it provides a mechanism for
incorporating individual contributions into the group process.
As is true in the community, individual members of the large group function
with varying degrees of competence and interest and with different priorities.
Large-group transactions usually are not oriented towards serving the immediate needs of individual group members. Instead, thejob of the large group
is to generate ideas, determine goals, gather information, set agendas, and assign tasks to smaller subgroups. The large therapeutic group provides a forum
in which a brain-injured person can relearn how to communicate effectively in
a formal, structured, and controlled social situation.
7.2.2 The Cluster Group

In the smaller therapeutic ("cluster") groups, members follow the agenda and
carry out the tasks assigned to them by the large group. The focus of the small
group is on the present, and the assigned tasks usually are performance-oriented.
In this situation, the term peiformance includes the verbal mediation necessary
to carry out the task. In some cases, the tasks may be primarily verbal in character, for example, obtaining information about a proposed trip. These small
therapeutic groups function in a less formal manner, and their mode of operation varies according to the perceived needs and abilities of a "like" group or
cluster of clients.
In a sense, the cluster group serves as a mediator between the relative impersonality oflarge-group decisions and the needs and desires of the individual
brain-injured person. In this group, individualized short-term task-related
goals are defined, and feedback is provided about the effectiveness of each
client's approach to his or her part of the assigned group task. For the small
group to be an effective therapeutic vehicle, the efforts of each of the group
members must contribute significantly to achieving group goals, that is, each
person's contribution must be necessary for the group's success.

5. Traumatic Brain Injury

119

TableS-I.
More abstract, verbal

Large Group (future-oriel/ted)


Short-term goal: effective management ofleisure time
Agenda: planning and carrying out a picnic for clients, staff,
and friends-choosing theme, location, and date
Cluster Groups (presellf-oriel/ted)
Assignments: Group 1-invitations and information;
Group 2-food; Group 3-entertainment; Group 4transportation
Example: Cluster Group #2
(assigned area-food)
Tasks: (1) menu development; (2) shopping; (3) cooking;
(4) serving and cleaning up
Illdividual client tasks
Clients #1 and #2-develop a shopping list from the menu
and shop for food, paper goods, etc.

More concrete performance

Task-related individual therapeutic activities: sequencing,


communication, eye-hand coordination, deductive
reasoning, management of time and space, organization of
information

7.3 Individual Therapy

The need for specific individual therapy is determined from an interpretation


of the results of tests (neuropsychological and other types) as well as from
observations of the client's behavior in both large and small groups. Individual
therapies usually are present-oriented. Therapeutic exercises are designed to
strengthen the client's impaired abilities or, if that is not possible, to help him
or her to develop compensatory strategies. In practice, of course, "individual"
therapy always involves two people (a pair) and sometimes three (a triad).
7.4 Relationship of Group Activity to Individual Therapy: An Example

A large group started with the universal goal, "learning to live independently."
The group decided that, for one month, clients would work on exercises
directed towards the more effective management of leisure time. The agenda
for the subsequent week consisted of activities necessary to plan and carry out
a picnic. The large group also decided on the theme for the picnic, the site, and
the date. Each cluster group was assigned a set of tasks relating to some aspect
of the agenda, and a flow-chart was developed (Table 5-1).
In such a project, each individual contributes to the group endeavor while
engaging in therapeutic activities designed to meet his or her special needs.
The end-product is tangible and is a true sample of the "real world."
8 SOME THOUGHTS ABOUT SUCCESSFUL THERAPEUTIC INTERVENTIONS

In general, the more that two people share experiences, the more likely it is
that they will use similar problem-solving strategies. However, a brain injury
disrupts many of the commonalities that formerly existed between the injured

120

I. Theory and Intervention

person and other people. For this reason, strategies that are useful to the therapist may be less useful, or may even be incomprehensible, to the client. A
didactic approach to cognitive rehabilitation assumes that a strategy that will
"work" for the therapist will also "work" for the client. This assumption fails
to recognize the extent of the estrangement that is experienced by brain-injury
surVIvors.
Therapeutic efforts will be successful only after a sense of commonality
and sharing has been established between the injured person and his or her
therapist (or therapists). This can only be accomplished through the active
participation of both parties involved in the therapy.
Therapeutic interventions must be relevant not only to the client's impairments but also to the client's individuality and to the environment in which
the client exists (or will exist). Rehabilitation exercises must be functionally
oriented and should have a "real world" significance.
Placing too much emphasis on the outcome of a therapeutic activity-that
is, on the success of the client in reaching a goal or in producing a product,
rather than on the developing strategy or process-can lead to an adversarial
relationship between the injured person and his or her therapist.
Clients must be aware that they have something to gain and that problems
in certain areas of functioning have been contributing to their failed attempts
to reestablish themselves in the community. This does not mean that every
brain-injured person must confront his or her deficit, but rather that rehabilitation activities should be designed to demonstrate to the clients' satisfaction
that they still possess areas of competence and are therefore still worthwhile
people. With each success, the clients' sense of self-value increases and the need
for denial decreases.
Therapeutic activities should build systematically toward short-term and
long-term goals. Challenges should progress from concrete to more abstract,
from simple to complex, from unimodal to multimodal, and from the present
to incorporate the past, and then to anticipate the future.
Each therapeutic exercise should have readily discernible steps, stages, or
expected responses that will serve as landmarks to permit the therapist and
client to follow the progress that is being made. Initially, landmarks are provided by the therapist. Later, the client should select appropriate landmarks
from a group of possible landmarks. Still later, the client should develop his or
her own landmarks-that is, should anticipate and self-monitor.
The structure in which therapeutic activities are embedded must be flexible
enough to change in response to the client's changing competence and needs.
Certain information must be shared with the client to foster a feeling of
community and to provide a basic orientation to the therapeutic task. The
information should include: 1) the purpose of the activity (e.g., to improve the
ability to plan effectively); 2) the short-term goal (e.g., to plan and carry out
a project); 3) the starting point or the "givens" (e.g., materials to be used,
other persons involved, time scale); and 4) the steps or stages that should be
traversed.

5. Traumatic Brain Injury

121

The procedure or strategy for carrying out the task is developed by the client,
who is assisted by the therapist. This problem-solving activity is the essence of
the therapy.
8.1 Responsibility, Uncertainty, and Stress

Of necessity, the attitudes of the professionals who are responsible for carrying
out a program of cognitive rehabilitation must reflect the uncertainty that is
felt by their clients. That is, the therapists know the rehabilitation goals, but at
the outset, like the clients, they cannot be sure of the best path to follow to
achieve them. A.s much as the therapy staff might like it to be otherwise, this
lack of clarity and the accompanying tension are inevitable when working
with brain-injured persons. The staff tension could be eliminated by removing
choices and standardizing the sequences of rehabilitation activities, but this
"lock-step" approach fails to acknowledge the individuality of each client and
therefore falls short .
. The following example illustrates a typical therapeutic task and the problems that it presents to both clients and therapists: Each of us must know the
meaning and significance of comparative words such as lesser/greater, lighter/
darker, and higherllower. If we cannot comprehend these relationships, our
concept of the world is limited to absolutes and our ability to communicate
effectively with others is severely compromised. Unfortunately, this is just the
way in which many brain-injured people perceive their world. It is most appropriate, then, for the rehabilitation team to implement a series of therapeutic
exercises that are designed to improve the clients' ability to deal with these
difficult relationships.
The initial problem is how to find a way to explain the nature of these exercises and their significance to brain-injured individuals so that they will understand them. In practice, this is impossible to do in words without using a series
of high-level abstractions, which in most cases are incomprehensible to individuals who have suffered a serious brain injury. A more effective way to help
brain-injured people to develop a rationale for the therapeutic exercises is for
therapists to demonstrate the relationship "greater/lesser" (perhaps using peas
or marbles). However, the clients may still not understand an explanation of
why this is an important task to engage in. Even if the clients are given a more
active role in the exercise so that they, rather than the therapists, manipulate
the items, they may not comprehend an explanation of the value of the task. In
both of these situations, the therapists' input necessarily will be limited to an
appeal to the clients to have faith that there are good reasons for working on
the exercise. A more effective approach sidesteps the necessity for explanations in the abstract by presenting the brain-injured individuals with a task
that requires that the relationships "greater/lesser" be handled in order to
reach a concrete end-product. The answer to the question, "Why is, it important for me to develop a strategy for managing comparatives?" now is easily
comprehended by the clients: "It is necessary in order to get an acceptable
end-product. "

122

I. Theory and Intervention

In the course of developing an effective strategy for handling comparative


relationships, a therapist has no choice but to follow the brain-injured person's
lead, interrupting only in order to call his or her attention to an error. At this
point, the therapist makes additional information available and encourages the
client to reassess the situation before proceeding with the task. In this approach,
the therapist's role is reflective, rather than controlling, and tentative, rather
than decisive. For these reasons, participation in the process necessarily produces some level of tension in the therapist. Anxiety great enough to impair
the function of a therapist indicates some problem in the treatment relationships. However, the absence of tension strongly suggests that the therapist is
maintaining his or her comfort by overcontrolling the process of rehabilitation.
Stress in members of the therapy staff can be engendered by a number of
non-client-related factors as well, such as inadequate numbers of staff, insufficient equipment, limited physical space, slow-moving institutional bureaucracy, insufficient funds, and unreasonable time allocations. There is probably
no rehabilitation program in existence that does not suffer from one or more
of the above-listed shortcomings. Even under the best circumstances, therapy
staff must be prepared to work within the same constraints and with the same
limitations that every independent person contends with in his or her own
community. Therapists must expect no less from themselves than they do
from their clients: that is, to acknowledge their own assets and limitations and
to work with both, whether these are imposed upon them by a brain injury or
by the environment.
Granted, therapists must plan and guide their clients' rehabilitation according to the clients' needs, but invariably they must do so in a therapeutic environment that is shared by others who have different needs and in the face
of constraints that are imposed by a relatively uninterested "real world."
Therapists can and should seek counsel from others when the need arises, but
they may not give the responsibility for their clients' therapy over to someone
else, any more than can clients give responsibility for themselves over to other
people and at the same time progress toward independence.
In the effort to help brain-injured people toward independence, staff members do have a few significant advantages over those available to their clients.
In the first place, therapists understand the ways in which a TBI can impair
physical and cognitive function. Each therapist has been trained in methods
that can help the injured individuals improve their level of competence. Perhaps
even more important, staff members are able to-or at least should be able
to-maintain an attitude of objectivity in their work with brain-injured people,
who usually need assistance in recognizing their assets and liabilities and in
recognizing when they are (or are not) behaving in an effective way.
To maintain a therapeutic environment, it is equally important for staff
members to view their own transactions with clients and other professionals in
an objective manner. The same uncertainties that promote staff tension make
it more difficult for individual therapists to separate their needs from those of

5. Traumatic Brain Injury

123

their clients. However, it is essential for each staff member-no less than for
his or her clients-to maintain the ability to self-monitor. Consultation and
supervision, provided by senior staff members or even by peers, should be
available to help the individual staff member maintain his or her objectivity.
8.2 Other Therapeutic Activities

The activities of the large context-forming group and the small cluster groups
do not constitute the total program of rehabilitation. Other group activities
and individual therapeutic activities serve as additional vehicles for carrying
out both the group agenda and the individual treatment plans. Such activities
may take place in groups that focus on body awareness and movement, selfexpression and self-awareness (drama), biofeedback and relaxation, recreational and vocational activities, and counseling and personal growth.
Most brain-injured people continue to need some basic rehabilitation therapies for a protracted period-for example, physical therapy, speech and language therapy, and some aspects of occupational therapy. These therapies, which
are usually delivered one-to-one, arc more medical in character; for this reason,
they should be viewed as necessary but essentially unintegrated experiences
in a client's program of rehabilitation. For example, sessions of individual
physical therapy fit into the brain-injured client's flow of experience in much
the same way as a series of regular visits to a physician's office for treatment
of a non-life-threatening ailment fit into the life-flow of a non-head-injured
person. Both are of immediate importance, but neither is an integral part of the
person's life experience. If medically oriented therapies still must constitute a
major portion of a client's rehabilitation, the client may not yet be ready for a
transitional situation.
9 THE IMP ACT OF PSYCHIATRIC DISORDERS: SOME COMMENTS

There is little doubt that a TBI predisposes the iruured person to the development of psychiatric problems. In fact, it is now recognized that injury to
specific areas of the brain is associated with specific psychopathological symptoms. (For example, individuals who have sustained left frontal damage frequently exhibit signs of depression [10, 11].) As with any other behavior, the
character and severity of psychopathological behavior are influenced by the
injured person's long-standing personality style, as well as by the environmental factors that are acting upon the person. In any given case, there is
no simple formula that will permit us to predict the results of the interplay
between the brain injury, the preinjury personality, and the environment. The
complexity of these interactions makes it difficult to describe them in simple,
down-to-earth terms. Despite this, the therapist must try to find a way to provide a valid, clear, and concise explanation for the injured person's disturbing
behaviors.
In the preceding pages, I have described the impact of a TBI on the quality
of the injured person's life, as well as its impact on the lives of his or her family

124

I. Theory and Intervention

and friends. Such factors as missed opportunities, un met challenges, unsuccessful transactions, failed transitions, loss of a sense of community, ill-defined
contexts, and disconnection from a sense of self are best conceived of as the
result of interplay between the brain injury, the person's preinjury personality,
and the physical and social environment. These factors do not unfold sequentially but rather are interactive, each contributing to the brain-injured person's
disability.
The character of postinjury psychiatric symptoms is determined by the set
of factors that most influence the individual's view of himself or herself relative
to other significant people in his or her life. For example, some people who
suffer a head injury after they have been successful in their occupations complain about the opportunities they are missing because they arc no longer able
to function well enough to meet challenges that they formerly could have
managed with ease. They despair because they no longer receive recognition,
promotions, invitations, and the like. For these individuals, the theme is loss,
and the psychiatric symptomatology is that of depression.
For other brain-injured people, the most obvious problems seem to revolve
around the inability to establish an appropriate context and the loss of a sense
of community. Because of this set of circumstances, the injured individuals are
unable to fathom the motives of others, and they tend to be uncertain about
the meaning of communications, both verbal and nonverbal. The ensuing
psychiatric problems arc paranoid in character. If, in addition to the loss of
commonality and the difficulty of establishing a context, injured persons
experience a significant loss in their sense of self, their few remaining links to
reality can break, resulting in a paranoid psychosis.
Commonly, following traumatic brain injury, a person's obsessive personality traits, which formerly had been adaptive, become so intensified and
so rigid that the new learning that is necessary for successful rehabilitation
becomes impossible. This occurs when the brain-injured person becomes aware
of his or her unsuccessful transactions and his or her failures to make the transitions from one life-stage to another that are required in order to maintain a
connection with "significant others." In the absence of the ability to find
alternative behaviors-and in a last-ditch effort to maintain relationships with
other people-the injured person intensifies behaviors that have already been
found to be wanting.
Similarly, every other significant psychiatric disorder suffered by traumatically brain-injured people can be described in terms of the several factors
that result from the interplay of the injury, the preinjury personality, and the
environmental circumstances.
10 CONCLUSION

The foregoing discussion does not mean to imply that there is no reason for
professional counselors to have an understanding of the classic theories of psychopathology, any more than it holds that knowledge of the neuroanatomy

5. Traumatic Brain Injury

125

and the pathophysiology of brain injury is unnecessary for the neurosurgeon


or the neuropsychologist. It docs, however, support the view that this type of
information means little to brain-injured people and to their families.
Likewise, the average therapist who works in a program for the brain-injured
will find little in the accumulated data that will help him or her to comprehend
the impact that the injury has made on the survivor because, as we have seen,
this impact is the product of the interplay of forces. Only by dealing with it in
the form in which it is experienced by our brain-injured clients can we develop
a relevant approach to brain-injury rehabilitation.
REFERENCES
1. The Ralldom House ColleRe Dictionary, rev. ed. (1975). Random House, New York.
2. Lichtenberg, P. and Norton, D. (1970). Cognitive aIld Melltal Development in the First Five Years
of Life. National Institute of Mental Health, Rockville, MD.
3. Bro~nell, H., Potter, H., Bihrle, A. and Gardner, H. (1986). Inference deficits in right brain
damaged patients. Brain Lang. 27, 310-321.
4. Gardner, H., Brownell, H., Wapner, W. and Michelow, D. (1983). Missing the point: The
role of the right hemisphere in the processing of complex linguistic materials. In Cogllitille
Processes ill the RZ~lzt Hemisphere, Perelman, E., ed., Academic Press, New York, pp. 169-191.
5. Weinstein, E. and Kahn, R. (1955). Denial of Illlless, Charles C. Thomas, Springfield, lL,
pp.3-9.
6. Engel, G. (1962). Psychological Developmellt in Health alld Disease. W.B. Saunders, Philadelphia,
p.365.
7. Sandifer, P. (1946). Anosognosia and disorders of body scheme. Brain 69, 122-137.
8. Luria, A.R. (1973). The Working Bra ill , Basic Books, New York, pp. 30-33.
9. Schank, R.C. and Abelson, R. (1977). Scripts, Plalls, Goals and Understalldillg. Lawrence
Erlbaum, Hillsdale, NJ.
10. Robinson, R. (1986). Post-stroke mood disorders. Hosp. Pract. (April 15), 83-89.
11. Robinson, R., Lipsey, J., Rao, K. and Price, T. (1986). Two-year longitudinal study of poststroke mood disorders: Comparison of acute-onset with delayed-onset depression. Am. J.
Psychiatry 143, 1238-1244.

6. THE NEUROPSYCHOLOGICAL INVESTIGATION AS


A THERAPEUTIC AND REHABILITATIVE TECHNIQUE

ANNE-LISE CHRISTENS.EN

1 INTRODUCTION

This chapter will outline the specific neuropsychological investigation that is


structured on the basis of A.R. Luria's work, mainly described in his books,
Traumatic Aphasia [1] and Higher Cortical Functions in Man [2]. It is suggested
that a qualitative examination along these lines must be accomplished to plan
rehabilitation and neuropsychological treatment strategies for brain-injured
patients. The models that Luria initially developed-as well as subsequent
additions and modifications-will be explored, not only as a method for
neuropsychological examination, but also as rehabilitation and neuropsychological treatment tools.
It is the intent of this chapter to emphasize the phenomenological aspects
that show how this investigation is in accordance with the insight obtained by
the growing neurobiological knowledge about the functioning of the brain.
2 NEUROPSYCHOLOGY: A BRIEF REVIEW OF LURIA'S THEORY

Luria [2] has described how a "troika" of psychologists (Vygotsky, Leontiev,


and Luria) in the U.S.S.R. in the 1920s and the early 1930s worked with ideas
in physiology and psychology that later were manifested in the new science of
neuropsychology. It was not until the late 1950s and 1960s, however, that the
group's publications, earlier available only in Russian, were more generally
available to Western psychologists. Luria's Restoration of Brain Function after
War Injuries [3] appeared in Russian in 1948 and in English in 1963. Traumatic
Aphasia [1] was published in Russian in 1947 and appeared in English in 1970.
127

128

I. Theory and Intervention

The comprehensive and more extensive book Higher Cortical Functions in Man
[2] is a translation (in 1966 and 1982) of the Russian book from 1962. These
books combined theory \vith the method of the neuropsychological investigation. Higher Cortical Functions in Man [2] and later The Working Brain [4] made a
great impression on the Western neuropsychological world. It was evident
that new insight into the organization and function of the brain was presented
in these works. Teuber [5] wrote in a preface to The Working Brain: "Here then
is a book, written by a master in his chosen field. Its translation marks a further
step in the mutual recognition of common values in scientific endeavors in
East and West" (p. xiv). Pribram [6], in a companion preface, stressed the
book's "scientific endeavor, while at the same time guarding the spirit and
substance of Soviet experience in this area of science" (p. xv).
2.1 Theoretical Tenets

In Luria's approach to neuropsychology, some main trends are discernible.


The neurophysiological tradition in Russia-from Sechenov to Pavlov, and
Bechterev to Anokhin-had given evidence of the organization of cerebral
reflexes and of the principles of "reafferentation" and "feedback." (Afferent
refers to information that comes into the brain from the central and peripheral
nervous system.) In this light, Luria and his collaborators had reconsidered the
concept of "psychological function" [2]. Their conclusion was the assumption
that complex behavioral processes are not "localized" but are distributed
throughout the brain in "functional systems." These systems are organized so
that each cortical zone contributes to behavior in a specific way-in accordance with its hierarchical position and governing principles. Therefore, the
coordinated working of all cortical areas responsible for the elements of a
complex behavioral act is necessary for the act to be performed in a precise
and smooth way.
The only effective way of obtaining knowledge of the cortical areas and
their interrelated and coordinated functions is through a qualitative analysis of
psychological processes. These processes belong to the basic knowledge not
only in general psychology, but also in experimental and developmental
psychology.
In Luria's early works (from the 1920s and 1930s), the ideas of Vygotsky
were fundamental. In one of the most recent publications [7] and also in some
Russian artIcles written with Homskaya [8], Luria emphasized that Vygorsky
had attributed special value to the limits within which the analysis of behavior
must be held. According to Luria [9], Pavlov had originally stressed the
importance of breaking down complex processes of behavior into their constituent parts, which had opened up possibilities for qualitative analysis. In
contrast, Vygotsky argued that breaking down an element of behavior into its
simplest parts does not conserve the qualities of the whole any more than, for
example, decomposition of water (H 2 0) into hydrogen and oxygen conserves
the original property of water. The procedure can only confirm that behavior

6. Neuropsychological Investigation

129

is made up of many isolated reflexes. The purpose ofVygotsky's analysis was


to break down a phenomenon not into "elements" but rather into "units"each of which conserved the properties of the whole in its simplest form. This
means that "behavioral units" are not clements of the reflexes themselves but
are "complex forms of mediated activity ... that arise in society and in history
and ... constitute the essential components of complex human mental activity"
[9]. The basic assumption of the Luria group, then, is that the higher psychological processes are social in origin, are structured through the mediation
of speech, and function consciously in a self-regulated manner [10]. The
essence of "humanness"-consciousness-develops as a result of people's
social experiences, but the foundation of the consciousness is greater than
merely the sum of its elementary reflexes.
According to developmental theorists, the child is not born conscious or
possessing voluntary activity. The higher psychological functions develop in
the course of the child's interaction with its mother, with other people, and
with the human-made environment-the culture-into which he or she is
born. Voluntary activity develops primarily through interaction, and later
becomes a means of organizing behavior in which commands and speech are
in ternalized.
The child's behavior is made to happen by the interaction with the mother,
who names objects and gives instructions. Naming things allows the child to
modify its environment-for example, naming can provide a means of solving
problems. This can occur when the child is able to name his or her own needs
(e.g., hunger).
2.2 Functional Factors, Brain Structures, and Effects of Lesions

Voluntary movements are an example of a functional system in which different cerebral structures participate in an integrated and complex way. Every
structure makes its own functional contribution to the whole functional system
of voluntary movements. Depending on which structure is damaged, there may
be a specific disturbance of this functional system.
The affected structure can be identified only after the symptom(s) have been
analyzed in detail. According to classic neurology, disturbances of voluntary
movements (apraxia) correspond simply to lesions located in the parietal and
anteroparietal areas. However, a more accurate analysis has demonstrated not
only that the structure of voluntary movements is extremely complex, but
also that such movements involve the integrated participation of more than
one brain structure [11].
Basic clements of the functional structure of a voluntary movement are 1)
the kinesthetic aJference (i. e., the combination of kinesthetic signals concerning
muscle tone, joints, etc., of the limbs in movement); 2) the synthesis oj visuospatial aJference (i. e., the combination of signals relating to the spatial coordinates of the limbs); 3) the kinetic organization (i. e., the consecutiveness and
the "melodic synthesis" of the movements); and 4) the intentional aspect (i.e:,

130

I. Theory and Intervention

the goal). Controlling these elements of voluntary movements are various


cerebral structures: the sensory areas of the postcentral cortex, the parieto,?ccipital areas, the basal ganglia and the premo tor areas, and the frontal lobes,
respectively. Different disturbances of voluntary movements (i.e., different
forms of apraxia) occur, depending on which of these areas is damaged; they
can be differentiated diagnostically into kinesthetic apraxia, spatial apraxia,
kinetic apraxia, and intentional apraxia [2].
2.3 Functional Units

Luria identified at least three principal "functional units" in the brain [4] and
argued that the participation of these functional units is necessary for any type
of mental activity. At the clinical level, disturbances of consciousness can be
attributed to, and differentiated with respect to, impairments in the operation
of these units.
Luria described these functional units in order of their localization:
The first unit maintains cortical tone and waking state and regulates these in
accordance with the actual demands confronting the organism. The corresponding structural parts of the brain are the reticular formation, the higher
regions of the brainstem, the thalamic region, and the limbic system. The
inclusion of the medial zones of the cerebral hemisphere in this unit also means
that inclinations and emotions are controlled from this unit.
The second unit obtains, processes, and stores information. It is located in the
lateral regions of the neocortex on the surface of the hemispheres and occupies
the posterior regions, thus including visual (occipital), auditory (temporal),
and general sensory (parietal) regions.
The third unit programs, regulates, and verifies mental activity. The structures of this unit are located in the anterior regions of the hemispheres anterior
to the precentral gyrus.
Cytoarchitectonic studies [2] have shown that all three units are hierarchical
in structure. In the second and third units, at least three cortical zones are built
one above the other. The primary zone of each receives impulses from or sends
impulses to the periphery, spreading excitation gradually and thus modulating
the whole state of the nervous system. The secondary zone processes information or prepares programs, and the tertiary zone carries out and controls
the program and is therefore responsible for the most complex forms of
mental activity. The last two cortical zones consist of isolated neurons that are
capable of sending single impulses along their long axons and that operate
according to the "all or nothing" law.
In the first unit, disturbances in the systems of instinctive "food-getting"
and sexual behavior (which also include the simplest metabolic processes
connected with respiration and digestion) may be due to lesions in the brainstem and archicortex. These are considered to be the first and deepest source of
activation of the brain; the second source has to do with the orienting reflex.
For example, disturbances that have the character of a generalized lack of tonus

6. Neuropsychological Investigation

131

are attributable to the lower regions of the reticular formation, whereas phasic
disturbances are connected with higher regions of the brain stem. The third
source of activation has to do with the coordinated work made possible by the
descending and ascending connections of the lower system of the reticular
formation of the brainstem, with the thalamus and the higher levels of the
cortex responsible for the formation of intentions and plans.
The disturbances in consciousness in these deep medial parts of the brain are
followed by affective changes and characteristic defects of memory that can
vary in degree. A generalized asthenia is present; reactions become slow;
fatigue develops rapidly; the voice may become "aphonic"; the emotional tone
is depressed or indifferent. Sometimes anxiety develops and can become so
acute that it takes on catastrophic proportions.
3 LURIA'S NEUROPSYCHOLOGICAL INVESTIGATION:
GENERAL CONCEPTS

The following section describes various aspects of "the Luria Neuropsychological Investigation" (LNI) [12]. The procedure grew out of the work that
Luria carried out in the 1940s and 1950s.
3.1 Theoretical Basis for the Investigation

The main purpose of studying what happens to the higher cortical functions in
the presence of brain lesions is to attempt to explain which syndrome of
disturbances of mental activity results from the fundamental defects. This
knowledge is essential in planning treatment for the patient. The disturbance
in mental activity in the presence of brain lesions is always a result of neurodynamic changes, which are characteristically found in nerve tissue that has
been damaged. Therefore, when we examine a patient, our results should
indicate not only the general pattern of change taking place in the mental
functioning, but also the neurodynamic changes underlying the disturbance.
Such information eventually assists us in the diagnoses of brain lesions (i. e.,
the behavioral changes determining the components of the brain lesions).
The behavioral changes observed with various circumscribed brain lesions
are often very similar. Furthermore, simple observations can merely indicate
some of the disturbances that affect the patient's general behavior and perceptual activity. Frequently we cannot establish the basic factors that are
responsible for these disturbances, nor can we evaluate and discriminate between symptoms that are due to various causes and that differ in their internal
structure. Special methods of investigation are required to establish the precise
components and the significance of a symptom, to describe the defect, and to
differentiate its underlying factors.
The investigative methods known from experimental psychology and psychophysiology create the background for the specific examination whereby a
defect can be demonstrated with the greatest possible clarity and whereby its
structural organization can be analyzed in the greatest detail. These methods

132

I. Theory and Intervention

constitute what we will henceforth call the neuropsychological investigation. The


neuropsychological investigation differs from psychometric tests because it
purports to analyze the defects qualitatively instead of formally and quantitatively. In contrast to certain approaches that focus on psychometric tests, it is
not based on a preconceived classification of "functions" in accordance with
contemporary psychological ideas that by no means always reflect the forms
of disturbance of mental processes that actually result from brain lesions.
However, formal structured tests do contribute to the data gathered in the
neuropsychological investigation.
The neuropsychological investigation is focused on "syndrome analysis"
and is concerned with the factual pathology of higher mental functions. It
differs from the factor-analytical approach by being directed towards the
investigation of the organization of mental processes in a single subject.
A number of constraints are attached to the use of the neuropsychological
investigation. Because they are dealing with actual patients in clinical situations, clinical neuropsychologists cannot single out a process that interests
them and study it under specially created conditions. In their task of diagnosing a patient's condition, they will not know which process to focus their
scrutiny on until they have made preliminary studies of the patient's mental
activity. Moreover, they cannot allow their investigation to take too long;
frequently, 30 to 40 minutes is the longest period of examination that the
patient in an acute state can tolerate. Later sessions of the examination may
increase to 11/2 hours at a time. In addition, techniques have to be appropriate
for use at the patient's bedside, where the examination must often take place.
The neuropsychological investigation in itself is only a component of the
entire clinical investigation of the patient. This includes a thorough anamnesis
(i.e., medical and psychological history, including observations about the
patient from the family and others); detailed observation of the patient's
behavior while in the consulting room or hospital; analysis of the neurological
symptoms; and a series of additional objective examinations-otoneurological,
ophthalmoneurological, roentgenologic, electroencephalographic, and biochemical. By these means, the foundations are laid for the topical diagnosis.
Like the clinical investigations, the neuropsychological investigation must be
based on sound ideas about the possible types of disturbance that may be
encountered in brain lesions. It is absolutely necessary that neuropsychologists
have solid knowledge of the syndromes of neurophysiology and anatomy that
arise from brain lesions in various locations. Otherwise, it will be impossible
for them to direct their investigations toward the discovery of one of those
syndromes.
A qualitative analysis of the structure of the discovered defect can be made
after there has been an investigation into the state of the individual "analyzers"
(i. e., the subject's auditory, optic, kinesthetic, and motor analysis and synthesis). Investigators must discern whether a particular defect is based on a
disturbance of relatively elementary components of the particular mental

6. Neuropsychological Investigation

133

activity or whether it is due to a disturbance in the organization at a more complex level of activity. The investigators must also decide whether a particular
symptom is the primary result of a disturbance of some special feature of the
functional system under investigation, or a secondary (systemic) consequence
of some primary defect.
3.2 The Rationale for Conducting the Neuropsychological Investigation

A qualitative analysis of the level of functioning of the patient is thus the most
important facet of the neuropsychological investigation. The patient's reactions
and responses to every interaction with the examiner or investigator have to be
scrutinized, so that the examiner can understand the dynamics of the functioning taking place. The investigation has to be structured in accordance with
this goal; to the degree possible it has to fulfil the character of an experiment,
in which the examiner varies and controls conditions in the attempt to clarify
the nature of the patient's psychological processes.
The examiner not only should identify the disturbed functions but also
should clarify the means by which the patient is trying to cope with the
problems he or she is presented with (as well as the way he or she is making
use of any intact functions in those compensatory efforts). The data we need
have to be extracted from what we can observe, and this can only be provided
through an exploratory clinical investigation. This calls for an individualized
examination of each patient. The content of the questions and tasks has to be
the same, but the time spent, the formulation of the questions, and the amount
of support given have to vary not only in accordance with the physical state of
the patient, but also with his or her premorbid personality characteristics.
Tasks may need to be reformulated or changed in order to elucidate conditions
under which they are solvable. After questions have been responded to, they
may be repeated in order to illustrate learning possibilities or they may be
changed, for example, to demonstrate the effects of emotional tone.
The reliability of the tests is also an important issue. In some of the latest
literature about test validation, a number of warnings against relying too
heavily on deviations from norms in individual cases have been expressed [13].
The variability between individuals seems to be increasingly acknowledged
[14]. The clinician may need to look for confirmation from other tests or seek
corroborative evidence from the daily life experiences of the individual.
Accepting variances is important, but more important is another trend that
has also been emphasized in the recent literature: the analysis of the component
parts of the tests being used [15]. In the neuropsychological investigation, the
tests for the initial examination are especially selected on the basis of their
importance in pinpointing the functional level of the basic analyzers of audition, vision, and motor and kinesthetic abilities in the brain. Furthermore, the
tests are selected in order to be as simple as possible, consonant with the
specific goals of providing clear information. These demands on the tests have
made it possible to compare the patient's behavior, when the tests are solved,

134

l. Theory and Intervention

with the results of the new imaging methods, blood flow (e. g., rCBF),
computerized tomography (CT scan), and magnetic resonance imaging (MRI).
The comparisons have given us a deeper understanding not only of the areas in
the brain where the activity is increased for a specific test but also of the
variations in normal people as well as in brain-injured patients. The simple
tests provide the possibility of analyzing a disturbed function from various
aspects. They also provide an opportunity to illuminate how a disturbance of a
specific component affects other areas of functioning [15].
The importance of a neuropsychological tool for localization has diminished,
but there is still a need for a thorough understanding of the inner structure of the
neuropsychological processes that take place and that are elucidated through
this kind of investigation. The meaningful rehabilitation planning depends
upon such information. Even the imaging techniques, especially the rCBF,
seem to be able to shed a light on the most effective rehabilitation procedures.
[15].
3.3 The Investigation: General Procedures

The neuropsychological investigation is preceded by a conversation with the


patient, during which information is obtained about the history of the present
condition as well as the general state and particular aspects of the patient's
mental activity.
The investigation begins with a series of preliminary tests. In this second
stage, a relatively large number of tests that reveal the various aspects of the
patient's mental activity have to be included, but the duration of each test must
be short. The tests must all be relatively standardized in character, and they
must include only those items known to be within the grasp of any normal
subject, even a relatively uneducated one. However, the complexity of the
tests must vary from patient to patient, in keeping with each patient's premorbid level. The primary aim at this stage is to discover the state of the
individual analyzers (optic, auditory, kinesthetic, and motor) and to ascertain
the various structural levels of the patient's mental processes, including 1) the
level of direct sensorimotor reactions, 2) the level of memory organization, and
3) the level of complex, mediated operations (in which a leading role is played
by the connections of the speech system). A disturbance of one or more of
these processes may be the direct result of a lesion of a particular zone of the
cerebral cortex.
The third stage of the investigation, the selective stage, must be devoted to a
more detailed exploration of any groups of mental processes for which the
preliminary tests have detected the presence of definite defects. Although it
must be built upon the foundation of results obtained in the first stages, it must
also take into account the facts that are obtained in the course of the second
phase of the investigation itself. This part of the examination is thus strictly
individualized: It is more complex, yields richer results, and calls for greater
flexibility in the conduct of the examiner.

6. Neuropsychological Investigation

135

The tests included in this part of the investigation must be integrated tests
that examine complex forms of activity, the performance of which may be
disturbed in different ways, depending on the functional systems built up by
the individual during his or her development, but also dependent on the type
of lesion. These tests examine repetitive and spontaneous speech, writing,
reading, comprehension of written materials, and the solution of problems, as
well as memory. Each of these complex forms of mental activity is accomplished with the participation of the group of basic analyzers from the aforementioned principal areas of the cerebral cortex, as well as the secondary
cortex. The difficulties experienced by the individual patient in the performance
of these tests will reveal the particular type of disturbance of the activity in
question.
The choice of adequate methods of investigation is of utmost importance in
this third stage, but so is the manner in which the experiments are carried out,
as well as the way the results are analyzed. There should be less concern about
whether a problem has been solved and more concern about the method by
which it has been solved.
It is not enough merely to carry out a particular experiment in a standardized
manner. The experiment must be suitably modified so that the conditions
making the performance of the test more difficult can be taken into account, as
well as those that enable compensation to take place. In other words, investigators must make all possible use of the highly discriminatory devices at their
command. Among the special methods that may enhance the analysis of the
neurodynamic defects are a change in the tempo of the investigation, presentation of the stimuli at a faster rate, or extension of the scope of the task-all of
which may easily induce a protective inhibition. Furthermore, it is important
to observe the development of fatigue during the course of the experiment; the
fact that fatigue does not develop uniformly in different types of activity,
especially during activity associated with different analyzers, may be particularly important. Examiners must also try to elicit functions that have remained
intact: They need to determine not only the residual forms of analysis and
synthesis being used by the patient in order to perform a task, but also ways in
which the patient reconstructs a disturbed activity by bringing into play
surviving analyzers and by transferring the solution of a problem to a level at
which the processes are carried out with the closest participation of the system
of speech connections. Investigators should consider variability and flexibility
to be requirements in the conduct of the examination; static standardized
techniques in these situations must be emphatically discouraged. Only if
these requirements are satisfied-requirements that demand knowledge and
experience-will this kind of clinical neuropsychological investigation prove
effective.
The fourth, and final, stage of the investigation is the formulation of a
clinical neuropsychological conclusion, based on the results obtained. The
fundamental defect must be identified, ways in which this defect is manifested

136

I. Theory and Intervention

in the various forms of mental aCtiVIty must be described, and, as far as


possible, the pathopsychological factors underlying the defect must be indicated. A thorough and complete analysis may then serve as a basis for an
assumption of a possible localization of a lesion being responsible for the
observed phenomena. Only then can investigators begin to distinguish the
relative importance of the general cerebral factors that are associated with
local lesions of the brain in an individual patient, as well as the extent to which
they are associated with the lesions.
How constant and reliable are the results of these tests? The statistical approach used in laboratory experiments-in which little variability in numerical
values is considered a sign of reliability-is impossible in a neuropsychological
investigation. The individual tests are usually administered only a few times,
the number being limited by the short duration of the investigation. The need
to carry out a large number of different tests in order to examine as many
aspects of the patient's mental processes as possible cannot be met; furthermore, if the same experiment is repeated with a patient for too long a time, the
results begin to change.
The reliability of the results can, however, be ensured by syndrome analysis-that is, the comparative analysis of the results of a group of assorted tests
and the determination, from these results, of general signs, which are then
grouped together into a unified syndrome. This grouping is possible because
the presence of a primary defect that interferes with the proper function of a
given part of the brain inevitably leads to disturbances of groups of functional
systems-that is, it leads to the appearance of a symptom-complex, or
syndrome, composed of externally heterogeneous symptoms (which, in fact,
are internally interrelated).
For instance, writing, like the pronunciation of words, has as one of its
components the reception of acoustic elements of speech. Operations involving spatial relationships or calculation, which externally are very different
functions, also possess a common link-simultaneous spatial analysis and
synthesis. The reliability of the experimental neuropsychological investigation,
carried out under clinical conditions, is thus obtained not only by way of
distinguishing the fundamental defect, but also by demonstrating how the
defect manifests itself through changed activity, coming to the fore in the
complex system of disturbances that arise. Hence, if the results obtained in
different tests are compared, and a common type of disturbance affecting
different forms of activity is observed, the results of the investigation can be
considered reliable and acquire clinical significance.
If all the aforementioned conditions have been met and a hypothesis regarding localization of the brain damage has been reached, assessment of
emotion and personality will be an important component of the general
clinical conclusion. In this way, the neuropsychological analysis in combination with the personality assessment can provide information about the braininjured patient that can further clarify the characteristics of the damage and its

6. Neuropsychological Investigation

137

consequences-functional as well as pathological. The latter aspect of the


investigation is bound to become increasingly important in the future.
4 THE PHENOMENOLOGICAL APPROACH OF THE INVESTIGATION

The restrictions and demands of any clinical examination hold true also for
Luria's Neuropsychological Investigation. Although the ideal experimental
situation-in which variables can be held constant, hypotheses can be formulated, and logical conclusions can be drawn-can never be fully realized in
the examination of individual patients, the clinician should strive to be as
precise and scientifically rigorous as possible, as well as critical of his or her
interventions. However, a flexible approach to the examination is important;
the neuropsychologist should be a skilled observer of the patient's reaction in
all situations and should be able to invent small experimental situations and
formulate hypotheses that can be verified or rejected by more specific examinations. Luria has compared the work of the neuropsychologist with that of a
detective; it has to be rigorous, but it is more complex and less logical than a
scientific experiment.
The neuropsychologist has to be aware of the individual differences that
may change the procedure. For example, patients may be in bed, unable to
move from a supine position, or may tire easily, so that their reactions may
give the impression of disturbances that are actually only secondary.
Patients' cooperation in the examination process is extremely important,
and one of the neuropsychologist's main tasks is to enlist such cooperation. A
means to this end can be to explain the purpose of the examination so that
patients can better understand their psychological functioning: what is easy for
them to do, how they do it, what gives them trouble and why, how they
succeed, and what resources they use. Once the patients have achieved some
knowledge about psychological processes and their development, the neuropsycologist may ask meaningful questions that foster confidence and belief in
the psychologist's power to help. Only if this background of cooperation
exists can the results of an investigation be considered sound and valid.
Cooperation is even more important in the planning of the rehabilitation
program: If patients have not cooperated fully in the examination, they are
unlikely to have confidence in the explanation of the results. If patients do not
recognize the disturbances of their behavior, they may not participate fully in
the program.
Some examples from Luria's Laboratory at the Bourdenko Neurosurgical
Hospital may serve to illustrate the examination process in Luria's hands. (The
author observed these examples on visits to Luria's Laboratory in 1973 and
1975.) The neuropsychological examinations were performed either in the
office, which the psychologists shared, or in the hall of the ward. This meant
that several people were present and could participate. At one end of the room,
Luria occupied his chair and table; beside him was an armchair for a guest, and
opposite was the patient's chair. The younger psychologists who occupied the

138

I. Theory and Intervention

other desks in the room often participated in the examination. Before the
patient arrived, one of them had undertaken the task of reviewing the available
information about the patient, including the results of examinations already
performed (e. g., otological, ophthalmic, EEG, neurological). This preliminary
information assisted the neuropsychologists in questioning the patient correctly, thus giving evidence of their concern and emphasizing their capabilities
for helping. The setting was a "social" one; the patient was shown interest and
offered help, and usually patients in this situation responded with accentuated
openness and trust.
The ability to create an optimal atmosphere for the individual patient is a
prerequisite for the neuropsychologist's investigation. Theoretical studies and
clinical experience with a variety of brain-injured patients have provided
guidelines for making the investigation procedure smoother, less tiring, and
less disturbing to patients [12].
Information obtained in this manner determines the course of the investigation. It is not desirable for all patients to be given the same questions and the
same tasks in the same order. Consideration has to be given to the specific
condition of the individual patient, the degree and extent of the lesion, and the
presence of any disturbances of consciousness, as well as the patient's age,
pre trauma life, interests, likes, and dislikes. This does not mean, however, that
the examination can proceed without order or plan. The course of the examination must be purposeful; every step must be carefully coordinated with the next.
For the neuropsychologist to gain full knowledge of the patient's characteristic
ways of functioning, he or she must be able to analyze the psychological
manifestations, make variations in the conditions, or repeat some of the tasks
for the sake of control or comparison, perhaps mentioning specific research
procedures.
The neuropsychological investigation can be divided into four stages. These
stages can be performed in an abbreviated manner, or over a period of time,
depending upon the complexity of the referring question.
4.1 Stage One: The Preliminary Conversation

The aim here is to obtain information about the history of the patients' present
condition, to evaluate the general functioning of the patients, and to define the
particular aspects of the patients' actual mental activity. The more careful the
attention paid during the conversation by the skilled examiner, the more
precise and meaningful the subsequent investigation will be. At this stage, the
neuropsychologist makes hypotheses with the intent of identifying pathological as well as intact processes. Subsequent procedures are designed to
confirm, modify, or refute the hypotheses that have been formulated.
The main areas of concern in this preliminary conversation are the patients'
consciousness, premorbid level, and attitude, not only toward the illness and
their actual situation, but also towards their surroundings. In addition, the

6. Neuropsychological Investigation

139

neuropsychologist must obtain information regarding the extent and character


of the patients' complaints.
The disturbances of patients' consciousness can be manifold. Generally
speaking, these disturbances can be attributed to injuries to the first of the
three areas of the brain that Luria described as "principal functional units"
(see Section 2.3 for a detailed discussion about those functional units). Identification of these disturbances is an important part of the neuropsychological
investigation.
Patients may have defects in orientation with respect to their surroundings,
the date, and the time of day. Confabulations may be noticed; the patients may
believe they are at home or at their job, or that their relatives are present.
When asked the name of the examiner, they may confidently give the name of
some person they know well. This is especially characteristic when the lesions
are in the anterior parts of the limbic regions. The memory disturbance reflects
the low tonus and is easily influenced by even the slightest distraction.
To determine patients' pre morbid level, the examiner may ask questions
about the patients' name, residence, family, and job, as well as about current
and past events in the news. If patients give inaccurate responses, further and
more specific questions have to be asked, to eliminate the effect of a possible
loss in selecting the right answer. This loss of the ability to select an answer
may be the result of lesions affecting the brain as a whole or by frontal or
frontotemporal lesions of various types.
The examiner not only should ask about patients' attitudes about their
illness, their actual situation, and their surroundings, but also should ascertain
the adequacy of the patients' self-estimation. For example, are patients able to
assess their own reactions? Do they lack introspection? Do they avoid recognizing their difficulties? Patients with frontal lobe lesions generally will not
admit their failures. Patients with lesions predominantly in the right hemisphere do not mention their defects, but, when confronted with them, admit
their presence and appear to be unhappy about them. Patients with lesions in
the left hemisphere know their difficulties, but they try to hide them by
compensating arduously.
Patients' complaints play an important role in directing the subsequent
stages of the neuropsychological investigation. The complaints may give evidence of the level of consciousness, or they may even have a character that has
a direct bearing on the topical diagnosis. The neuropsychologist listens to
what the patients say as well as how it is said. What is the impact of the
complaints on the patients' behavior and adaptation? What is the precise
content of the complaint?
Deficits have many facets. Memory disturbances, for instance, may be
manifested in very different ways-for example, the forgetting of intentions, a
difficulty in finding names, and problems of registering what is happening,
keeping track of events, or remembering what happened a moment ago or
what happened years ago (i. e., what used to exist as part of one's background

140

I. Theory and Intervention

and orientation). The onset of impairment is important: Were there special


conditions or other symptoms at that time? To the trained neuropsychologist,
the verbalizations of the patient may reveal aphasic disturbances even before
the patient is aware of them.
It is especially in this area that the neuropsychologist's broader clinical
experiences are important. He or she must be able to differentiate not only
between disturbances in functional systems caused by lesions in various brain
areas, but also between disturbances due to neurotic or psychotic (schizophrenic or manic-depressive) syndromes and difficulties caused by either
delimited focal lesions or more generalized diffuse injury.
If the examiner keeps in mind the first of Luria's units and its basic influence
on the overall functioning, the preliminary conversation can provide a great
many details that should be worked on and controlled for in the following
stages. Although the preliminary conversation is important, patients should
not be questioned for so long a time that they become tired before the
examination itself is performed. Especially when patients are tested at the
bedside, it may be necessary to base the hypothetical considerations on less
information, due to the limitations of clinical data obtained. In such cases the
demands for control, which are emphasized in Stage Four, must be observed.
4.2 Stage Two: Preliminary Investigation

This stage begins with a series of preliminary tests. These tests are short and
standardized at such a level that anyone who does not have an organic brain
lesion can perform them (including people with a poor formal educational
background). The complexity of the tests should be determined by the patient's premorbid level and should cover the various aspects of mental activity.
The primary aim at this stage is to discover the status of the individual
analyzers-visual, auditory, kinesthetic, and motor. These analyzers provide
information for the major areas in Luria's second functional unit, which is
responsible for registration, analysis, and memorization of information. These
abilities are located in the primary areas in the occipital, temporal, and parietal
lobes of the cerebral cortex.
The sensorimotor areas of the evaluation are included in the part of the
investigation in which the motoric analyzer is examined. It may be necessary
to analyze the means that the patient is using to perform a task and to change
the stimuli and the situations if the patient's responses are not fully comprehensible. This implies that the investigation cannot follow a strict scheme.
If the patient experiences difficulties in a certain area of functioning, the
examiner may want to investigate further within this area and test the elements of the functional system in which disturbances are suspected. It is also of
interest to evaluate whether a change in conditions, such as providing partial
or additional information, giving more time, or giving emotional support,
improves performance.

6. Neuropsychological Investigation

141

Another purpose of this stage of the investigation is to ascertain the various


structural levels of the patients' mental processes. This is done by examining
the level of direct sensorimotor reactions, the level of kinesthetic organization
of activity, and the level of complex, mediated operations (in which a leading
role is played by the connections of the speech system). A disturbance of one
or more of these processes may be the direct result of a lesion of a particular
zone of the cerebral cortex.
The motor functions of the hands and mouth (i.e., oral praxis) are examined. The tasks provide the examiner with much information. The kinesthetic basis for the movement is revealed and so is the dynamic organization.
Acoustic perception is examined with tasks that include tones and taps, melodies and rhythms. Finally, in this stage, information is sought about the
patients' perception, sensory stimulation, spatial experience, right/left orientation, body image, and possibilities for change.
4.3 Stage Three: Selective Investigation

The third stage is the selective part of the investigation. At this time, the
neuropsychologist examines in detail any groups of mental processes in which
the preliminary tests have detected the presence of definite defects. Testing
is thus based on the results obtained in Stage Two. Stage Three is strictly individualized, is more complex, and may yield richer results than the previous
stages. However, great flexibility on the part of the examiner is still called for.
The examiner must make use of the highly discriminatory devices at his or her
command. The tests included in this part of the investigation examine understanding of spoken words, sentences, and logical grammatical structures, as
well as repetitive and spontaneous speech. Patients' reading and writing skills
are examined, along with comprehension of texts, ability to solve problems,
memory functions, and other higher intellectual functions.
The examination of speech includes tests for receptive as well as expressive
speech. In practice, the two categories are examined at the same time, sometimes by the same methods. The tests for the examination of receptive speech
or language range from simple tests of phonemic hearing, word comprehension, and understanding of simple sentences, to more complex tests that
measure patients' understanding of logical grammatical structures. For
example, the following tests may provide information about receptive speech
functions: "Draw a cross beneath a circle; a square to the right of a circle but
to the left of a triangle," or "IfI had breakfast after having read my newspaper,
what did I do first?" [12].
The range of tests for examination of expressive speech goes from articulation of sounds, to repetitive speech, to nominative function of speech, and
lastly to narrative speech. Examples of these tests are repetitions of single
words, words in a row, complex sentences, naming from description, and
determination of categorical names.

142

l. Theory and Intervention

Finally, predicative speech (which expresses not only events but also relationships) in reproductive and productive forms is investigated. Examples are
presenting an action-oriented photograph to the patient and discussing a
specific subject in the picture.
Memory tasks also go from simple to complex, including interference,
retrieval, and various forms of recall. The learning process is examined by
asking patients to memorize a series of unrelated words. The patients' control
over the number of words they can process is a valuable part of this test. When
a patient has repeated the row of words several times, the examiner asks:
"When you hear this series again, how many words do you think you will be
able to remember?" The patients are assessed as to how accurately they predict
how many words they will remember. The most complex tests illuminating
higher intellectual processes have to do with concept formation at various
levels. An example of a task illustrating discursive intellectual activity is the
following: "A son is five years old. In 15 years his father will be three times as
old as he. How old is the father now?"
These complex tasks involve the secondary and tertiary areas of the second
functional unit. The more complex the tasks become, the more influential is
the role played by the third unit, which is dependent on the frontal areas and
has the function of programming, controlling, and regulating mental activity.
The effects of frontal lobe lesions are extensive; regulation oflevels of activity
may be disturbed and so may complex motoric tasks. Due to loss of intentions and plans, visual search may be incomplete, verbal tasks may be
responded to in an uncontrolled manner, and learning may lack organization
[3].
4.4 Stage Four: Formulation ofa Clinical Neuropsychological Conclusion

The fourth stage of the investigation is the formulation of a clinical neuropsychological conclusion based on the results obtained from the examination
of the patient and a comparative analysis of the data. The fundamental defect
must be identified; how the defect is manifested in the various forms of mental
activity must be described; and-as far as possible-the pathophysiological
factors underlying the defect must be identified.
A thorough and complete analysis may then serve as a basis for assuming
that a possible focal lesion is responsible for the observed phenomena. Only
then can the investigator begin to distinguish the relative importance of the
general cerebral factors that are more or less associated with localized lesions of
the brain. The analysis is not easy, and Luria [4] himself suggested two types
of investigation that complement each other to solve the problem of analysis.
Mecacci [11] described it in this way:
On the one hand, one must determine which are the various symptoms produced by
damage to a cerebral structure; on the other hand, one must determine which are the
disturbances produced by different lesions and structures into a particular functional

6. Neuropsychological Investigation

143

system. It has been seen that a lesion in the parieto-occipital area produces a serious
alteration of the visuo-spatial organization of voluntary movements (spatial apraxia).
This same lesion, however, produces disturbances of other functions, all implicating a
factor that is, so to speak, "spatial" (mathematical operations, logical rclations, etc.)
Other functions in which this factor is not concerned (comprehension of spoken
language, comprehension of music, etc.) are not disturbed by lesions of parietal
occipital areas. (p. 122)

A diagnostic proposal of a focal lesion is possible only if all functional


systems that include a disturbed factor are affected. The functional systems
that are affected by a lesion may vary, according to the individual lesion.
However, the consistency of disturbances within the functional systems is a
necessary condition for the reliability of a diagnostic suggestion.
Mecacci's citation [11] of Teuber's principle of double dissociation is used
as the tool for experimental control of the investigation in such situations.
Double dissociation requires that Symptom A appears in lesions of Structure
X, but not in lesions associated with Structure Y, and that Symptom B appears
with lesions of Structure Y, but not in lesions associated with Structure X.
Luria [2] has provided an example of double dissociation: He reported that if
there is a lesion in the left parietal area, but the left occipital area is left intact,
then the investigator would expect impaired performance on motor tasks, but
average performance on visual tasks. These tasks distinguish between the areas
of impairment.
Whenever such dissociation is lacking, specificity in the effects oflesions has
not been demonstrated. The task of the examiner is to prove that general signs
arc present in patients' responses to assorted tests and that these signs can be
grouped together into a unified syndrome.
If the neuropsychological investigation is performed by a skilled examiner
according to these guidelines, not only can it provide a topical diagnosis, but
(of even greater importance) it can provide insight into the behavioral disturbances associated with brain dysfunction in the specific patient and can assist in
the development of the rehabilitative program.
4.5 Case Illustration: H.H.

H. H. was 45 years old at the time of the rupture of a sacculate aneurysm at the
anterior communicating artery (March, 1987). From the local hospital, he was
transferred to the neurosurgical department of the major university hospital.
He was fully awake and oriented at the time of the transfer. CT scans confirmed bleeding from the subarachnoid space. A ventriculogram and an angiogram of the right carotid artery were performed, showing an aneurysm at the
right cerebral artery. The patient underwent a craniotomy the following day.
The aneurysm was broad-based, including almost the total communicating
artery, making it impossible to clip the sac of the aneurysm. Instead, the area
above the communicating artery was clipped. The immediate postoperative

144

I. Theory and Intervention

period was unproblematic, but after a few days H.H. became motorically
agitated. He later became disoriented and confused. A control CT scan revealed a small infarction in the basal parts of the frontal lobes. At the time of
H.H. 's transfer back to the local hospital, his family was informed about the
somewhat dubious prognosis.
The agitation was treated with chlorpromazine and haloperidol without
effect, and H. H. was transferred to a psychiatric ward. At the time of the
admission, he was confabulating without awareness of the operation and
without any insight into his situation. He had a clear memory of his earlier life
and eagerly discussed ethical and moral problems. Although his attitude in
the beginning was kind and friendly, he easily got into conflicts with his
"psychopathic and idiotic" fellow patients. During the stay in the ward, his
demented state seemed to progress; he walked slowly with his arms hanging
down, looking like a chronic psychiatric patient. The cause might have been
an overreaction to the medications.
H.H. visited his home several times during weekends. He recognized the
surroundings, felt at home, and managed far better than in the hospital.
During the subsequent period, he participated in occupational therapy individually for one to two hours a day. His main problems were memory
difficulties, but perseveration and lack of structure were also present. He was
only slightly aware of his situation. His wife, who was a nurse, had to handle
the severity of her husband's illness alone. However, she had been getting
good support from the hospital.
H.H. was then referred to the Center for Rehabilitation of Brain Damage at
the University of Copenhagen. Because three other patients of more or less the
same age and also of the same level of pre morbid intelligence were under
treatment at the Center, a special program was created specifically for this
group, called the "Group of Four, " starting in January, 1988; it was planned to
extend for 41f2 months, 4 days a week, 4 hours a day.
In accordance with the theoretical attitude at the Center, the very first step in
H.H. 's rehabilitation was a thorough evaluation with Luria's Neuropsychological Investigation.
During the preliminary conversation, H.H. appeared open, with an eager
gleam in his eyes. There was an alertness in his behavior that seemed to reflect
his original personality. Simultaneously, he interrogated the examiner, wanting to understand the intent of the questions in order to answer as thoroughly
as possible. He was oriented in time, place, and person. He was aware of his
current problems. His major difficulty was a memory problem, although he
had also been bothered by some mood swings. He was also able to provide
anamnestic data, but his presentation had a repetitious character, as if it were a
series of sentences that he had memorized by rote learning. The examiner's
questions made unpleasant disruptions in his presentation. The disruptions
made him start over again, and the information already given was repeated. In
situations like these he looked bewildered, but he soon collected himself. It

6. Neuropsychological Investigation

145

was quite clear from his statements that before the injury he had been a highly
intellectual man with command over his life and his working situation. He
seemed to have been able to react emotionally-for example, he talked about
his four children in a warm and loving way-but self-reflection did not seem
to have had a predominant place in his personality structure.
He described himself as ambidextrous; he had always written with his right
hand, but for practical matters like sweeping and shoveling, he used his left
hand.
From thefirst stage of the investigation (i.e., the preliminary conversation), it
was thus deduced that the subsequent testing needed to reflect H.H.'s higher
intellectual functions (e.g., the strateg~es characteristically used by him for
solving problems and presenting solutions). However, thorough investigation
of the processes of perception, coding, learning, and recall, as these processes
are represented in memory, seemed to be the most important element to
disclose in this case. Information had been obtained that his orientation was in
order, so far as the imminent questioning was concerned. However, his ability
to keep information in mind in a sequential order seemed to have suffered, and
taking in new information problematically affected his thought processes and
disrupted the strategies he formerly used for presenting old material.
Examining the analyzers or functioning of the primary areas in the brain was
the first part of the second stage. The patient reacted in a smooth and precise
way to all motor tasks; there was no tendency to perseveration, and his ability
to carry out an oral movement on command outside the real situation
suggested a preserved high level of organization of actions. Simple acoustic,
kinesthetic, and tactile tests caused no difficulties.
The visual tests showed a pattern in which all kinds ofless-complicated test
material were reacted to without any trouble. If complex visual input was
only available for a short period of time, however, the patient requested further information; he became bewildered, which caused lack of structure and
systemization.
The patient's high intellectual abilities were evident, and the functioning of
the primary areas was unproblematic as long as the input could be taken in
within a short span of attention.
The more individualized examination that is characteristic of the third stage
confirmed that H.H.'s level of verbalization and abstraction was very high.
The various tasks within this section (logical grammatical tasks and requirements to produce spontaneous, narrative speech) presented no difficulties, and
an additional high-level category test was solved accordingly. Reading and
writing were performed easily and fluently. Asked to retell what had been
read, the patient recalled almost verbatim about ten components of a story but
then declared that all the rest was totally lost. When he was encouraged to use
his logical abilities, it was possible for him to remember an additional number
of details.
Arithmetical tasks within the rather easy amount of material contained in

146

I. Theory and Intervention

the investigation were solved in a superior way, although tasks with more
components tended to be rejected or requests were made for repetition of the
problems. Learning, however, caused difficulties. The investigation of direct
retention of memory traces again gave evidence that the amount of input that
the patient was capable of retaining was restricted in all sensory spheres. Tasks
illustrating changes in colored figures, amount of figures, and kinesthetic
movements of the hands were only partly solved when the presentation time
was five to ten seconds, whereas the acoustic traces of rhythmic taps, words,
and short sentences seemed of slightly longer duration. Presented with a series
of completely unrelated words that was too long for his memory span, the
patient was able to remember five or six. Repeating the words to him twice
had no effect. However, when the repetition was combined with activation (in
the sense that the patient was asked to judge his own performance), the
number of elements remembered increased to eight.
The investigation of indirect memorizing gave equivalent information about
both memory and intellectual processes. A series of ten words had to be
memorized by using appropriate pictures as aids for each word. In the first
phase of a task in which words and pictures were presented, H.H. only
remembered four. When he was asked to explain the logical train of his
thoughts, the number increased to eight. In a variant of the task, H.H. was
asked to memorize a series of 14 phrases or abstractions (e.g., "a deaf old
man," "a hungry boy," "cause and effect") by drawing certain signs or
pictures to help him remember. He remembered five correctly and five in a
paraphasic way: "debts" became "debtor," "varsel" (Danish for "warning of
tragedy") became "advarsel" ("warning of danger"). Once again, activation
had a strong impact on performance.
Tests examining intellectual processes showed superior functioning. The
patient's understanding of thematic pictures and texts was excellent and so was
his concept formation. He had no difficulties in reasoning and developing
strategies when the task was presented in writing. He worked fast, but occasionally too fast, so that the solution became incorrect. When he was confronted with the incorrect solutions, it was discovered that the failures were
usually due to his lapses in memory.
In conclusion, the immediate impression this patient gave was of a premorbidly active, highly intellectual, and effective man, in charge of his life
and his emotions, but not necessarily much preoccupied with his psychological
structure.
The investigation gave no evidence of disturbance in the primary cortical
areas. However, his span of attention was severely affected. As soon as
material presented to all his senses became too complex, H. H. became bewildered and his functioning was disturbed.
The individualized, more specific examination gave evidence that his learning and recall were severely affected by his restricted attention span, and
he had not yet developed any compensation strategies. Bewilderment and
confusion added to his problems.

6. Neuropsychological Investigation

147

The examination showed, however, that whenever it was possible to activate the patient and to stress his awareness of the problems and his own
participation in solving them, his performance improved radically. Finally, his
main problems were his lack of precise awareness, his restricted attention span,
and his subsequent problems with the sequence of presented information. His
intellectual capacity was high, and his verbalization was excellent. The main
task for the rehabilitation had to be concentrated on developing the patient's
strategies to observe, to take in observations in a logistical sense, and to make
use of mental aids (in the form of notes, Dictaphone, etc.) in order to obtain
maximal structure.
5 LURIA'S THEORY OF REHABILITATION

5.1 Recovery of Function

Luria, who was trained in both psychology and medicine, worked with braininjury survivors in a rehabilitation center in the Urals during World War II.
His goals were to have the rehabilitation programs of his hospital rest on firm
scientific ground and to define precisely the loss of function associated with
injuries in various locations in the brain. In addition, he hoped to distinguish
between the primary disturbances (resulting directly from localized injuries)
and their secondary effects [3]. As noted, Luria emphasized the necessity of
reorganizing at least two distinct components in every brain injury.
First, disorders of brain function may be the outcome of the destruction of
tissue and its replacement by scar tissue. The results are irreversible functional
changes that are the direct results of a lesion. These functional changes can
only be restored by major reorganization of the cortical processes.
The secondary disturbance may be a result of temporary malfunctioning of
synaptic transmission. In these cases, Luria proposed, it might be possible to
restore functions to their original form during consistent physiological therapy
and treatment, called "deinhibition" or "deblocking." The idea was to increase synaptic activity by using drugs that facilitated synaptic transmission.
Luria suggested that small doses of neostigmine might be used to suppress
cholinesterase production for selected cases of brain injury. Various behavioral
methods can also be used to help these patients [2].
Luria reported that observing brain-injured patients can provide insight into
what may be called "substitutive compensation" -for example, the use of one
hand when the other is paralyzed, or the incorporation of the visual system
into the locomotor system if disturbances in the proprioceptive impulses
impair a patient's ability to walk [2]. If a patient uses a walking cane, tactile
sensations are incorporated into the act of walking, and then walking is
accomplished on the basis of a reorganized functional system. Luria refers to a
case (described by Gelb and Goldstein) of a patient who was able to identify
white and colored spots but was unable to recognize objects or letters [1]. The
patient compensated for this defect by substituting movements of the hand or
eyes for the impaired function of visual integration. By tracing out the contour

148

I. Theory and Intervention

of an object with a finger or with his eyes, he was able to "synthesize" the
object's structure and thus facilitate recognition. Similarly, by outlining
letters, he developed a new functional system that made reading possible.
Functional reorganization can take place in either of two basic ways. In one
way, the same functional system is transferred to a new level of organization;
this is referred to as "intrasystemic reorganization." It can be carried out on a
more primitive, automatic level, or it can be transferred to the level of higher
cortical processes-for instance, by employing speech. In the other way, the
patient learns to rely on a different functional system. This type of compensation is called "intersystemic reorganization."
The majority of compensatory mechanisms that develop belong to one of
these two types of functional reorganization. In both, recovery is brought
about by the incorporation of some new afferentation (i.e., information from
an undisturbed area of the central and peripheral nervous system) into the
disturbed functional system.
As a result of research on animals, the following relationships between
functional systems and afferents have been described: A functional system
cannot exist without a constant afferent nerve supply. Each functional system
possesses a particular group of receptors, which together form a specific
"afferent field" that ensures the normal working of the functional system [2].
The number of afferent impulses required for the working of any functional
system decreases with practice, so that only a small group of receptors is in
active use. One of these stands out as "the dominant receptor," and the rest
remain in a latent state, forming a reserve of afferent impulses for that particular functional system. The quickness and ease of reorganization that takes
place within a functional system may be related to a rich supply of afferent
impulses.
Structures that are much more complicated may be present in human activity. Leontiev [17] has shown that defects in one system (the proprioceptive
afferent system) may be overcome by input from a second system (the visual
one). The result of this reorganization yields a compensatory movement. For
example, the range of a movement of a patient's injured limb was determined
to be at a certain level, and when the patient was asked to touch a visible point,
the range could be extended. The range could be extended further by having
the patient reach for an object at a suspended height.
Several factors are important for the development of compensatory mechanisms. One of the most important factors is the location of the lesion. The
lesion may be peripheral, or it may be in primary, secondary, or tertiary areas.
In the primary areas, there will be a defect of a specific function, but all
complex afferent syntheses directing that function will still be present. Destruction of the secondary and tertiary cortical areas will cause increasing
disintegration. These areas presumably endow the excitation that arises in the
primary areas with a definite functional organization, generalize the excitation,
and prepare the excitation for participation in corresponding functional sys-

6. Neuropsychological Investigation

149

tems. Luria stated that observations of patients who had lesions of secondary
and tertiary areas showed that their functions are always of a generalized
character. For example, the parieto-occipital areas of the cortex cease to relate
to visual activity, but remain as areas for spatial and simultaneous organization
of experience. Similarly, the temporal areas remain primarily concerned with
the organization of successive sensory impressions, and the premotor areas
remain concerned with the regulation of successive motor impulses. The
injuries may vary in character, depending on the part that the destroyed area
plays in the integration of a functional system.
Intersystemic and intrasystemic reorganization can be useful after patients
have sustained lesions. Both types of reorganization are possible when basic
motivation is preserved and when patients can take an active part in the
rehabilitation process, can recognize deficits, and can make special efforts to
overcome them.
Luria stressed that disturbance of a function at a high level of integration
docs not necessarily imply a complete loss of function. In this connection, he
opposed the technique in which preserved automatic functions are focused
upon, because this would only lead to mechanical learning. He agreed that
these steps can be helpful in the very early phases after injury, but-if they are
not followed by conscious compensation-they will provide little further
assistance. Even if residual forms of an affected function disintegrate further at
the beginning of a retraining program, the reorganization may be proceeding
successfully.
After an injury within the brain, the reorganization of "functional systems"
is believed to occur in the same way that reorganization and compensation take
place if a peripheral organ is injured. If the injuries do not affect the apparatus
directly concerned and do not prevent the development of new functional
connections between the different parts of the brain, then compensation for a
defect by functional reorganization may be possible.
In the Jirst form of reorganization, the functional system reorganizes automatically and quickly, without the patient's being aware of it. (For example, a
patient breaks an arm and then uses the other arm for all motor operations.)
This is especially the case for rather elementary functional systems. In the
second form, there is an intrasystemic reorganization of the preserved links; this
can usually be achieved by special and long-term training and, it is hoped, will
lead automatically to an alternate method of operation. (For example, if a
patient has sustained phonemic disruption in language, the therapist returns to
the affected phonemic structures and trains the person on those). Finally, in the
third form, there is restoration by intersystemic reorganization. This requires a
long period of training, involving maximum participation of the patient's
consciousness, and only gradually leads to automatization. (For example,
when the motor system has been impaired but the language system has been
spared, the language system can be used to enable the motor act to be
performed. )

150

I. Theory and Intervention

As might be expected, the restored actIVIty requires great effort and is


carried out extremely slowly at first. The amount of effort reflects the degree
to which the recovered function differs psychologically from the original
function. The aid and direction of a therapist are usually necessary during the
first state of recovery. The therapist's role is to identify for the patient the
methods that will allow him or her to compensate for the defect while mastering specific tasks. The training has to be carried out systematically and must be
carefully guided. The intention is to support the patient's own compensatory
efforts and to prevent the development of behavior reactions that in the long
run may inhibit the processes of adaptation and learning.
5.2 Rehabilitation and Recovery

The patient's participation in the planning process is enlisted as early as


possible. The patient's attendance is considered important because, as Luria
stressed, continuous feedback between therapist and patient (e. g., about disturbed functions and the effect of training) is necessary for the rehabilitation
process. The structure of treatment is thus adapted to the patient's situation.
Family members are also included in the planning process as early and as often
as possible.
The second stage in the rehabilitation program has two aims: first, disinhibition or deblocking of temporarily reduced secondary areas of the brain;
and second, reorganization of disturbed functions. The disinhibition period is
necessary due to the acuteness of the injury. The staff spends a great deal of
time helping the patients facilitate and practice their preserved automatic
functions (usually lower level hierarchical functions), which the patients do by
singing well-known songs, playing games, doing simple arithmetic, or performing practical daily tasks. The purpose of these activities is to support
patients emotionally and to stimulate and challenge their motivation.
It is important to emphasize involuntary fluent reactions by means of tasks
such as drawing, reporting news, or writing notes. The activities are chosen in
accordance with the individual patient's habits and experiences. This period is
usually rather short. If it is successful, the patient will have developed a
realistic understanding of his or her situation and will be able to collaborate in
further planning.
The retraining of lost functions follows the deblocking period. The reorganization of these functions can be either intrasystemic or intersystemic.
Intrasystemic reorganization occurs when behaviors are integrated to a lower
level (i.e., become automatic) or when they are integrated to a higher level
(i. e., made more conscious by introducing language). Intersystemic reorganization occurs when an entirely different functional system has to be created. In
the latter case, specific cognitive tasks are planned so that elements from the
intact functions are coordinated in some way. For example, if a patient has
articulatory difficulties due to a sensorimotor deficit, intact visual functions
can be used for the treatment of expressive speech.

6. Neuropsychological Investigation

151

The training is always directed towards the patient's strengths. The same
steps are followed in daily exercises with the final goal of creating an automatized process that eventually can reach full integration. In fact, the program
is continued to the point of overlearning, to make the behavior systematized.
The process is discussed continuously with the patient, who receives feedback
after every session.
The training period for patients lasts at least six months. The patients are
then discharged to their homes for a period of readaptation. The neuropsychologist maintains contact with the patients, to watch development, to make
new assessments when required, and to initiate new training periods when
needed.
Rehabilitation programs that are developed for patients must be strictly
individualized Gust as the neuropsychological investigations were). There are
four m;uor rules or guidelines for planning a patient's rehabilitation program.
First, diagnostic qualification of the defect (i. e., a thorough analysis of the
disturbances combined with a precise knowledge of the intact functions)
should be made. Patients should be given exact and complete information
about their condition, since, for their successful training, it is necessary for
them to be fully aware of their defects and the implications those defects have
on various functions.
The second rule is that intact functions are made use of in the training of the
disturbed functions. Examples of this come from all spheres of functioning.
For instance, if a patient has a focalized occipital injury and is not able to see,
and therefore to copy, written words or drawings, there is no sense in making
him or her practice copying. Instead, the patient can be trained to use any
intact kinesthetic movements for performing a copying task. It likewise makes
no sense to train a patient with sensorimotor disturbances to articulate words
unless intact areas (e.g., functional visual areas) are included in the training.
The third rule involves using automatized, lower level functions. In the case
of a patient with a focalized occipital injury, "automatic writing" can be used.
For example, the patient may be asked to write his or her name, short
common expressions, and so on.
The fourth rule is that the program has to be systematized and then repeated again and again so that internalization occurs. The goal is a systematic
reorganization of functioning.

6 SUMMARY

Neuropsychological assessment and treatment represent an entire theoretical


model of approaching rehabilitation. Recent advances in the fields of neuropsychology and rehabilitation have pointed out the need for such an integrative approach. The qualitative assessment seen in the Luria Neuropsychological
Investigation, combined with what is found in the multiple modern approaches (e.g., Boston Approach to standardized tests [18]), combined with

152

I. Theory and Intervention

modern radiological techniques, provides a wealth of data on which to base a


plan for neuropsychological treatment.
Successful treatment and rehabilitation programs depend on a thorough
understanding of brain functioning, recovery patterns, and treatment and
rehabilitation concepts. Views about brain activity included from Luria, as
well as newer concepts about brain functioning, build the theoretical background for the work.
From this conceptual framework, theory, assessment, and treatmentlrehabilitation are a dynamic process. Because of the structure of the neuropsychological investigation, it is not only an examination per se but rather a
treatment technique.
The art of rehabilitation is still incomplete. It has not yet fully developed in
accordance with the possibilities at hand. Until now, rehabilitation has mainly
consisted of so-called "cognitive retraining," initiated on the basis of the
results obtained from testing the patients. Cognitive retraining has mainly
been performed as an attempt to train the disturbed functions of patients. In
some settings, the retraining has been combined with group psychotherapy or
individual psychotherapy. Such therapy is often of substantial benefit to the
patients.
The rehabilitation process is not static, but rather dynamic. Luria himself
was one of the first to recognize this and eagerly incorporated new discoveries
(e.g., Teuber's) into his thinking. However, there are some fundamental
principles that need to be agreed upon.
The first is the recognition of the examiner as an experimental clinician. This
clinician should be fully aware of the components of interaction, using himself
or herself as one of the most important tools fulfilling the aim of the evaluation or investigation. This awareness includes a thorough understanding of the
neuropsychological functioning of the group being examined, as well as of
personality functioning and rehabilitation/treatment concepts.
The clinician must also have a thorough understanding of the components
of all of the tests used in the neuropsychological analysis, as well as how the
analysis will be integrated into an assessment of general and psychological
functioning. Of greatest concern to the neuropsychological treatment after
brain injury is the explicit analysis of the psychological function of the individual
patient. How the person is making use of the intact areas of his or her brain,
and the compensatory efforts that have developed since the injury, are all
scrutinized to provide a partial framework for the recovery pattern and,
ultimately, the blueprint for treatment. In the future, the rehabilitation strategies developed by educators for retraining learning disabilities need to be
extended and improved to aid m the development of the compensatory
functions after brain injury.
Finally, the clinician needs to be committed to exammmg and exploring
methods that can elicit and utilize the personality structure of the braindamaged patient and should integrate this knowledge into the neuropsy-

6. Neuropsychological Investigation

153

chological treatment. The establishment of a relationship, as well as clear


communication and feedback, is essential in this process. If the client comes to
feel better about himself or herself, he or she may be motivated to try a new
method of rehabilitation that the therapist thinks would make a significant
difference in the person's recovery and level of functioning.
Of crucial importance is the collaborative bond between two human beings
(i.e., the patient and the therapist) directed towards the rehabilitative process
with the goal of a higher level of personal functioning for the patient.
REFERENCES
1. Luria, A.R. (1970). Traumatic Aphasia. Mouton, The Hague (English edition) (original work
published in Russian in 1947).
2. Luria, A.R. (1982). Higher Cortical Functions in Mall, 2nd ed. Basic Books, Inc., New York:
(original work published in Russian in 1962; first edition published in English in 1966).
3. Luria, A.R. (1963). Restoratioll of Brain Ftlliction Alier War Illiuries. Pergamon Press, London:
(original work published in Russian in 1948).
4. Luria, A.R. (1973). The Workillg Braill. Penguin, London.
5. Teuber, H.-K. (1966). Preface. In Higher Cortical Fllllctiol15 itl MOIl, 2nd ed. Luria, A.R. Basic
Books, New York, pp. xi-xiv.
6. Pribam, K.H. (1966). Preface. In Higher Cortical FUllctions ill Man, 2nd ed, Luria, A.R. Basic
Books, New York, pp. xv-xvi.
7. Luria, A.R. (1979). The Makillg of Milld, Harvard University Press, Cambridge, MA.
8. Luria, A.R. and Homskaya, E.D. (1966). The Frolltal Lobes alld Regulatioll of Psychological
Processes. Moscow University Press, Moscow. (in Russian)
9. Luria, A.R. (1979). Preface. In Bmill alld History. Mecacci, L., Brunner/Maze!, New York,
p. xiii.
10. Luria, A.R. (1966). HUlIlall Bmill and Psychological Processes. Harper and Row, New York (first
published in Russian in 1963).
11. Mecacci, L. (1979). Brain alld History. Brunner/Mazel, New York.
12. Christensen, A.-L. (1979). Luria's Neuropsycholo,~ical Illvestigatioll, 2nd ed. Munksgaard,
Copenhagen (first published in English in 1974).
13. Goldberg, E. and Costa, L.D. (1986). Qualitative indices in neuropsychological assessment:
An extension of Luria's approach to executive deficit following pre-frontal lesions. In Neuropsychological Assesslllfl1t of Nellropsychiatric Disorders, Grant, I. and Adams, K. M. eds., Oxford
University Press, New York, pp. 48-64.
14. Meier, M.j., Strauman, S. and Thompson, G.W. (1987). Individual differences in neuropsychological recovery: An overview. In Neuropsychological Rehabilitatioll, Meier, M.J.,
Benton, A. and Diller, L., eds., Guilford Press, New York, pp. 71-110.
15. Christensen, A.-L., Jensen, L. R. and Risberg, j. In press. Luria's neuropsychological and
neurolinguistic testing. Submitted to Journal of Neurolinguistics.
16. Christensen, A.-L. (1984). The Luria method of examination of the brain-impaired patient. In
Clinical Neuropsychology-A Multi-Disciplinary Approach, Logue, P.E. and Schear j.M., eds.,
Charles C Thomas, Springfield, IL, pp. 5-28.
17. Leontiev, A.N. (1928). The Developmfl1t of Memory. Krupskaya Academy of Communist
Education Press, Moscow. (in Russian)
18. Milberg, W.P., Hebben, N. and Kaplan, E. (1986). The Boston approach to neuropsychological assessment. In Neuropsychological AssessllleJIt of Neuropsychiatric Disorders, Grant, I. and
Adams K.M., eds., Oxford University Press, New York, pp. 65-86.

II. REHABILITATION PROGRAMS:


APPLICATION OF THEORY

7. INTERVENTIONS IN THE INPATIENT SETTING

M. ELIZABETH SANDEL

1 INTRODUCTION

In the inpatient setting, neuropsychological treatment of brain-injury survivors


cannot be understood without considering their medical treatment. The model
is a medical one, for obvious reasons. Patients in an inpatient rehabilitation
facility are often in a precarious medical condition. Their assessment should
include ruling out medical causes for continuing coma and lack of recovery as
well as ruling out neurological causes.
This chapter will address the delivery of services to patients in inpatient rehabilitation, with a focus on the organization of services for improving the
neuropsychological function of the brain-injured individual. The treatment
begins in the coma-emerging stage and extends to the development of an outpatient program, in most cases when the patient has achieved a level of at least
some independence. If the patient reaches a plateau before this goal is reached,
a residential program or extended-care facility may be the best alternative.
2 EPIDEMIOLOGY: THE SCOPE OF THE PROBLEM

2.1 Brain-Injury Incidence

Attempts to determine the exact incidence and prevalence of brain injury in the
United States or other countries have not been entirely successful for a variety
of reasons. For instance, many studies use inconsistent definitions of brain or
head injury. Some patients are arbitrarily excluded on the basis of etiological
factors or severity of injury. Hospital statistics are often unreliable, and not all
patients are admitted to acute care hospitals. Regional incidence varies.
157

158

II. Rehabilitation Programs

Nevertheless, the incidence of externally produced brain injury in the


United States is probably about 200 to 400 per 100,000 people, depending on
the region of the country, urban-rural population ratios, and age and socioeconomic factors within the population [1]. Based on 1984 statistics, Kraus and
his colleagues [2] have estimated the total number of new cases in that year
to be 470,000. They calculated the number of severe cases to be 40,000. The
National Head Injury Foundation (NHIF) estimates that 140,000 people die
each year from head injuries and that 700,000 require inpatient treatment [3].
2.2 Brain-Injury Disability

Determination of the exact incidence of disabling brain injury is also difficult.


However, these figures are more elusive because of the difficulty in defining
disability and because most of the data are not compiled in a systematic fashion
by rehabilitation centers, where most patients are evaluated and treated. Many
patients discharged from acute-care facilities and from emergency rooms are
never evaluated by a specialist in physical medicine, neuropsychology, or
brain-injury rehabilitation. Impairments and disabilities may be unidentified
and therefore may not be "counted" in epidemiological statistics.
Kraus [4] has attempted to arrive at an estimate of the incidence of braininjury disability in the United States by using the incidence rates from seven
incidence studies. He calculated the total number of brain-injury survivors discharged each year from hospitals by level of severity on admission as 320,000
(mild), 37,200 (moderate), and 16,800 (severe), for a total of 374,000. This
calculation is based on a brain-injury incidence of 200 per 100,000. Obviously,
the figures exclude patients not hospitalized, although they may have been
evaluated in an emergency room.
Based on disability prevalence figures, it has been estimated that the total
number of people in the United States who are living with disability after brain
injury is about 1,000,000 [5]. Probably more than 75,000 Americans sustain
brain injuries each year that result in long-term disability [4].
3 REHABILITATION INPATIENT PROGRAMS

3.1 Numbers of Programs

Many forces have joined to accelerate the development of brain-injury rehabilitation centers and programs in this country. The National Head Injury
Foundation, legislative mandates, and insurance industry responses have
played strong roles. NHIF estimates that there are currently approximately
500 programs in this country that are categorized as brain-injury rehabilitation
programs (J.M. Williams, Director of the Clearinghouse, NHIF, personal
communication, March 1988).
There is geographic disparity in the distribution of programs, with a concentration of brain-injury rehabilitation beds in urban areas. However, statistics
indicate that the need for beds is not being met even in large metropolitan
areas. In Houston, for example, a city with three major trauma centers, there

7. Interventions in the Inpatient Setting

159

are 875 new brain-injury survivors per year, and only 45 designated beds in
two institutions, with fewer than 100 persons admitted per year [6].
3.2 Admission for Inpatient Rehabilitation

Which patients receive inpatient rehabilitation? When are they referred for
inpatient services? Who determines which patients are appropriate candidates
for inpatient rehabilitation services?
Obviously, the answers to these questions vary, depending upon 1) the patient population, 2) the regional organization of trauma services, 3) physician
referral patterns, 4) the availability of rehabilitation beds and physiatrists (rehabilitation medicine physicians) to staff them,S) inequalities of insurance
coverage that affect admissions, and 6) differences in admission criteria from
one rehabilitation facility to the next. As Aronow [7] points out:
Traditional inpatient programs appear to treat only a subpopulation of severe TEl
[traumatic brain injury], selected primarily from the more severe end of the continuum
of severity, or those patients with evidence of physical disability. (p. 33)

Admission criteria for inpatient rehabilitation units and facilities vary,


depending on the services provided. For example, some facilities are able to
provide mechanical ventilation and tracheostomy care. Pediatric or adolescent
patients require staff members with special expertise. Other inpatient programs may not have staff members who can provide behavioral management,
or they may be ill-equipped to provide a suitable environment for patients
who need specialized neurobehavioral and neuropsychological treatment.
The status of rehabilitation medicine as a "shortage" specialty, which has
been established by the Graduate Medical Educational National Advisory
Committee study [8], contributes to the problem of rehabilitating the braininjured patient. In a recent survey of practice characteristics of physiatrists [9],
patients with pain complaints were listed as the number one diagnostic group;
in the typical work week of a physiatrist, physical medicine services were
provided twice as often (48%) as rehabilitation medicine services (22%).
(Conditions treated in a physical medicine practice include, for example,
low back pain, sports injuries, and arthritis. In the practice of rehabilitation
medicine, the physiatrist treats the patient in an inpatient setting at least initially, with a team of other professionals delivering various services to maximize
the individual's functioning. Typical patients treated using this approach include survivors of brain injuries, spinal cord injuries, and strokes.) Although
the survey did not address the question of how many hours physiatrists spend
treating brain-injured patients, the figure appears to be small. Since 1983,
only about a dozen physiatrists have received postresidency training in a
fellowship program in brain injury rehabilitation. There are approximately
50 members of the Brain Injury Special Interest Group within the American
Academy of Physical Medicine and Rehabilitation. The total membership in

160

II. Rehabilitation Programs

the Academy is over 3,000 (Lawrence J. Horn, MD, Chair, Brain Injury
Special Interest Group, American Academy of Physical Medicine and Rehabilitation, personal communication, March 1988).
In contrast to the regionalized system of care developed in the 1970s for
patients with spinal cord injuries, no service delivery system currently exists
for brain-injured patients in the United States. Development of such a system
of care would only partially ameliorate the problem, however. Although
spinal cord injury centers operate in 15 national geographic areas, each year
fewer than 10% of new patients with spinal cord injuries enter a system of care
in one of these centers [6].
3.3 Commission on Accreditation of Rehabilitation Facilities

The Commission on Accreditation of Rehabilitation Facilities (CARF), formed


in 1966 as a national, private, nonprofit organization by an alliance of disabled
people, professionals, providers of services, and public and private agencies,
has been instrumental in promoting quality care for all disabled people [10].
The Commission's role in developing standards for inpatient care of the traumatically brain-injured has been crucial. Nevertheless, the incentive to obtain
accreditation has not been strong enough to compel many facilities to seek
accreditation. Only 75 brain-injury programs were accredited by CARF as of
January 1988 (C.K. Pierce, Associate Director for Programs, CARF, written
communication). In 1986, that figure was 44. If the estimate of 500 braininjury programs currently in existence is correct, only 15% are accredited by
CARF.
3.4 Cost Considerations

Cope and Hall [11] studied the costs of rehabilitation of brain-injured patients
admitted "early" (before 35 days post-injury) and "late" (after 35 days). Inpatient hospitalization stays for patients admitted late were twice as long as
those for patients admitted early. Costs were comparably higher for this "late"
group.
In Aronow's study [7], brain-injured patients who received rehabilitation
services achieved better long-term outcomes than did patients who had not
undergone formal rehabilitation, even though the patients who received services frequently had more severe injuries. It is obvious that the costs to society
would be reduced if patients were able to be less dependent on others and had
a greater chance of employment.
3.5 Establishing a Continuum of Care

The organization of regional trauma centers and the process of accreditation


for these centers have initiated changes in the treatment of the brain-injury survivor. One of these changes is the alignment of trauma centers and rehabilitation units and hospitals. This often takes place in informal ways, although the
Model Projects for Comprehensive Services to Individuals with Traumatic

7. Interventions in the Inpatient Setting

161

Brain Injury through the National Institute on Disability and Rehabilitation


Research (NDRR) of the U. S. Department of Education help to solidify these
ties in a few settings. The establishment of a "continuum of care" provides a
context for research as well as a service for patients and their families.
Physiatric and neuropsychological evaluations should be completed early in
the patient's acute hospitalization (e.g., in the trauma unit itself). This allows
for the interaction of these staff members with families as well as the use of
early intervention techniques to prevent later complications, both physical and
psychological. Transfer to an inpatient rehabilitation unit or hospital as soon as
the patient is stable medically and surgically is imperative. The physiatrist
should determine whether an inpatient or outpatient program is appropriate.
If the posttraumatic medical needs and physical impairments require inpatient
management, the patient's neuropsychological deficits should be treated in an
inpatient setting initially, followed by outpatient treatment or a transitional
living program.
In federally funded Regional Spinal Cord Injury Systems, an orthopedist, a
neurosurgeon, and a physiatrist often see the patient as a team in the emergency
room on admission. This allows key physicians involved in the care of the
patient to become involved in his or her care from admission, and thus a continuum of care is established early through these professionals. Such a system
would promote the establishment of a continuum of care for the brain-injury
survivor as well. The involvement of the physiatrist as well as the neuropsychologist early in the patient's hospitalization permits greater planning for the
rehabilitative aspects of care in the postacute setting.
4. THE PATIENT POPULATION

4.1 Patients Emerging from Coma

Coma is defined as "unarousable unresponsiveness" [12]. The Glasgow Coma


Scale, developed by Teasdale and Jennett in the early 1970s [13], is used to
document depth of coma in the early stages of the patient's hospitalization after
brain injury. (See Chapter 2, Section 5.2.)
The Glasgow Coma Scale is less helpful when a patient is able to open his
or her eyes, a state incompatible with the usual definition of coma. When the
patient has no apparent responsiveness to external stimuli other than a generalized response to pain or noxious stimuli, with eye-opening, the person is in a
so-called "vegetative state." The multiplicity of terms in the literature for this
state of "unconscious wakefulness" is testimony to the confusion about its
diagnosis. These terms have included apallic syndrome, coma vigil, prolonged
posttraumatic unconsciousness, akinetic mutism, parasomnia, anoetic syndrome, and
posttraumatic catatonia [14]. The definition provided by Jennett and Plum [15],
when they coined the term persistent vegetative state, is most helpful. In describing these patients, they stated that "they have periods of wakefulness when
their eyes are open and move; their responsiveness is limited to primitive

162

II. Rehabilitation Programs

postural and reflex movements of the limbs, and they never speak .... What is
common to these patients is the absence of function in the cerebral cortex as
judged behaviorally" (p. 734). When this state lasts more than two weeks, the
term persistent may be applied [15]. Other authors suggest that the condition
must exist for a year before the term persistent is applied [16].
Many vegetative patients show inconsistent responses over time, and therefore this label should be used cautiously. Some patients will be motorically
disabled to such an extent that they cannot respond verbally or with movement. Consciousness may be inferred if a primitive communication systemfor example, eye-blinks-can be instituted. Many of these patients are dismissed as vegetative in the acute care hospital, but, in the presence of family or
skilled, dedicated therapists, will respond and reveal their awareness, albeit
limited, of the environment.
Unfortunately, one of the most difficult problems facing the physician treating the patient emerging from coma is defining the patient's level of alertness
and interaction with the environment. Often the diagnosis of vegetative state
is reversed by team consensus after many hours of work with the patient and
the development of a rudimentary communication system. The input from
the family can be helpful, but it needs to be substantiated by treatment team
members to verify the consistency of responses.
When treating the patient in a coma or vegetative state, the team must
expect the possibility of awareness and speak as though the patient understands
the content of conversations, at least in part. On a coma-emerging rehabilitation unit, a fundamental neuropsychological principle must be that patients
hear and see-until they are able to tell us that they do not. This philosophy
provides a therapeutic milieu for the family as well as for the patient.
In one study, electrodiagnostic testing of patients in both coma and vegetative states showed normal brainstem auditory evoked responses, prolonged
central conduction time, and diminishing amplitude of the N20 (central)
response on evoked potential testing; these findings are constant in patients
who are observed clinically to change from the comatose state to the vegetative state [17].
The patient with the "locked-in syndrome" (or "ventral pontine syndrome")
must be distinguished from the patient in vegetative state. The term lockedin is often applied loosely, lacking etiological or anatomic specificity, to
patients with severe motoric deficits. It should be reserved for those patients
with the classic picture described by Plum and Posner [12]. Typically, these
patients are victims of brainstem stroke with limb and pseudobulbar paralysis
but unimpaired consciousness. They arc able to communicate only by means
of eye-blinks and other eye movements. Bilateral interruption of corticobulbar
and corticospinal tracts, usually with preservation of sensation, produces the
clinical picture. These patients live with the horror of being totally physically
disabled but mentally intact. They require intensive psychological intervention
on a regular and ongoing basis. The staff must recognize that although the

7. Interventions in the Inpatient Setting

163

patient's motor pathways are damaged, his or her sensory pathways may be
intact. Analgesia must be given for any procedure producing intolerable pain.
In addition, decubiti (bedsores) may be extremely painful.
Some authors have argued that nonvascular causes, namely multiple sclerosis
[18] and trauma [19], can produce the same syndrome, without vascular insult
to the ventral pons. In another case, bilateral midbrain infarcts produced
a locked-in syndrome, perhaps justifying the use of the term in a variety of
clinical conditions, with diverse pathology.
Although this syndrome undoubtedly occurs after head trauma, very few
head trauma victims are simply locked-in in the true sense (i.e., with preserved
cognitive function). If an injury is of sufficient magnitude to damage the brainstem severely, it also probably exerts a force that damages the cerebrum to a
clinicially significant degree. One study [20] has documented that traumatic
hyperextension of the head produces specific lesions in the medulla and pons,
namely tears and hemorrhages in the pyramids from overstretching. The cerebrum escaped injury in 12 of 21 cases, but all cases were fatal. If the hyperextension force is strong enough, presumably the patient does not survive
because of damage to cardiac and respiratory centers.
The success of so-called "sensory stimulation" or "coma arousal" programs
for patients in a coma-emerging or vegetative state is claimed but unsubstantiated. No one would argue the importance of preventive measures to decrease
the incidence of contractures, decubiti, deep venous thrombosis, gastrointestinal bleeding, and infection. However, the use of "sensory stimulation
techniques" to promote arousal and improve outcome cannot be justified, unless research is conducted to substantiate the effectiveness of these techniques.
Rader, Alston, and Ellis studied the effects of a sensory stimulation protocol
on severely brain-injured patients in vegetative state (unpublished paper,
1987). They noted immediate changes in certain indices such as respiration,
pulse, blood pressure, eye-opening, and motor response, but no effect on "the
level of [cognitive] functioning ... as a result of interdisciplinary rehabilitative
intervention and sensory stimulation ... for 19 subjects over a three-month
period. "
Nevertheless, one can argue for the use of a protocol of this nature for
research and monitoring of patient responsiveness. Whyte and Glenn argue
that the goals for sensory stimulation include a provision of environmental
stimulation and a means for monitoring patient response [21].
Various pharmacological interventions have been suggested for use in
patients in a vegetative state. In one study [22], a patient with "akinetic
mutism" responded to lergotrile and bromocriptine (dopamine receptor
agonists) but not to L-dopa or methylphenidate (dopamine mimetics). (See
also Chapters 3 and 4.)
Special services must be provided for the families of patients who do not
emerge from the vegetative state. In many ways, the family becomes the focus
of treatment. Psychological services must be available for the families of this

164

II. Rehabilitation Programs

severely injured population with poor outcomes. Often these services need to
be available for long periods of time as the family attempts to move from the
initial crisis to at least a measure of resolution. Because final grieving cannot
occur until the patient dies, many families have significant difficulty reaching
any resolution. Denial is common in many family members for months and
even years. In rare cases, the denial is so extreme that the family member may
develop a special "communication system" with the patient, who is clearly
unable to respond to the environment or any individual when examined by
professionals. This constitutes a delusional system and may be highly resistant
to psychological or psychiatric treatment.
Quality-of-life interventions are a part of any program for the patient in a
vegetative or severely disabled state. Quality oflife in this context may simply
mean the provision of comfort, or it may mean promoting interactions
between family members and the patient. Therapeutic recreation specialists
trained in brain-injury rehabilitation often can suggest unique approaches to
improve quality of life.
After many months have passed without improvement in the patient's level
of awareness, the family of the patient may begin to seek information from the
physician or social worker about reasons for the lack of change. If the physician
believes, on the basis of the type and severity of injury, that recovery or even
slight improvement in the patient's level of arousal cannot be expected, the
time has arrived for discussions of providing" comfort care" -in other words,
nursing and therapeutic interventions to prevent pain and suffering, without
intensive treatment to prolong life. The decisions concerning "do-not-resuscitate" and "comfort care only" orders must occur after trust has developed
between the physician and the patient's family. As Berrol [23] states, "In
the final analysis the determination of proportionality between benefits and
burdens of a treatment should involve the traditional triumvirate of the patient, the family and the physician" (p. 285). Usually these discussions only take
place in regard to the patient at Level I or II (and occasionally III) on the Rancho
Los Amigos Scale (Rancho Scale) [24-25]. The Rancho Scale is commonly
used to evaluate emergence from coma (see Table 7-1). The physician must be
Table 7-1. Rancho Los Amigos Scale: Cognitive levels associated with traumatic brain injury
Levels of response
I.
II.
III.
IV.
V.
VI.
VII.
VIII.

None
Generalized
Localized
Confused, agitated, inappropriate
Confused, non-agitated, inappropriate
Confused, appropriate
Automatic, appropriate
Purposeful, appropriate

Sources: [24, 251.

7. Interventions in the Inpatient Setting

165

fully informed about state laws regulating such decisions and the policies and
procedures of his or her own institution. (For a further discussion of legal
issues, see Chapter 15.)
Medical criteria for establishing brain death, or irreversible cessation of
brain function (including the entire brain, i.e., cerebrum and brainstem), have
been discussed for years by many groups of researchers and medical practitioners, including the Ad Hoc Committee of the Harvard Medical School to
Examine the Definition of Brain Death, in August, 1968, and other later
groups [26]. Statements regarding the brain-dead patient cannot be applied
to patients in coma or vegetative state. The issue of defining life and death
becomes even more difficult when applied to the latter groups. Thus, decisions
about the withholding or withdrawing of treatment are equally more difficult.
A medical ethics committee is essential to monitor decision-making on these
issues. General guidelines for the establishment and functioning of these types
of committees have been developed by the Judicial Council of the American
Medical Association [27].
Those patients who do not enter the category of vegetative state as they
emerge from coma often pass through a series of stages that can be categorized
by use of the Rancho Scale (see Table 7-1). The behavior of patients who
become agitated during the period of coma emergence is much like that of
patients in acute confusional states from metabolic causes or those with acute
psychotic reactions. They may hallucinate, confabulate, and show total disorientation and paranoia. These patients appear to have prolonged periods of
posttraumatic amnesia, an index of the severity of the injury. One study [28]
demonstrated that agitation and restlessness portend a good recovery.
The medical treatment of the patient emerging from coma and displaying
agitation is focused on protecting the patient through the judicious use of
medications. Often, the use of medications to foster normal sleep-wake cycles
is sufficient. Some studies have demonstrated that medications, particularly
phenothiazines, may slow the recovery process [29]. Other studies [30] have
refuted this argument, demonstrating no significant difference in outcome
between a haloperidol-treated and a nontreated group of brain-injured
patients.
Many medications with psychogenic properties have been suggested for
use in this population of patients, and are reviewed elsewhere [31]. (See also
Chapter 4.)
The behavior of patients emerging from coma with so-called agitation is
remarkably similar to the phenomenon of akathisia, or motor restlessness,
described in the neuropsychiatric literature. In fact, the brain-injury rehabilitation literature infrequently defines agitation, and perhaps the phenomenon is
in many instances identical to that described as akathisia. Originally described
by Haskovec [32], akathisia was thought to be a psychiatric condition produced by anxiety and hysteria. However, more recently akathisia has been
described as a state that occurs following the use of neuroleptic drugs [33]. A

166

II. Rehabilitation Programs

postulated cause is the competltlve blockade of meso cortical postsynaptic


dopaminergic receptors [34]. Treatment consists of withdrawing antipsychotic
medication (or other dopamine antagonists such as reserpine or metoclopramide) and instituting measures to protect the patient and reduce environmental
stimulation. The use of benzodiazepines, beta-adrenergic blockers, and
anticholinergics has been recommended [33].
Agitation, the term most often used in the brain-injury rehabilitation literature to describe the motorically active patient, perhaps should be replaced
by the term akathisia, which may provide a basis for understanding the biochemical and anatomical origin of this phenomenon: damage to mesocortical
or mesolimbic dopaminergic pathways. Then, perhaps, other terms could be
used to describe other behavioral phenomena that involve motor as well as
verbal behavior. Considering the frequency of use of neuroleptics for the
agitated brain-injury patient in acute care settings, at least some of the patients
manifesting akathisia may be manifesting an iatrogenic form of the syndrome.
Patients emerging from coma constitute the most medically unstable group
of patients who survive a brain injury. They require a staff of nurses, respiratory therapists, and physicians trained to care for this population. Although
stable upon transfer to an inpatient rehabilitation facility, these patients can
become medically unstable at any moment, requiring emergency treatment.
Proximity to an acute-care hospital is mandatory.
The management of primary and secondary medical problems and the prevention of complications require the coordination of the medical, nursing, and
respiratory staff. This coordination can be accomplished by any physician
dedicated to the comprehensive treatment of these patients. The philosophy of
the physiatrist with a sense of primary responsibility for the patient is most
compatible with the demands of this role [35]. Medical treatment of the patient
in coma or vegetative state in a rehabilitation facility is reviewed elsewhere
[21].
4.2 Patients Undergoing Active Rehabilitation

Defining the neuropsychological syndromes most common to the population


of brain-injury survivors is difficult. Indeed, this is the chief reason that research in the area of behavioral management has often been scant or conflicting. The terms agitation, aggression, and behavioral dyscontrol, for example, have
no doubt been applied to a very neuroanatomically and behaviorally diverse
group of patients. These terms ought to be discarded, if useful research is to
be promoted. Controlled drug trials [36] are needed. However, identification
of behavioral and cognitive subgroups, described in consistent patterns by
researchers, is necessary to permit the advancement of our understanding of
the behavioral disturbances of this population and to determine which treatments are successful.
The availability of magnetic resonance imaging (MRI), positron emission
testing (PET), and cerebral blood flow (CBF) studies may allow neuroanatomy

7. Interventions in the Inpatient Setting

167

to be more closely correlated with clinical observations. However, the success


of this correlation will largely depend on the careful delineation of behavioral
and cognitive subgroups.
In arriving at a decision about the use of pharmaceutical agents for behavioral sequelae of brain injury, one can always cite at least one case report to
support the use of any psychotropic, or even "atypical" [37], medication.
More useful is the information gained from controlled trials with groups of
patients. A single-subject, controlled, randomized approach can also be
valuable [38].
Answers to the following questions provide information that is useful for
arriving at a consistent approach to the management of behavioral disturbances
after brain injury:
1. How is the "behavioral problem" best defined? The description of behavior may vary from observer to observer. Therefore, specific information must be culled before interventional decisions are made. This
information includes a) verbal (What does the patient say?); b) physical
(What does the patient do?); c) temporal (How long does the behavior
last?); and d) onset-termination (How quickly does the behavior escalate
or terminate?).
2. Was the patient comatose for any period of time, and if so, for how long?
How long was the period of posttraumatic amnesia?
3. What is the patient's cognitive profile? Is his or her attention and concentration affected? What is the degree of memory impairment? What is
the recovery curve for cognitive functions?
4. How does the patient's behavioral problem interfere with his or her
treatment? Is it a safety problem for the patient? Is it a safety problem for
other patients?
5. Does the behavioral problem respond to nonpharmacological treatment?
If so, what type of treatment (e.g., redirection or reduction in environmental stimulation)?
6. What medications is the patient taking that might contribute to either
cognitive deficits or behavioral problems?
7. Can precipitating circumstances be identified? Do these involve personto-person encounters or other encounters, such as experiences in frustration
or failure?
8. Are there manifestations of other psychic disturbances, such as depression?
9. Is there any premorbid history of childhood developmental or adult behavioral problems or a family history of behavioral or psychiatric disturbances? Does the person have a previous history of brain injury?
10. Does the patient have evidence, clinical or electrical, of a seizure disorder?
Has a search been made for temporolimbic seizure activity?
After this information is obtained, case-controlled, randomized studies of
pharmacological agents should be initiated in all centers involved in the treat-

168

II. Rehabilitation Programs

ment of brain-injured patients. Consistent data collection will allow for the
advancement of research in these areas, rather than simple recapitulation of old
data and anecdotal literature.
The behavioral dysfunction that occurs after traumatic brain injury most
often includes some form of behavioral or emotional dyscontrol. This type of
dysfunction often requires planning careful strategies at regular interdisciplinary
team meetings. Occasionally, it may become necessary to hold ad hoc meetings
to address shifts or accelerations in aggressive or erratic behavior.
Other types of dysfunction, such as denial, egocentrism, depression, and
withdrawal can be addressed by the team on a regular basis but rarely require
emergency meetings unless suicide is a concern. Overtly suicidal behavior
often precludes treatment on a rehabilitation unit or in a rehabilitation hospital,
unless the staff is trained to manage a suicidal patient. The development of
liaisons with psychiatric facilities to expedite the emergency transfer of suicidal
patients is important for any brain-injury inpatient program.
Other potentially self-destructive behaviors, such as hyperphagia and other
eating disorders, require behavioral modification approaches by the entire
treatment team, primarily nursing and dietary services.
The development of behavioral management strategies for patients in an
inpatient unit is the responsibility of the entire team. However, discussions
about the institution's philosophy and approach to patients with behavioral
dysfunction must take place on a regular basis among the members of the staff
who are responsible for institutional policy. Often the institutional philosophy
or policy regarding these types of patients changes as a facility evolves. Likewise, individual philosophies of staff members may differ, and these differences should be discussed. Ultimately, the policies and procedures of the
institution must reflect an institutional consensus on these issues. State laws
may grant more or less latitude in the development of policies-for example,
those addressing restraint of patients.
Every inpatient institution for patients with traumatic brain injury should
develop a statement of patients' rights. Patients need to understand their rights
and require education about their responsibilities within the inpatient setting.
Ideally, if the patient is able to read, he or she should be given a copy of this
information. General institutional policies may also establish a basis for linking
performance or acceptable behavior with certain privileges. For example,
regular participation in therapy may result in more opportunities to participate
in activities outside the building, such as community outings. These policies
must be consistently applied to all patients, however.
Each inpatient facility must develop a policy and procedure concerning the
use of mechanical and pharmacological restraint. Obviously, these policies
must conform with state regulations. Physical or pharmacological restraint
must always be considered a last resort but may be necessary to protect the
patient from injury.

7. Interventions in the Inpatient Setting

169

4.3 Pediatric Patients

The special problems of brain-injured children call for a treatment team with
specialized training. An excellent discussion of the treatment of the child with
brain injury is contained in a series of articles in a recent journal [39].
Relatively poorer neurobehavioral outcome after brain injury in childhood
and adolescence has recently been correlated with the presence of focal and
diffuse lesions on CT scans and a coma duration of greater than one month
[40]. As in the adult population, behavioral and emotional disturbances probably occur most often in patients with frontal injury.
5 THE STAFF

5.1 Medical Services

The medical care on a brain-injury rehabilitation unit or other rehabilitation


facility is usually coordinated by a physiatrist, but the physician may have
training in a field other than physical medicine and rehabilitation. The Standards Manual for the Commission on Accreditation of Rehabilitation Facilities
(CARF) [10] states:
In those programs where more than one physician is involved in performing the
functions with the same patient, there should be clear identification of physician
responsibility. In the rehabilitation setting, the rehabilitation physician should have the
final authority and responsibility for the care of the person served (p. 46).

Consultation with other physicians is often required because the survivor of


traumatic brain injury usually has multi-organ and multisystem impairments.
CARF requires that the following specialties be available on a consultation
basis: general surgery, neurology, neurosurgery, internal medicine, ophthalmology, orthopedic surgery, otorhinolaryngology, pediatrics, plastic surgery, psychiatry, pulmonary medicine, and urology, in addition to physical
medicine and rehabilitation [10].
The medical and rehabilitation aspects of patient care can be coordinated
best by the physiatrist, who is usually the physician identified as primarily
responsible for the care of the disabled patient. However, in many settings,
given the shortage of physiatrists interested or trained in brain-injury rehabilitation, another physician may assume the role of team coordinator and primary
physician. The physician must have knowledge in the area of neuropharmacology and behavioral neurology, regardless of specialty training or board
certification.
5.2 Nursing and Respiratory Care

The management of patients in any rehabilitation inpatient setting requires


expert care by nurses with training in both acute care and rehabilitation
nursing. In addition, when many patients emerging from coma and patients

170

II. Rehabilitation Programs

in a vegetative state are being treated in one unit or center, a respiratory


therapy staffis often required because of the many active and potential pulmonary problems of this population. Ventilator programs require a staff with
even more extensive expertise and should be structured so that pulmonary
medicine coverage is always available.
In an inpatient setting in which many patients have tracheostomies and
where there is the potential for respiratory or cardiac emergencies at any time,
the nurses and physicians should have critical-care experience and the ability
to respond to any medical emergency. For those who treat patients who are
emerging from coma, training in rehabilitative nursing is important, but
equally important is training in a critical-care or emergency unit. In an active
rehabilitation unit, training in rehabilitation nursing techniques becomes more
important.

5.3 Neuropsychological Remediation

Neuropsychological services are a crucial aspect of the treatment for the


survivor of a traumatic brain injury. In a brain-injury rehabilitation inpatient setting, these services are ideally available to provide the following
more specific types of evaluation and treatment: 1) cognitive evaluation and
remediation; 2) behavioral evaluation and the development of behavioralmanagement and behavior-modification strategies; 3) personality assessment;
4) neuropsychotherapeutic intervention; and 5) family counseling.
These services are provided by professionals with specific training in these
areas. The staff of neuropsychologists treating the brain-injury survivor must
individually or collectively be able to deliver these services. Because of the
broad nature of this view of "neuropsychological services," it may be necessary to designate certain professionals for certain "areas of expertise." For
example, a behavioral clinical psychologist may perform a very different
assessment and design a very different intervention strategy than the evaluation and treatment recommended by a neuropsychologist. These disciplines
need to be integrated into the general field of clinical neuropsychology, because of the depth and breadth of expertise needed with this type of patient.
Training the nursing staff to manage the neuropsychological and behavioral
aspects of traumatic brain injury is crucial to the success of any program. This
can be accomplished by the use of didactic programs, role-playing, and observation of behavioral interventions by well-trained staff. The use of a crisismanagement team can aid in the management of patients and the education of
other staff. One-on-one staffing for the patient requiring behavioral management mwo' be available at all times and can be provided by well-trained
nursing 01 uehavioral neuropsychology staff.
Behavioral neuropsychology "rounds" can be helpful for facilitating communication among team members concerning the management of patients
with behavioral dysfunction. These team meetings should be conducted by the

7. Interventions in the Inpatient Setting

171

behavioral neuropsychologist and physician trained in the use of pharmaceuticals in the brain-injured population and should occur on both a routine and
emergency basis.
Social service provides a key role in the counseling of families and in the
provision of emotional support to the p;1tient and family. Social workers
skilled in family therapy techniques can provide desperately needed services.
Co-treatment with a clinical neuropsychologist can be effective for some families. Supportive counseling for the family that denies their need for psychological services is often best provided by a social worker, whose role may
not be as "stigmatized."
Supportive counseling for families goes hand in hand with education, because in many families anxiety is caused by the lack of understanding about the
patient's problems and about his or her prognosis. The physician provides
education to the family, but nurse-clinicians or case-managers who have
experience and training in brain-injury rehabilitation can augment the educational process for families and patients. Resource information (e.g., from the
National Head Injury Foundation) provides reading material that can then be
discussed in greater detail with the physician or nurse-manager.
5.4 Cognitive Remediation

Cognitive remediation is the domain of more than one specialist on the


rehabilitation team. The coordination of cognitive remediation strategies is
crucial and should be achieved through the team conference or through the
designation of a "cognitive team leader" from one department. Typically, the
speechllanguage pathologist, the occupational therapist, the therapeutic recreation specialist, the special education staff member, and the vocational
specialist participate in cognitive remediation efforts under the direction of the
neuropsychologist and the physiatrist. The roles and approaches used by the
different disciplines may overlap, and therefore a delineation of responsibilities
is necessary. In the interests of cost-containment, the responsibility for cognitive remediation may need to be limited to one or two disciplines. The
physiatrist or neuropsychologist must assume a key role in decisions related to
treatment strategies and intensities, especially in an atmosphere in which
enthusiastic therapists may make individual rather than team decisions about
therapy.
In the United States, the speech/language pathologist usually provides the
linguistic-based assessment of mental function and plans strategies for improving the patient's communication, both in articulation and pragmatic language skill areas. The special education staff member offers background in
assessment of academic functioning as well as training and education in learning deficits, strategy formulation, and general cognitive deficits. This evaluation is used in conjunction with the neuropsychologist's assessment in order to
provide the best methods of retraining. In addition, the education professional

172

II. Rehabilitation Programs

can provide important realistic assesments of the patient's ability to perform in


an academic setting. Cognitive remediation strategies (e. g., the use of a memory
log) are often instituted by these team members.
The cognitive assessments performed by the occupational therapist include
the evaluation of visual-perceptual function, and these assessments are best
performed under the direction of a physician trained in neuro-ophthalmological
evaluation and treatment. In addition, the combined expertise of a neurooptometrist and a neurologist trained in eye disorders can provide a pragmatic
approach to these deficits and guidance to the occupational therapy department.
Occupational, recreation, and vocational therapists are often the team members most involved in promoting the patient's "deinstitutionalization" or
"community reentry." Therapeutic-recreation professionals provide cognitive
evaluation and remediation, often using methods not apparent to the patient.
The use of leisure activities to promote a sense of self-worth and enjoyment
should not be underestimated. The vocational specialist also uses the patient's
job itself as a "tool" for cognitive remediation.
5.5 Physical Restoration

Physical impairments are the rule rather than the exception after a traumatic
brain injury. These impairments are often the result of direct trauma to the
brain; however, orthopedic injuries and peripheral-nerve injuries are also quite
common. Secondary complications often occur before the patient reaches the
rehabilitation inpatient setting. These include contractures, scoliosis, physical
deconditioning, and decubiti. All of these conditions require treatment by a
team of professionals, including the physiatrist, physical therapist, occupational therapist, rehabilitation nurse, and therapeutic-rea creation staff member.
The treatment of each of these primary and secondary conditions is a subject
covered extensively elsewhere [41]. Physical restoration ultimately affects
neuropsychological outcome by promoting the patient's sense of independence
and well-being.
5.6 Vocational! Academic Reentry

Vocational or academic reentry is seldom accomplished immediately after the


patient is discharged from inpatient rehabilitation. However, the initation of
academic and vocational evaluations is appropriate for those patients with
a good prognosis for recovery. This facilitates the early identification of
factors that might promote or interfere with return to work. It also facilitates
appropriate placement for postacute rehabilitation.
In some cases, however, academic reentry can be achieved when a patient is
discharged from an acute inpatient rehabilitation setting. Such a placement
may be in modified classroom settings or with tutors in the home. The
inpatient rehabilitation setting provides a transition for some patients who will
return to school but who require more physical interventions and a structured

7. Interventions in the Inpatient Setting

173

setting that cannot be provided at home. The transition is best achieved if


special education personnel are on staff in the inpatient facility. These professionals work closely with school district professionals to foster academic
reentry.
5.7 Driving Ability Assessment

The importance of a driver's license to an individual's independence in industrialized societies does not need to be argued. State regulations concerning the
driving rights of a brain-injured individual vary. In Pennsylvania, the "Handicapped Driver Reporting Law" (75 Pa. C.S. 1518) requires that "all physicians
and other persons authorized to diagnose or treat disorders and disabilities
defined by the Medical Advisory Board must report to the Department of
Transportation, in writing, within 10 days," any patient with "epilepsy,"
"mental deficiency," "mental or emotional disorder," or "any other condition
which in the opinion of the examining licensed physician, could interfere with
the ability to control and safely operate a motor vehicle." This information
must include the patient's full name, date of birth, and address. In states such
as Pennsylvania, the physician may be held liable for injuries sustained by
individuals in an accident caused by a brain-injured patient who is under his or
her care for any of these conditions.
The American Medical Association offers information on the medical conditions affecting drivers [42]. It is the physician's responsibility to counsel the
patient about the impairments that might affect his or her driving. This should
be discussed prior to the patient's discharge from the inpatient facility. The
psychological effects of receiving this information must be treated with appropriate sensitivity. Reassurance and education are important. The opportunity to undergo a driving evaluation should be offered to the patient if he
or she is medically cleared to drive.
Any patient who has had a significant head injury should be assessed by an
occupational therapist skilled in evaluating driving competence. This usually
involves a predriving evaluation and an on-the-road test.
Visual-perceptual deficits, impulsivity, and impairments of judgment were
the most frequently noted problems of brain-injured drivers in one study
[43]. In the same study, traditional neuropsychological tests did not predict
fitness to drive. In another study, perceptual training was associated with
improved driving performance and correlated with the degree of improvement in perceptual skills [44].
6 COORDINATION OF INPATIENT SERVICES

6.1 The Comprehensive Care Plan

In most rehabilitation settings in the United States, a care plan is devised to


assist the team in coordinating the patient's treatment. This care plan is often
used as a means of facilitating discussions about the patient at team conferences

174

II. Rehabilitation Programs

Table 7-2. Common problems among inpatients with


traumatic brain injury admitted to a rehabilitation facility
Lack of responsi veness
Protein-calorie malnutrition
Dysphagia and other eating disorders
Decubiti
Seizures
Incontinence ofbowcl and bladder
Respiratory insufficiency
Visual-perceptual deficits
Impairment of other senses: hearing, olfaction
Behavioral dyscontrol
Potential for injury
Impairment in problem-solving and judgment
Denial of disabilities
Cognitive dysfunction
Language/communication dysfunction
Mobility deficits
Dependence in self-care activities
Pain
Disrupted family relationships

Table 7-3. Grading system for charting improvement in problem areas*


Example: visual-perceptual dysfunction
1. Perceptual deficits prevent performance of any visual-perceptual tasks.
2. Perceptual deficits prevent performance except with maximum assistance.
3. Perceptual defi>:its can be overcome with some assistance and supervision using strategies to
compensate.
4. Perceptual deficits do not prevent patient from independently compensating and performing all
tasks.
5. No perceptual deficits exist.

* Developed by G. Bergman and M. E. Sandel, Mediplcx Rehab-Camden, Camden, NJ, 1988, copyright
pending.

and as a method of charting the patient's progress and goals. The care plan is
usually problem-oriented, and the problem list varies with the needs of the
disabled population receiving treatment.
In the inpatient setting, this problem list can be quite lengthy. A sample list
is shown in Table 7-2.
This type of care plan is by its very nature interdisciplinary, because staff
members from many disciplines may treat the same problem but may use
different approaches.
A grading system may be used to designate the patient's status in each
problem area. This makes possible a systematic method of charting progress in
resolving specific problems. See Table 7-3 for an example of this charting
system.
6.2 The Program Prescription

The program prescription represents the physiatrist's orders for therapy for a
specific patient. This prescription should be individualized, because each pa-

7. Interventions in the Inpatient Setting

175

Table 7-4. Sample program prescription for active rehabilitation patient


1. Physical therapy
Range of motion to all extremities-active assistive
Muscle re-education
Bed mobility, sitting balance, transfers, standing balance, progress to ambulation training
when standing balance is adequate
Strengthening and conditioning exercises to all four extremities
Equipment evaluation
2. Occupational therapy
Upper extremity range of motion-active assistive
Self-care evaluation and treatment
Visual-perceptual evaluation and treatment
Training for dominance change
Equipment evaluation, with evaluation of orthopedic needs of both upper extremities
Home visit for architectural modifications
3. Speech/language pathology
Dysphagia evaluation and treatment
Communication evaluation and treatment
Cognitive-linguistic evaluation and treatment
4. Therapeutic recreation
Leisure education
Community reintegration
Adaptive aquatics program
Horticulture therapy program
5. Special education
Cognitive evaluation and retraining
Evaluation for academic reentry
6. Vocational services
Vocational assessment
Vocational counseling
Work adjustment training
7. Psychology
Individual psychotherapy
Neuropsychological assessment
Behavioral management
8. Social services
Family counseling and education
Family support group intervention
Discharge planning

tient is unique. Use of an open-ended form is usually the best way to avoid
"cookbook" orders. In such a program, treatment can vary in intensity and
direction, and patients can enter at different phases of treatment. Treatment
orders will vary in both the types of interventions and the number of hours of
treatment per week. A sample program prescription is shown in Table 7-4.
The program prescription outlines the needs, treatment modalities, and
amount of treatment for the patient.
6.3 Admission Conference

Shortly following admission, after the team members have been able to evaluate the patient, it is helpful to have a meeting to coordinate the patient's
program-for example, to design the comprehensive care plan and the patient's

176

II. Rehabilitation Programs

Table 7-5. Examples of group therapies for inpatients with traumatic brain injury
Environmental awareness group
Orientation group
Community living skills group
Exercise group
Pragmatic language group
Social skills group
Group psychotherapy
Work adjustment group

schedule and to determine the approximate length of stay. The patient's insurance representative should participate in this planning process (and should
be a member of the planning team from admission to discharge).
6.4 Team Conference

Team conferences for each patient are usually conducted weekly. This gives
the team an opportunity to coordinate individual therapies with other team
members and to set priorities. Decisions about the patient's readiness for
group treatment can also be made by the team at this time. A list of possible
group therapies for brain-injured patients in the inpatient setting is included in
Table 7-5.
6.5 Discharge-Planning Conference

A discharge-planning conference is helpful as the patient's discharge approaches. This allows the team to plan the final stages of the inpatient program
and to institute plans for the next stage of treatment, which might be outpatient or day treatment, transitional living, or even long-term placement in a
residential facility or nursing home.
6.6 Professional Roles

The attempt to foster the interdisciplinary nature of treatment planning carries


with it the risk of blurring professional boundaries. This may result in either
the "drop the ball" syndrome (one team member expects another to address a
particular problem) or in the "don't do my job" syndrome (one team member
has conflicts with another because of overlapping responsibilites and authorities). These issues need to be squarely faced, especially if they interfere with
patient treatment or staff satisfaction. A strong team leader, usually the
physiatrist, can facilitate constructive communication and resolution of these
issues in most cases.
7 CASE ILLUSTRATION: HILARY

Hilary is a 17-year-old girl who was leaving a school dance when she was
struck by an automobile traveling at high speed. Prior to her injury, she had
completed 10th grade at a local high school; she was an average to belowaverage student with reading difficulties. According to her mother, Hilary was

7. Interventions in the Inpatient Setting

177

being evaluated for a possible learning disability. She had competed on the
track team and was a gifted athlete, despite idiopathic scoliosis.
As a result of the accident, Hilary's left tibia and fibula were fractured, and
she sustained an acromioclavicular dislocation (dislocated collar bone). She
had lacerations of both legs, a left supraorbital hematoma, and fractures of
both top front teeth. When she was admitted to the Regional Trauma Center,
her right eyelid was drooping (ptosis), and her right optic disc was pale. Both
pupils were poorly reactive to light. Her Glasgow Coma Scale score on
admission was 6. An external fixation device was placed on her left leg. She
was initially placed on a ventilator, but eventually this was discontinued,
although the tracheostomy tube remained in place. A gastrostomy tube was
inserted for provision of adequate nutrition. Hilary's initial CT scan showed a
midbrain contusion and a small hematoma in the left frontoparietal region.
When she became medically stable, at about five weeks postinjury, she was
transferred to a rehabilitation facility. At that time, she remained in a coma.
However, she was beginning to show some motor restlessness (akathisia). Her
right eyelid continued to droop. Her right pupil was not reactive to light, and
the left was minimally reactive to light.
At about three months postinjury, Hilary began to demonstrate the ability
to communicate-for example, by using a hand-squeeze or by responding to
the command to "raise your leg." At the same time, she became very difficult
to restrain, and eventually she was placed in a Craig bed (a padded bed that
allowed her to rest on the floor). During this period, it was difficult to position
her properly because she required elevation with pillows for tube feeding.
Sometimes she sat upright or reclined in a wheelchair, and tube feedings were
given in those positions as much as possible.
During that time, the speech therapist began to work closely with Hilary to
increase her oral-motor strength and to evaluate how safe it would be to begin
oral feeding. Initially, she was able to take in only small amounts of food and
was unable to initiate a swallow. Gradually, however, she was advanced from
thickened liquids to a chopped diet and then eventually to a regular diet. This
progress in feeding took place over several months. A videofluorographic
study substantiated the safety of her swallow mechanism and gave the team
confidence to proceed with the feeding program.
Initially, Hilary was incontinent of bowel and bladder. After her transfer to
the rehabilitation facility, her Foley catheter was removed and she was placed
in diapers. Gradually, however, as she became more alert, she was able to ask
to be toileted. Over the course of several months in rehabilitation, she became
continent of both bowel and bladder.
Hilary eventually underwent placement of a rod in the left tibia, and the
external device was removed. She was then able to ambulate using a patellartendon-bearing brace, which decreased the amount of stress on the fracture
site. During the course of her rehabilitation, as she assumed an upright posture
and began ambulation training, her scoliosis began to worsen. X-ray films

178

II. Rehabilitation Programs

demonstrated the progression of her spinal curvature, and she was fitted with a
Boston scoliosis brace to prevent further progression of this curvature.
Once Hilary was alert and medically stable, she underwent a resection of
part of the clavicle on the left side, primarily because of pain secondary to the
dislocation. She was placed in an upper-body cast to prevent movement
during the process of healing. After several weeks, the cast was removed, and
Hilary continued to have more function in the left arm.
Hilary demonstrated increasing visual ability. She began to open her right
eye. She was evaluated by a neuro-optometrist who noted aberrant regeneration of the right third cranial nerve, with oculomotor problems and pupillary
defects. After glasses were prescribed, her vision improved significantly, and
she was able to see for functional tasks, including reading.
Neuropsychological evaluation revealed that Hilary had significant cognitive
deficits. Her Wechsler Memory Scale score was 57, in the mentally deficient
range. Her verbal abilities were superior to her visuospatial abilities, but both
were at a very low level. She had problems with fine motor control, visual
scanning, comprehension of verbally presented information, word-finding,
and learning new information. She frequently confabulated (i.e., she substituted
information when she was unable to remember).
At the time of her neuropsychological evaluation (at approximately five
months postinjury), her academic skills in reading, math, and spelling were at
the third grade level, or second percentile.
An MRI scan of her brain performed six months after her injury showed
encephalomalacia (softening) in the left parietal white matter, and abnormalities
of the splenium of the corpus callosum suggestive of hemosiderin (iron)
deposits. These findings were consistent with a shearing injury of the diffuse
axonal type.
At this time, Hilary's speech was characterized by deficient respiratory
capacity for phonation. The speech therapist taught her to pause frequently
and prolong her phonemes. In addition, she was given respiratory breathing
exercises designed to decrease breathiness and produce strong voicing. Hilary
had difficulty using the appropriate pitch, prosody, and intonation; she spoke
in a flat, hypernasal, mono pitched voice. Verbal cues and auditory feedback
from a tape recorder improved her speech. She spoke too softly for particular
situations, and her rate of speech was too rapid at times. Again, auditory
feedback and verbal cueing facilitated increased loudness and decreased rate of
speech.
During Hilary's final months in inpatient treatment, the psychologist focused
on identifying and exploring social relationships and family issues, as well as on
problem-solving concerning these issues. A related goal was to increase her
ability to accurately define feelings in various situations and to identify thoughts
associated with these difficulties. In contrast to her behavior in previous
months, she began to demonstrate more willingness to discuss sensitive issues.
She began to initiate discussion without prompting, although memory dif-

7. Interventions in the Inpatient Setting

179

ficulties sometimes interfered with this. She admitted to feelings of frustration,


anger, and hurt resulting from conflicts. Ways to express these feelings were
evaluated in therapy, and she was able to come up with modified ways of responding. Her socialjudgment in some instances was questionable, but she was
able to reflect on whether she had exercised good or bad judgment at various
times. The focus of therapy at discharge was on self-esteem, particularly selfevaluations about her manner of communication and her appearance.
Prior to leaving the rehabilitation setting, dental restorative work was completed, and both her front teeth were capped. As a result of the dental procedures and the overall improvement in her condition, including weight gain,
Hilary began to have an improved sense of self-worth.
Hilary was discharged approximately six months after herinjury. She weighed
120 pounds, her weight before the accident. She was walking with a straight
cane, with close supervision. The left lower leg brace had been removed. She
continued to have some difficulty with fine motor coordination but was fully
independent with all self-care activities. Her speech was dysarthric (imperfectly
articulated), but she was able to communicate intelligibly, and she had begun
reading. Her behavior was appropriate, and she was fully cooperative with
the rehabilitation program. She was discharged to an outpatient program to
continue work on improving her cognitive function, her psychological
adjustment, and her mobility skills. Her goal was to return to school within
6 to 12 months and to run track again. The treatment team and her family
remain cautiously optimistic.
8 CONCLUSIONS

The neuropsychological treatment of the brain-injury survivor in the inpatient setting of a rehabilitation facility requires a full understanding of the
medical and neurologic consequences of the injury and the impact of these
factors on the emotional and cognitive functioning of the individual. In other
settings, the medical aspects of the patient's care are either overriding, as in the
acute care hospital, or relatively less important, as in an outpatient program. But
in the inpatient rehabilitation environment, the neuropsychological treatment
of the individual cannot be divorced from the medical aspects of treatment.
For the coma-emerging patient, neuropsychological progress depends in
large part on the patient's medical stability. It is essential to rule out all possible
causes of coma, including, for example, infection and endocrine disorders.
The patient's recovery depends more on a stable medical course than on any
coma stimulation program.
As the patient recovers, as illustrated in the case report (Section 7), the
medical issues pale in importance, and physical functioning becomes the chief
focus of therapy. The more active patient then becomes more of a challenge to
the staff because of the potential for injury. The interaction of diminished but
improving awareness of the environment allows for more interventions but
also the chance for psychic distress and physical injury.

180

II. Rehabilitation Programs

Functional gains are possible only when the patient begins to have the ability
to interact effectively with the environment, both physically and cognitively.
It is at this point that the actual rehabilitation process begins. Prior to the
achievement of functional gains, the patient's recovery can only be recorded in
neurological terms, not rehabilitative or functional terms. The patient's awareness of his or her functional achievements as well as deficits is necessary for
further recovery in neuropsychological function.
The inpatient rehabilitati~n team should include physical therapists, occupational therapists, speech/language pathologists, psychologists, therapeutic recreation specialists, special education professionals, vocational therapists,
nurses, respiratory therapists, and physiatrists. Each member of the team
contributes to the patient's neuropsychological recovery by providing the
patient with approaches to compensate for his or her deficits and by fostering
independence.
REFERENCES
1. Thai, E.R. (1987). Initial management of the multiply injured patient. In Head Injury, 2nd ed.
Cooper, P.R., ed., Williams and Wilkins, Baltimore, pp. 34-50.
2. Kraus, J. F., Black, M.A., Hessol, N. et a!. (1984). The incidence of acute brain injury and
serious impairment in a defined population. Am. J. Epidemiol. 119, 186-201.
3. National Head Injury Foundation. (1987). Trauma: The Silent Epidemic (pamphlet), NHIF,
Framingham, MA.
4. Kraus, J.F. (1987). Epidemiology of brain injury. In Head Illjury, 2nd ed., Cooper P.R. cd.,
Williams and Wilkins, Baltimore, pp. 4-15.
5. Vogenthaler, D. R. (1987). An overview of head injury: Its consequences and rehabilitation.
Brain Injury 1, 113-127.
6. Committee on Trauma Research. (1985). Injury ill America: A Continui11g Public Health Problem
(Contract No. DTNH22-84-C-0781). National Academy Press, Washington, DC.
7. Aronow, H. V. (1987). Rehabilitation effectiveness with severe brain injury: Translating research into policy. J. Head Trauma Rehabil. 2, 24-36.
8. Lane, M.E. (1984). Preparing for the 1990's: A challenge to the speciality ofPM&R. Arch.
Phys. Med. Rehabil. 65, 740-741.
9. Gonzales, E. G., Honet, J. C. and LaBan, M. M. (1988). Physiatric practice characteristics:
Report of a membership survey. Arch. Phys. Med. Rehabil. 69, 52-56.
10. Commission on Accreditation of Rehabilitation Facilities. (1987). Stalldards Mallual for
O~~al1izatio11S Servillg People with Disabilities. CARF, Tucson, AZ.
11. Cope, D. N. and Hall, K. (1982). Head injury rehabilitation: Benefits of early rehabilitation.
Arch. Phys. Med. Rehabil. 63, 433-437.
12. Plum, F. and Posner, J.B. (1966). The diagl/Osis of Stupor and Coma. F.A. Davis Co., Philaddphia.
13. Teasdale, G. and Jennett, B. (1974). Assessment of coma and impaired consciousness: A
practical scale. Lancet 2, 81.
14. Bricolo, A. (1976). Prolonged post-traumatic coma. In Handbook of Clinical Neurology,
Vol. 24, Vinken, P.J. and Bruyn, G.W. cds., American Elsevier, New York, pp. 699-755.
15. Jennett, B. and Plum, F. (1972). Persistent vegetative state after brain damage. Lancet 1,
734-737.
16. Berrol, S. (1986). Introduction. J. Head Trauma Rehabil. 1, viii.
17. Hansotia, P. L. (1985). Persistent vegetative state. Arch. Neurol. 42, 1048-1052.
18. Forti, A., Ambrosetto, G., Amore, M. et al. (1982). Locked-in syndrome in multiple sclerosis
with sparing of the ventral portion of the pons. Ann. Neurol. 12, 393-394.
19. Britt, R.H., Herrick, M.K. and Hamilton, R.D. (1977). Traumatic locked-in syndrome.
Ann. Neurol. 1, 590-592.
20. Lindenberg, R. and Freytag, E. (1970). Brainstem lesions characteristic of traumatic hyper-

7. Interventions in the Inpatient Setting

181

extension of the head. Arch. Pathol. 90, 509-515.


21. Whyte,]. and Glenn, M.B. (1986). The care and rehabilitation of the patient in a persistent
vegetative state.]' Head Trauma Rehabil. 1, 39-53.
22. Ross, E.D. and Stewart, R.M. (1981). Akinetic mutism from hypothalamic damage:
Successful treatment with dopamine agonists. Neurology 31, 1435-1439.
23. Berrol, S. (1986). Condition for the management of the persistent vegetative state. Arch.
Phys. Med. Rehabil. 67, 283-285.
24. Hagen, C, Malkmus, D. and Durham, P. (1979). Levels of cognitive functioning. In Rehahilitatioll of the Head Illjured Adult: Comprehellsive Physical Mallagelll1'llt, Professional Staff
Association of Rancho Los Amigos Hospital, Inc., Downey, CA, pp. 87-88.
25. Malkmus, D. (1983). Integrating cognitive strategies into the physical therapy setting. Phys.
Ther. 63, 1952-1959.
26. Campos-Outcalt, D. (1984). Brain death: Medical and legal issues. J. Fam. Pract. 19,
349-354.
27. Council report: Guidelines for ethics committees in health care institutions. (1985).]. A. M. A.
253 (18), 2698-2699.
28. Reyes, R.L., Bhattacharyya, A.K. and Heller, D. (1981). Traumatic head injury: Restlessness
and agitation as prognosticators of physical and psychologic improvement in patients. Arch.
Phys. Med. Rehabil. 62, 20-23.
29. Feeney, D.M., Gonzales, A. and Baw, W.A. (1982). Amphetamine, haloperidol, and experience interact to affect rate of recovery after motor cortex injury. Science 217, 855-857.
30. Rao, N., Jellinek, H.M. and Woolston, D.C (1985). Agitation in closed head injury: Haloperidol effects on rehabilitation outcome. Arch. Phys. Med. Rehabil. 66, 30-34.
31. Cope, N.D., ed. (1987). Psychopharmacology. J. Head Trauma Rehabil. 2, 1-112.
32. Haskovec, L. (1909). L'akathisie. Rev. Neurol. (Paris) 95, 1107-1109.
33. Gibb, W.R.G. and Lees, A.J. (1966). The clinical phenomenon of akathisia. J. Neurol.
Neurosurg. Psychiatry 49, 861-866.
34. Marsden, CD. and Jenner, P. (1980). The pathophysiology of extrapyramidal side effects of
neuroleptic drugs. Psychol. Med. 10, 55- 72.
35. Felsenthal, G., Cohen, B.S., Hiton, E.B. et al. (1984). The physiatrist as primary physician
for patients on an inpatient rehabilitation unit. Arch. Phys. Med. Hehabil. 65, 375-378.
36. Cope, N.D. (1987). Psychopharmacological considerations in the treatment of traumatic
brain injury. J. Head Trauma Rehabil. 2, 1-5.
37. Horn, L.J. (1987). "Atypical" medications for treatment of disruptive, aggressive behavior
in the brain-injured patient. J. Head Trauma Rehabil. 2, 18-28.
38. Evans, R.W., Gualtieri, CT. and Patterson, D. (1986). Treatment of chronic closed head
injury with psychostimulant drugs: A controlled case study and an appropriate evaluation
procedure.]. Nerv. Ment. Dis. 175, 106-110.
39. Jaffe, K.M., ed. (1986). Pediatric head injury. J. Head Trauma Hehabil. 1, 1-96.
40. Filley, C M., Cranberg, L.D., Alexander, M. P. et al. (1987). Neurobehavioral outcome after
closed head injury in childhood and adolescence. Arch. Neurol. 44, 194-198.
41. Garland, D., ed. (1987). Orthopedic management. J. Head Trauma Rehabil. 2, 1-96.
42. Doege, T.C and Engleberg, A.L., eds. (1986). Medical COl1ditions Affectillg Drivers. American
Medical Association, Chicago.
43. van Zomeren, A. H., Brouwer, W. H. and Minderhoud, J. M. (1987). Acquired brain damage
and driving: A review. Arch. Phys. Med. Hehabil. 68, 697-705.
44. Sivak, M., Hill, CS., Henson, D.L. et al. (1984). Improved driving performance following
perceptual training in persons with brain damage. Arch. Phys. Med. Rehabil. 65, 163-167.

8. RESIDENTIAL TREATMENT

MEREDITH M. SARGENT

1 INTRODUCTION

1.1 Overview of the Problem

The incidence of head injuries in the United States is of epidemic proportions


(although the number of cases is still an approximate figure) [1, 2]. The incidence has also increased dramatically over the past 15 years, as improved lifesupport technology has resulted in the survival of severely injured patients. It
has been estimated that between 1972-1973 and 1974-1975, the incidence of
major head injury in the 17 to 44 age group almost doubled-from 351,000 to
609,000 cases, respectively. It is also estimated that currently a minimum of
400,000 cases occur annually in the United States. Adolescents and young
adults between the ages of 15 and 24 are most at risk. Individuals in this age
range who survive traumatic head injury constitute a survivor group whose
members are severely disabled at a very young age, have most of their lives
ahead of them, and are faced with multiple deficits. There is a growing recognition of the need for rehabilitation programs directed toward improving
functioning of these patients.
1.2 Need for Treatment

Four or five years ago, few facilities were concerned with the long-term treatment of brain-injured patients. Staff members in acute rehabilitation hospitals
were just beginning to recognize that needs of brain-injured patients were
different from those of other disability groups and were beginning to address
183

184

II. Rehabilitation Programs

those special needs. Today, hundreds of programs in the United States offer
services to the brain-injured. For this reason, it is essential that we begin to
identify factors that contribute to effective treatment. As health costs have
increased, interest in applying cost-benefit analyses to rehabilitation has grown.
The documented costs of head injury ($3.9 billion in 1980) make this a particularly critical issue.
Analyses of programs for people with other chronic diseases indicate that
rehabilitation that has a goal of decreasing dependency has been extremely
cost-effective. To date, the field of stroke rehabilitation has received the most
attention with respect to cost-benefit issues. For example, when patients who
had been living in nursing-home facilities or other institutions with no services
were admitted to rehabilitation facilities with appropriate treatment, the degree
of improvement and resulting discharge from institutions justified the cost of
providing therapy [3]. Similarly, spinal-cord-injury patients showed lower
nursing costs and improved employment rates when they were treated in comprehensive, multidisciplinary spinal-cord-injury centers than when they received unspecialized care. Moreover, there appeared to be greater benefit for
patients who were referred early rather than late, although equivalence in
severity of the two groups was not established [3].
Data show that early referral of severely brain-injured patients to an acute
rehabilitation setting results in their reduced need for services, with a potential
savings per patient of up to $40,000. This additional money would otherwise
have been spent for acute hospital care [4]. In this study, the early and late
referral groups were matched on indices of severity.
After acute rehabilitation, brain injury as a disorder is similar to other chronic
diseases, in that those patients who are unable to be managed at home are
generally sent to nursing homes or institutions. This is an inadequate solution,
not only in terms of cost but also for quality of life. This scenario is especially
likely for the severely disabled, who leave the acute rehabilitation setting
before reaching their maximum level of functioning and independence. Also,
the emphasis on physical deficits during acute rehabilitation does not prepare
brain-injury survivors with new social and cognitive skills to compensate for
deficits in those areas.
After they return home, brain-injured individuals are isolated, lonely, frustrated, and bored by their inactivity. They tend to lose any gains they have
made and become more dependent [5]. In addition, many brain-injury survivors may have appeared to be functioning at a higher level in the hospital
than was really the case. (For example, a hospital setting provides order,
structure, and routine in an organized environment; corridors, bathrooms,
floor surfaces, and other aids minimize the impact of physical deficits. These
conditions are tremendously helpful to brain-injured individuals).
Discharge to the "real world" is accompanied by tremendous confusion; the
patient is overwhelmed by stimulating sights, sounds, activities, and interactions. Limitations imposed by physical deficits become pronounced because

8. Residential Treatment

185

the individual must maneuver on all types of terrain and in a variety of settings.
Former friends stop visiting, as they either move on to new activities or are
uncertain about how to respond to the differences they perceive in the survivor.
The severely disabled patient is also unable to resume academic or vocational
activities. All of these factors contribute to the increasing loneliness, frustration,
boredom, and dependency on the family.
On their part, family members become extremely stressed as they attempt
to manage the brain-injured patient at home. This is a particularly demanding
task if the individual is physically or verbally aggressive. However, the braininjured person who is passive and who lacks the initiative to become involved
in tasks can be equally wearing; someone must be constantly interacting with
him or her. Early referral to a rehabilitation program, after the acute injury has
been treated, may prevent or at least minimize these stresses on the braininjured patient and the family. It is hoped that additional rehabilitation will
also make it possible for brain-injured people to live in less restrictive settings
(i.e., a group home or supervised apartment, as opposed to a nursing home or
institution) .
With the support of the National Head Injury Foundation, families as well
as professionals in the rehabilitation field are calling for effective postacute
rehabilitation for brain-injury survivors. As the number of patients needing
acute brain-injury rehabilitation services increases, there is also an increased
awareness of the benefits of early intervention (e.g., improved quality oflife,
as well as reduced costs) [4].
1.3 Theory of Change: Functional Adaptation
There are many explanations of recovery after brain injury. The relevance of
any specific mechanism of recovery may be tied to the length of time since the
injury and to the age of the patient. For example, such physiological processes
as edema (swelling), diaschisis (inhibition offunction in a region of the nervous
system due to a localized injury in another region with which it is connected
by fiber tracks), or inhibition (physiological suppression of neuronal activity)
are time-related [6]. They may be active immediately after the injury; some
functions return as these initial processes become resolved. On the other hand,
anatomical reorganization (whereby parts of the brain take over functions
originally performed in a part of the brain damaged by the injury) is an example
of age-related recovery; it seems to occur primarily in young people. It appears
that, after the brain reaches physical maturity, it is less able to assume a given
function in an area unrelated to that function.
In addition, mechanisms such as axonal regeneration or collateral sprouting
(attraction of neural sprouts from adjacent intact axons) are also assumed to
be limited in scope. As people come to recognize that these mechanisms are
limited, they become increasingly aware that recovery cannot depend solely
on these physiological changes.

186 II. Rehabilitation Programs

The mechanism of functional adaptation seems to be a more complete explanation for long-term change. In functional adaptation, the survivor uses an
alternate means to perform a specific action that a damaged portion of the brain
used to perform (for example, writing down and taking along a list of possible
topics for conversation on a dinner date rather than relying on memory). There
is some question about whether this mechanism will occur spontaneously.
However, brain-injured individuals can employ these methods deliberately to
help themselves compensate for deficits [6, 7]. For this reason, rehabilitation
programs should concentrate on teaching compensatory strategies to braininjured people [8]. Even though they may hope that their pretrauma skills will
be recovered, patients must learn to use their remaining strengths to function
as independently as possible.
Many of the techniques that have been used in treating brain injury can be
understood as building on the fundamental principle of functional adaptation.
Indeed, any substitution of one method of reaching a goal for a different method
utilizes this principles-for example, using pictures on signs instead of words,
writing memos instead of relying on memory, using a calculator to replace
basic mental arithmetic, and so on. Each of these strategies allows the braininjured person to perform a specific, important behavior with alternative means
that rely on remaining skills.
The emphasis here is both on an individualized program (i.e., determining
for each person the most important goals of rehabilitation and the specific
means of reaching those goals) and on function (i.e., focusing on practical
behaviors that will enable the person to improve the quality of his or her life).
2 TYPES OF PROGRAMS

2.1 Background

In the past, many brain-injury survivors were discharged directly to their


homes, with few recommendations for future services and little information
about how to acquire them. Today, such individuals are more often referred
for treatment to a postacute setting or outpatient program directly from the
acute rehabilitation hospital. Initially, as with many people who have survived
traumatic occurrences, the brain-injury survivors and their families were
advised that it was "enough" that the person had survived. In many instances,
the brain-injured individuals, especially when they were not physically disabled, were expected to resume their former lives.
When they were unable to do this, their families were often forced to turn to
psychiatric facilities for help. Many of the symptoms exhibited by brain-injured
people appeared to be similar to symptoms experienced by different psychiatric
populations [9, 10]. However, treatment modalities utilized in traditional
mental-health facilities were often ineffective with brain-injured individuals,
primarily because of their cognitive deficits [11]. Such deficits-including
impaired memory, disorientation, speech-language deficits, poor attention or

8. Residential Treatment

187

concentration, and disorganization-rendered traditional individual, group,


and milieu therapies all but useless.
Another aspect of brain injury that distinguishes it from many other disabling
conditions is the abrupt nature and development of severe disabilities in individuals who had often been functioning normally until the time of injury. The
discrepancy between old and new selves has to be understood and addressed
by individuals who are cognitively limited. Again, psychiatric facilities were
not prepared for this type of patient, especially those facilities that often offered
residential treatment to patients with long-standing problems.
Finally, the families of brain-injured individuals generally want and need to
remain very involved in all aspects of treatment. Residential treatment facilities
often have not developed a formal procedure to include family members in the
program. However, this inclusion is essential, not only because it fulfills the
immediate needs of the family but also because, in the future, it is the family
that will most often resume care and support of the brain-injured individual
after his or her discharge [12].
Because the available facilities were unable to meet the needs of brain-injured
people, new programs were developed. However, the rapid increase in the
number of programs in a field in which there has been, in the past, little formal
staff training and preparation has created major problems. Recently, a number
of training programs have been developed for staff who work with the braininjured. It is critical that knowledge and experience about what constitutes
effective treatment for the brain-injured be communicated, to make treatment
successful.
2.2 Basic Program Components

Services provided by postacute rehabilitation programs should involve the


following stages: assessment of the patient's functioning at the time of admission; development of a treatment plan based on the evaluations; implementation of the treatment plan; continuous monitoring of progress; planning for
discharge; discharge; and followup [13]. Beyond these very general guidelines,
more specific needs, in terms of staff and treatment, have been delineated for
acute rehabilitation. These include 1) a comprehensive team available to evaluate and consult on each case; 2) an interdisciplinary team that plans and manages
an individual program for each patient; 3) an integrated treatment philosophy
to address all problems of the individual (i. e., behavioral, cognitive, and emotional, as well as physical, needs) [14]; 4) use of admission, treatment, and
discharge criteria that are clear and measurable; 5) a well-defined method of
documenting assessments and progress reports; 6) commitment to long-term
followup, especially to provide support during the first few months after discharge; and 7) communication and cooperation with community agencies,
with the ultimate goal of facilitating the patient's interactions with and return
to the community. Whatever the specific type of facility under consideration,

188

II. Rehabilitation Programs

these basic characteristics will promote improved care for patients and enhance
communication among staff members, patients, families, funding agencies,
and the community.
2.3 Descriptions of Programs
2.3.1 Head Injury Task Force

The Head Injury Task Force of the American Congress of Rehabilitation


Medicine has broadly defined three major types of residential facilities: 1)
residential care programs; 2) residential treatment programs; 3) behavioral
rehabilitation programs. These three types of programs are differentiated by
the amount of structure and the intensity of the therapies provided [15].
Residential care programs emphasize supervision and care of the individual;
they provide appropriate social, leisure, physical, and emotional activities, as
well as assistance with tasks of daily living and monitoring of the patient's
health. Although such programs do not provide rehabilitation services, they
stress that patients should frequently be reviewed for rehabilitation potential.
Only those patients who are unable to benefit from, or are unable to obtain,
rehabilitation services should be admitted to residential care programs.
In contrast, residential treatment programs are defined as providing intensive rehabilitation services. These programs should make available physical,
occupational, and speech/language therapy; cognitive retraining; adaptive
daily-life skills and social skills training; psychological services; and vocational
and educational programs. The goal of treatment is to maximize independence
through intensive training, rather than simply to provide short-term or longterm life care. Treatment plans should be aggressive and should specify both
short-term and long-term goals, as well as clear criteria for discharge.
Behavioral rehabilitation programs are highly structured, intensively supervised settings that are meant to accommodate individuals who have problems
of physical or verbal aggression, who withdraw from social interactions or
activities, or have other maladaptive behaviors. Such behaviors are the primary focus of treatment, with the goal of allowing the client to return to a less
restrictive environment.
2.3.2 National Head Injury Foundation

The National Head Injury Foundation (NHIF) issues a directory of headinjury rehabilitation services, in which programs are categorized by the type of
service they provide and the types of clients they admit. The categories are
similar to those used by the Head Injury Task Force, although with somewhat
different labels; in the NHIF directory, the terms acute and long-term rehabilitation both refer primarily to rehabilitation hospitals that provide intensive
physical restorative services after injury.
Long-term rehabilitation programs also overlap with extended intensive rehabilitation programs. These programs are defined as providing therapies for
an extended period of time to seriously injured brain-trauma patients in a struc-

8. Residential Treatment

189

tured setting. Services should include cognitive retraining and speech/language


and occupational therapies; they should focus on activities of daily living,
social skills training, vocational programs, and academic programs.
Transitional living and independent living programs are meant to serve clients
at a more independent level, with emphasis on community interactions and
vocational activities, along with less staff supervision. These situations may be
incorporated into late rehabilitation settings for clients (as they improve and
require less assistance) or may be separate facilities to serve clients who are less
impaired or who have been discharged from rehabilitation programs.
2.4 Choosing the Right Program for Each Patient

The term head injury (more currently referred to as "brain injury") covers a
wide range of severity and disabilities. At one extreme are those patients who
remain in a comatose or persistent vegetative state and require specialized care,
primarily of a medical and physical nature. On the other hand, patients with
mild brain injuries may need only minimal intervention for a short period of
time on an outpatient basis to be able to resume their pretrauma activities.
Individuals who fall between these two extremes may need or may benefit
from residential programs for the kinds of disabilities they demonstrate.
Whenever possible, treatment programs for head-trauma survivors should
be situated in the patients' own communities-not only intensive treatment
programs, but also group homes and other supervised settings where the
individuals will live after completing treatment.
However, programs for brain-injured patients who are unable to control
their aggressive behavior will need to be able to contain those patients to
ensure safety. These behavior-management programs may also work with
amotivational clients to determine what incentives will encourage their participation in daily activities. The focus is on helping patients gain control over
their behavior, with the ultimate goal of discharge to less restrictive treatment
centers (with decreased staff supervision), where they will be able to interact
safely and successfully with the community. Highly specialized behavior management programs deserve a separate discussion and will not be addressed
in depth in this chapter.
Residential programs can best serve their clients if they offer them opportunities for differing degrees of independence and a variety of experiences prior
to discharge. For example, a client might start out in a supervised group-living
situation and later move on to apartment living at the same facility. When
this is not available, clients who are able to manage independent living with
occasional supervision will need to be discharged to an apartment or home
program to gain specific training in this type of life setting.
Brain-injured individuals need actual experience in each type of life setting,
since they usually are unable to generalize training from one setting to another
[16]. Talking about apartment life in a classroom while living in a dormitorytype setting will not prepare a person to actually live in an apartment.

190

II. Rehabilitation Programs

A distinction is made here between traditional residential programs and


apartment programs. There are relatively few apartment programs nationwide, although new ones are being developed each year; they generally serve
brain-trauma patients who either were mildly injured or were severely injured
but have reached a point in rehabilitation at which they can manage an independent, or nearly independent, life setting. They need to provide patients
with the opportunity to live in their own apartments upon admission to the
program or soon after admission, after some initial training. They generally
have a strong vocational-educational orientation and provide intensive training
in daily life skills (with much less emphasis on other therapies). These programs
deserve a separate discussion and will not be addressed in depth in this chapter.
All of the programs that have been discussed-behavior management, residential treatment, and apartment programs-are varieties of postacute rehabilitation. They offer intensive training, in contrast to life-care setttings that
provide safe and, it is hoped, stimulating environments but that do not provide
specific therapies.
This chapter will discuss long-term residential treatment for older adolescents and adults. The length of this kind of treatment for brain-injured
patients varies from 12 to 36 months. Length of stay may be shorter (it might
be determined in six months that a person will not progress and should be
discharged to a life-care setting) or longer (for those less disabled individuals
who can benefit more from treatment and thereby gain greater independence).
Work with brain-injury survivors is similar to, yet sometimes radically
differs from, more traditional forms of residential treatment. Because so little
has been written to date about residential treatment for this population, much
of what will be discussed here has evolved from my practical experience.
3 LITERATURE REVIEW

3.1 Methodological Considerations

Problems that are common to outcome research in mental health are relevant
to research on traumatic brain injury. The primary concern has been that outcome studies cannot both satisfy the conditions of sound experimental methodology and also be applicable to practice [17]. For research to satisfy true
experimental conditions (random assignment of patients, homogeneity of
subjects and disorders, pretreatment and posttreatment assessment), it has to
be conducted in a well-controlled environment. These conditions are required
to rule out alternative explanations of results and to establish connections
between treatment and outcome [18]. However, clinical settings often do not
lend themselves to satisfying these conditions. Therefore, outcome research
must begin to promote methods that can be implemented in clinical settings
while still yielding credible results. This issue is critical, since third-party payers
increasingly require "proof" that treatment is effective.
Two methods that should be considered for future outcome research are the
quasi-experimental approach and the multiple-case-study approach [19]. In the

8. Residential Treatment

191

quasi-experiment, not all conditions are controlled. For example, it may not
be possible to randomly assign patients to treatment groups or to have control
groups. However, given these restraints, it is still possible to evaluate the treatment groups for change. Multiple-case studies, which accumulate results from
a number of cases over time, can also be successfully implemented in a clinical
setting.
No matter what their approach, programs are going to have to assess their
effectiveness, and they must consider all alternatives to select a method that
can be implemented successfully [20]. Beyond those general concerns, research
in the field of neuropsychology requires special considerations. First, and most
important, it must be recognized that brain-behavior relationships are extremely complex. Many different factors may result in, or contribute to, a
particular behavior; no simple one-to-one correspondence should be assumed
[21]. Particular attention must be paid to such variables as age, sex, and
education-all of which, in isolation, affect performance regardless of the
patient's neurological status. Likewise, special attention must be paid to the
pretreatment and posttreatment assessments that have been chosen to document change. Whenever possible, measures should distinguish between patients'
real improvement from a treatment and the improvements seen on tests that
result from the practice gained during multiple testings. Control groups, which
could be used to evaluate the influence of such test-practice effects, are often
hard to organize in a clinical setting. However, case studies that demonstrate
similarity in severity of head injury and personal variables with good pretreatment and posttreatment assessments can be an extremely powerful alternative
to group comparison [22, 23].
3.2 Review of Outcome Research

Little research has been reported to date about the effectiveness of postacute
rehabilitation programs for severely brain-injured people. The literature has
largely focused on the effectiveness of specific interventions such as cognitive
remediation [24, 25]. The following is a briefliterature review of some evaluations that have been conducted to determine patient outcome in both outpatient
and inpatient programs.
3.2.1 Example of an Inpatient Postacute Rehabilitation Program

The Kemsley Unit at St. Andrew's Hospital, Northampton, England, is a


behavior-management program. Patients in the program have had severe
brain injuries, usually three to four years before. The average length of a stay is
12 to 18 months. Tokens as well as social reinforcements are used to teach the
patients socially accepted behaviors. The major goal of the program is to prepare patients to participate in traditional rehabilitation services that are unavailable to them because of their behavior. Typical targets of treatment include
disruptive behavior (verbal and physical), lack of motivation, and cognitive
deficits.

192

II. Rehabilitation Programs

Since the establishment of the unit in 1979, of24 patients who were followed
up (from 6 to 33 months after discharge), 16 (67%) were in environments that
were less restrictive than those they had been in prior to admission. Of those
16, 4 (17%) lived independently, and 12 (50%) lived under family supervision
but without additional help. Gains in activities of daily living were wellmaintained, although improvements in odd behaviors and in drive and motivation were not well-maintained. The majority of the remaining 33% who failed
to benefit had extremely diffuse damage [5].
3.2.2 Examples oj Outpatient Postacute Rehabilitation Programs

Although outpatient programs offer a variety of services, they primarily focus


upon cognitive and social problems of brain-injury survivors. Several postacute rehabilitation programs available to the brain-injured are described below;
all of these programs report some degree of success for their patients, either
compared with a control group or compared with the patients' own level of
functioning before admission. The reports about these programs did not evaluate benefits in terms of costs; however, increasing the independence of these
patients resulted in a decreased need for supervision and staff time. Overall,
then, programs that can help their patients be more independent will improve
their quality oflife while decreasing costs. In addition, these gains in patients'
independence provide tremendous relief to family members.
For instance, the Neuropsychological Rehabilitation Program (NRP) at
Presbyterian Hospital in Oklahoma City provides a structured program for
six hours a day, four days a week. Patients participate in both individual and
group sessions of cognitive retraining and psychotherapy for six months. At
discharge, patients obtained significantly better scores on the Wechsler Adult
Intelligence Scale (W AIS) Performance IQ, W AIS Block Design Scale Score,
and Wechsler Memory Quotient, along with reduced emotional distress compared to a control group. NRP patients also showed more improvement in
personality functioning relative to controls, with lower ratings on helplessness,
degree of social withdrawal, signs of general psychopathology, and restlessness
or hyperactivity [16]. Patients improved in the specific areas targeted by the
training program; however, these improvements did not generalize to other
areas. Only 50% of the NRP patients were gainfully employed (full or part
time), compared with 36% of the control group. It was concluded that specific
training must be provided in order for improvement to be achieved.
Unfortunately, few outpatient programs will accept severely brain-injured
patients. Most of these programs required that patients function at a fairly
high level, be aware of some of their problems, and be motivated to change.
Decreased initiative, motivation, and insight are very common for the braininjured and are among the greatest impediments to treatment [26]. One outpatient program designed to meet the needs of this population is the Adult
Development Center at the Santa Clara Valley Medical Center in California. It
provides brain-injured patients with training in basic academics, self-care,

8. Residential Treatment

193

homemaking, and social skills, as well as physical conditioning, for five hours
a day, four days a week. Participants either have been unsuccessful in other
programs or are too severely injured to enter other programs. Of 95 patients
who took part in the program from 1975 to 1981, 47% attained an improved
level of functioning and were able to move on to benefit from higher level
community agency services [27].
4 STAGES OF TREATMENT
4.1 Assessment and Admission

The primary goal of an initial screening procedure is to ensure that there is


an appropriate match between the brain-injured individual's needs and the
facility's resources. The screening process also gives the therapeutic team an
opportunity to begin to establish a relationship with the family-a relationship
that will be continued throughout the treatment process.
The focus of many residential treatment facilities is on cognitive, social, and
daily-life skills. Although physical therapy, occupational therapy, and speech
and language therapies are available, they should reflect the functional daily
skills that the survivor will need to reenter the community. Therefore, during
the screening procedure, it should be determined whether the individual has
progressed physically to the point of no longer needing intensive therapies to
the extent that he or she initially did, following the injury. Physical,
occupational, and speech/language therapists should have determined that the
individual has reached a point in his or her treatment that requires a setting that
is more community-based, and less intensive.
During assessment, it is also important to ascertain the client's degree of
aggressiveness or passivity. There must be a match between the client's needs
and the program's capabilities to encourage participation and maintain control.
In addition, each residential program needs to determine not only whether it
will treat a given patient, but also what range of problems it will address.
4.2 Evaluation and Treatment Planning

Following a patient's admission, there is generally an initial period of evaluation. For treatment to be most effective, evaluations in each specialty area
involve a two-step process. First, they need to be oriented toward practical
functions-for example, assessing what age-appropriate life skills the person
can or cannot perform. Next, the basic abilities that are essential to performing
these behaviors must be assessed; they will be the "building blocks" that will
be the focus of the patient's treatment.
For example, when a person cannot use money, the evaulation would determine whether this is because he or she cannot remember the value of the coins
or bills, or because he or she has difficulty with basic math skills (e.g., addition, subtraction). For the program to be comprehensive, this evaluation must
be repeated for each area of functioning, and the basic skills necessary for that

194

II. Rehabilitation Programs

area need to be addressed in the treatment plan. Of equal or greater importance


is the assessment of the person's strengths: What are the basic abilities that are
available to use in the treatment program?
The result of this initial assessment period is the development of a comprehensive treatment plan that is clearly individualized, is oriented towards
teaching practical life skills to increase the patient's independence, and is expressed in easily understood terms. The treatment plan should specify what
the patient's problems are, the methods that will be used to address them, and
the staff members who will be responsible for implementing those methods.
The residential program should designate when regular reviews of the treatment plan and patient's progress will be conducted and who will be involved
in those reviews. Even at this stage, there should be some tentative criteria for
discharge. Goals should be clearly stated in terms of attainable life skills.
Even though brain-injured patients have cognitive as well as physical deficits, they must have the major role in their own treatment programs. This
will include a) contributing to, and agreeing with, goals for treatment, b) participating in regular progress reviews and family conferences, and c) being
actively involved in making discharge plans.
The family's needs will be addressed more completely later in this chapter.
However, it must be mentioned that family members need to be consulted
early in the evaluation process about problems they have observed. There
must especially be discussion about expectations for treatment. The staff's
expectations may differ from the family's, and all of these may differ from the
expectations of the individual who is undergoing treatment. These differences
must at least be acknowledged, if not resolved. The staff will be assessing the
patient's and family's levels of emotional acceptance of the brain injury and its
repercussions. For instance, it is often only after several months of treatment
have passed that the survivor and the family members develop a realistic view
about what the survivor can achieve from treatment [28]. Likewise, the staff
may also be either overly optimistic or overly pessimistic, and they may have
to change their views about the patient's outcome as the treatment proceeds.

4.3 Course ofTreatm.ent

Much has been written in the psychiatric literature about the initial phase of
treatment, which is often referred to as the "honeymoon" period [29] because
patients usually maintain tight control over their behavior (i. e., are "on their
best behavior"). This may be less evident among brain-injured patients, because they may be less able to control problem behaviors. However, there may
be a similar "honeymoon" period of optimism about the treatment program.
Brain-injured patients are often unrealistic about their abilities, and they see
themselves as less impaired than others see them [30, 31]. Additionally, they
may have had to wait for admission to a treatment program or may have
worked very hard to qualify for admission. The patients' initial attitude may
be one of near euphoria, and they may have totally unrealistic expectations of

8. Residential Treatment

195

what the program can do. Most common is the patients' belief that the program
will restore their functioning to pretrauma levels-that they will be "normal"
again. The staff will need to communicate very clearly the reasons why they
are in treatment (i.e., what is "wrong") and the goals of treatment. These
will probably need to be written down and repeated during regular program
reVIews.
After assessment and program development, the initial months of treatment
focus on enhancing basic skills in all areas of functioning. For example, a typical weekly schedule for a brain-injured patient in residential treatment will
include physical therapy once or twice a week. This time will be used to maintain physical gains that have already been made and to provide additional training in balance and movement as needed (e.g., a patient may walk well on a
smooth, indoor floor surface but may be unsteady and may require help to
negotiate rough, uneven terrain outdoors). A physical therapist can also establish exercise and mobility programs that patients can follow, on their own and
with other staff members' help, to increase their strength and stamina.
Occupational therapy is usually scheduled at least twice a week and should
focus on daily life skills. Initially, the occupational therapist can help the patient
organize and complete a program of personal hygiene and room care. As each
person progresses, the emphasis should shift to performance of daily life skills
such as cooking, cleaning, and shopping. The occupational therapist actually
enhances performance skills by completing these chores with the patients. In
many cases, physical disabilities make the brain-injured patients slow and
awkward. The occupational therapist who focuses on performance skills will
also have to take the patients' cognitive deficits into account and help the
patients compensate for poor memory, disorganization, disorientationanything that interferes with the completion of daily life skills. An occupational
therapist can provide information about the most efficient ways to accomplish
these tasks so the patient can become more independent.
In speech and language therapy, group treatment is a valuable adjunct to
individual treatment. Individual sessions generally focus on quantity and quality
of speech; patients can often learn to improve or compensate for a variety of
speech problems. Group sessions can focus on social aspects of speech; patients
can engage in conversations that provide experience with different social
situations.
Cognitive remediation begins by helping patients develop, or compensate
for deficiencies in, basic skills such as attention, concentration, memory, and
organization. Although techniques can be used to improve memory, a system
for writing down important events and information is essential [32].
The efficacy of computer programs for cognitive training is still being
scientifically scrutinized. However, because brain-injured patients often have
difficulty with generalization, rehabilitation programs must provide them
with experiences of actual life events that they will encounter. For example,
disorientation is better addressed by acquainting the individual with his or her
current environment than by using pictures or stories about other places.

196

II. Rehabilitation Programs

Judgment and problem-solving will need much attention, since these skills
are diminished for almost all brain-injured people. A step-wise method of
working through a situation can be used with problems that a patient is actually
encountering. The person may also need academic instruction, and the facility
will be required to provide a certain number of hours of instruction each week
if the patient receives school funding. Depending on an individual's ability to
absorb new learning, he or she may be able to benefit from continued academic
work.
Along with the skills training, it is equally important to establish a recreational program for each patient. Most brain-injured people find it extremely
difficult to fill their free time. Severely brain-injured survivors need to have
activities scheduled for them, and supervisory staff must be available to ensure
that they actually participate in the scheduled activities. This means that the
evening and weekend staff members must be as well-trained as the weekday
staff about the effects of brain injuries and must be prepared to carry out a fulltime program during those hours. In particular, the activities must be ones
that the brain-injured patients can perform, given their cognitive and physical
deficits. For example, it should be anticipated that someone with a short attention span will need more staff support for engaging in frequently changing
activities than for watching a movie or television alone for an hour.
Concurrently, brain-injured patients need to be provided with counseling.
The major goals of psychotherapy during residential treatment are the acceptance of the changes brought about by the training and the development of the
skills needed to form relationships. The frequency of individual psychotherapy
sessions should be based on each patient's cognitive status. Meeting once or
twice a week for half an hour with a counselor can be sufficient to provide a
supportive relationship for a patient who is very limited cognitively [16].
Longer sessions can be made available to individuals who are better able to
engage in working through issues of how the injury has affected them and
their lives. For all brain-injured patients, group sessions are effective in providing support, feedback, and training in social skills [24]. Because brain injuries
often diminish patients' self-awareness, videotape replays of these sessions increase each group member's understanding about how he or she appears to
others. In these group sessions, members can confront their problems in a
warm, supportive environment, but even more attention should be given to
highlighting members' strengths.
Aggressive behavior and sexuality are two other issues that need to be addressed in the residential setting. Aggressive behavior can be a result of the
brain injury (e.g., loss of the ability to control impulsive behavior) as well as a
reaction to the brain injury. Patients must be helped to develop some means of
self-expression and self-control. Psychiatric consultation and psychopharmacologic agents are often useful as means to help patients control themselves. Disinhibition can also lead to sexual "acting-out" (i.e., physical involvement,
either heterosexual or homosexual, that is beyond what is accepted by the

8. Residential Treatment

197

community). However, sexuality is an important aspect of a person's life; staff


members need to help patients control their sexual acting-out, but should also
encourage them to engage in age-appropriate sexual behavior. Sex education
and the provision of information about birth control (and birth-control methods)
are important components of this process.
4.4 Encouraging Emotional Acceptance of the Head Injury

After they have been able to master basic skills, patients will move on to experiences with more complex tasks of daily living. These tasks include housekeeping skills, community interactions, and job skills. It is often at this point
in treatment that brain-injured patients begin to realize that certain disabilities
are permanent and that they will not return to their pretrauma existence. This
realization is often accompanied by anger, severe depression, and even the
potential for suicide. At times, brief psychiatric hospitalization may be required
to ensure the patient's safety.
The goal of treatment at this time should be to provide support while the
person grieves for the loss of his or her pretrauma self. For many individuals,
this is the first time they have had to grieve about what they have lost, and
they must be allowed to go through the process rather than be cheered up with
false hopes. The family should be made aware of the need to work through the
grieving process, so that they will not be frightened by the patient's depression
and also so that they can support the process.
One precipitant for the grief process may be the introduction of vocational
activity [33]. Patients begin to understand that they may not be able to return
to their occupation or enter the field that they would have chosen if the injury
had not occurred. Patients may also have the experience of being unable to
perform jobs that were easily accomplished pretrauma [2].
Many brain-injured patients come to realize that they will not be as independent as they had hoped they would be after their treatment has been completed. For instance, many of them will require a group home setting, or at
least a supervised apartment. All staff will be involved in supporting the patients
as they become more realistic and strive to achieve self-esteem in the face of
these disappointments.
4.5 Preparing for Discharge

The final phase of treatment must emphasize giving patients experiences that
will prepare them for the next life setting that they will encounter after discharge. For example, there should be work experience, so that each person can
comfortably assume the same type of work in a new community. In addition,
patients should be allowed to be responsible for as much of their own carepersonal hygiene, room care, cooking, and community interactions-as is
possible.
The staff will need to be actively involved with each patient's family, as
well as with the patient, in making and carrying out discharge plans. Again, as

198

II. Rehabilitation Programs

in finding a residential treatment facility, it will be a matter of matching the


patient's needs with available resources, and the process must take into account
the patient's level of functioning. A group home may be required if the patient
needs on-site supervision, whereas an apartment may be appropriate if the
patient is able to complete most daily tasks independently. Any setting must
have the capacity to provide some services, including support for the individual to engage in as much self-care as possible, as well as a structured work
experience and a structured recreational program. It is often helpful if ongoing
individual counseling and support groups are also made available in the new
setting. In addition, many brain-injured individuals will need to engage in a
program of physical exericse (monitored by a physical therapist) to maintain
their physical gains, strength, stamina, and general mobility.
5 CASE ILLUSTRATIONS

5.1 Kevin: A Successful Outcome

Kevin is representative of those brain-injured patients who are able to use a


residential treatment program effectively. His motor-vehicle accident occurred
during his last year in high school. He was in a coma for six weeks, followed
by a prolonged period of posttraumatic amnesia. His family had enough resources and knew enough about his needs to seek out an acute rehabilitation
program that specialized in head trauma, even though it was some distance
away from their small town. After his discharge from the acute center, he
finished high school and attempted to enter a local community college, but he
was unable to manage the work. His family searched for a program to improve Kevin's thinking skills, establish realistic vocational goals with him, and
develop his independent living skills.
On admission to a residential treatment program, Kevin showed some physical problems, including a right hemiparesis resulting in a slow, awkward gait
and restricted use of his right arm. Because he was right-handed, this interfered significantly with writing and other tasks. There was also a loss of hearing in his right ear and reduced senses of taste and smell. He had short-term
and long-term memory deficits and reduced auditory comprehension and was
unable to organize material by himself. His thinking also tended to be somewhat concrete and rigid, which made it difficult for him to shift ideas and
limited his problem-solving ability.
In spite of the family's efforts to treat Kevin as an adult, it was often easier
for them to do things for him, especially physical tasks. As a result, he was
becoming more and more dependent on them. In addition, all of his high school
friends had gone on to college or were working, leaving him without social
contacts.
Kevin might have been able to manage as a patient in an outpatient program, but there were no services available in his community. Furthermore,
one of his major goals was to become independent of his family; this make a
residential program away from home even more attractive.

8. Residential Treatment

199

A program that would focus on his major goals was designed with Kevin
and his family. Cognitive retraining addressed his memory and organizational
skills, and strategies that he could use were devised. An outline of questions
was prepared for him to use to lead him through situations, from defining the
problem to selecting a solution. In this way, he used the same sequence of steps
every time he encountered a problem.
Even with such assistance, Kevin did not have the cognitive capacity to go
on to college. He was able to take one course at a time for his own enrichment,
but if a course involved a lot of written material and numerous assignments, it
required too much expenditure of effort to make it worthwhile. Vocational
counseling focused on helping Kevin develop a career in an area in which he
had been interested before his injury: medicine and hospitals. He was willing
to investigate different types of hospital aide positions to determine which jobs
would be both challenging and possible, given his limitations.
Kevin tended to be passive, showing little feeling. However, he was extremely angry about his accident and needed help to express that anger. He
also had difficulty being assertive with friends and community contacts. Individual psychotherapy sessions first helped Kevin express how angry he was at
the accident for changing his life. Later, they established that he was still
worthwhile as a human being and that he deserved to be treated well by others.
Group psychotherapy helped Kevin with assertiveness skills. In addition, the
members addressed Kevin's rigid style, which tended to make him judge
others harshly if they disagreed with him.
Kevin's physical deficits prevented him from participating in most sports,
which resulted in his losing strength and stamina. Physical therapy helped him
develop a maintenance exercise program to address these concerns. Occupational and recreational counselors were concerned that Kevin be able to carry
through activities of daily living. He needed to develop specific skills such as
meal planning, cooking, cleaning, and shopping. However, he also needed to
learn how to take his slowness into account; it took him much longer than it
had before his accident to accomplish each task, and if he did not plan on the
extra time, he would not be able to finish all the necessary activities.
Kevin was also encouraged to find activities to improve his social life. He
enrolled in a YMCA and went to a community center to participate in its
programs.
As Kevin progressed, he was placed in an apartment setting for additional
training. It was important for him to learn to take public transportation, plan
his weekly schedule of work and leisure activities, and so on. He also needed
extensive help in arranging his apartment to compensate for his deficits. Written
reminders were placed around the apartment and special places were set aside
for lists of weekly and monthly chores.
When Kevin was discharged, the staff helped move much of this program to
a town near Kevin's parents. Strategies that had been developed for Kevin's
apartment were used in his new apartment. He continued his work as a hospital aide. A community center and YMCA provided him with recreation and

200

II. Rehabilitation Programs

social contacts. Adult night classes at the local high school were also a vehicle
for him to meet others. He and his family made a schedule for telephone calls
and visits so that he would not reestablish dependency upon them. A mental
health worker from the town's social-service department was engaged to come
by twice a week to check on him. Kevin also had a counselor who continued
individual counseling sessions throughout this transition.
This case illustrates how a very functional, goal-oriented residential treatment program can be successfully tailored to meet the unique needs of each
patient and his or her family. For example, Kevin was able to discuss his
problems and was highly motivated to change. His behavior was socially
appropriate, which enabled him to successfully initiate and maintain community contacts (i.e., participate in YMCA and other community activities)
without supervision. His family was aggressive in acquiring good treatment
for him and in openly challenging the staff with any questions they had about
the program. However, the family also was able to acknowledge the deficits
that resulted from Kevin's brain injury and discuss realistic goals for his futurethat is, they accepted him as he was after the accident. Although those deficits
altered his life, he was still able to achieve a relatively independent life and
maintain loving relationships with his family and friends.
5.2 Tom: Need for a Restrictive Environment

Probably the greatest impediment to patients' progress-beyond the issue of


insight and motivation-is the presence of physically aggressive outbursts.
Verbal outbursts are embarrassing for patients and affect their interpersonal
relationships. If these outbursts occur in public, they set the patients apart as
"different." Furthermore, physical aggression that is dangerous to themselves
or others means that the patients cannot remain in an open, community-based
program. Such behavior may require patients to spend some amount of time
in a closed treatment program that focuses on achieving behavioral control
prior to addressing other problems of living.
Tom is an example of a patient who could not be successfully treated in a
community-based residential treatment program. He had been living away
from home for several years before his motor-vehicle accident. He had never
done well academically and had gone to work right after high school graduation; he had taken great pride in his independence from his family. Although
his head injury left him intact physically, it left him with cognitive deficits that
made him unable to work and necessitated his return home to live with his
family.
Tom's major problem was a severe expressive aphasia. He was extremely
angry about the accident and his subsequent problems, which set up a vicious
cycle. He would have trouble talking, would become frustrated (which interfered even more with his speech), would become even more angry, and so
on. His family had tried to involve him in several outpatient programs, but,
although he would begin each one, he found the subject matter in cognitive

8. Residential Treatment

201

retraining programs to be too academically oriented and therefore irrelevant to


his goal of returning to work and living on his own. He preferred a residential
program as a way to be independent of his family and to engage in vocational
planning.
As a result of his outbursts, Tom's relationship with his family had become
very strained. Generally, Tom was verbally aggressive when he was angry,
but he could also be physically aggressive when he was drinking. He refused to
admit that he had a drinking problem and was often able to stay away from
alcohol. However, once he started to drink, he could not stop, and he became
extremely hostile and aggressive.
Tom refused to participate in programs that did not fit into his sense of what
he needed for himself. Therefore, his program needed to be very practical and
oriented toward vocational activities. The staff pointed out ways in which the
program would give him "tools" that would help him live on his own and
return to work. Soon after admission, he was helped to find a job on a construction crew. All therapies were geared towards job-related issues. Speech
therapy helped him to learn strategies to ease his expressive aphasia. Cognitive
therapy helped him develop aids to remember his assigned chores and perform
them in the appropriate order. His paychecks provided an opportunity to help
him make up a budget. Everyday materials such as newspapers and job applications were used to evaluate and improve his basic reading and writing
skills. Occupational therapists worked with Tom so that he would be able to
do basic household and cooking chores. Individual and group psychotherapy
proceeded very slowly, helping Tom to form relationships before beginning
to address his anger.
It seemed as though this program would give Tom the skills he needed to
successfully manage his life. Unfortunately, even with the gains he was making
in other areas, Tom found it difficult to use nonverbal and nonaggressive
strategies to vent his frustration. At best, he was able to leave the situation
and work off his frustration by himself, often through walking or exercise.
However, he entered a period of time that he found very stressful: He was
temporarily laid off from work, and simultaneously was rejected by a woman
he was dating. Tom began to drink and got into fights. He was unable to trust
the program to help him, and he became increasingly defensive, angry, and
threatening.
Tom was sent to a behavior-management program that was able to contain
him in a closed environment. There he was able to discharge his anger at the
accident and learn nonviolent ways to express his frustration about his language
problem in safety.
6 RESIDENTIAL TREATMENT AND THE SURVIVOR'S FAMILY

6.1 Effects of Traumatic Brain Injury on the Survivor's Family

It has been noted repeatedly [10, 12, 34J that the common sequelae of severe
brain injury are extremely difficult for family members to cope with. The

202

II. Rehabilitation Programs

patients' cogmtive and personality changes-particularly poor frustrationtolerance and physical and verbal aggression-are viewed as much more
troublesome than the physical deficits. Patients' apathy and lack of motivation-as well as their poor memory, organization, and problem-solving
skills-also increase their dependency on family members. However, there
are other characteristics of traumatic brain injury that make it stressful for
the family:
1. There is no warning or preparation for this tragic event. The family is told,
often in the middle of the night by the police, that an accident has occurred
and that they must rush to the emergency room. Other family members or
friends may have been involved in the catastrophe as well.
2. Family members may have to live for weeks or months with the uncertainty about whether the injured member will live or die and about how
much function he or she will recover.
3. Head trauma is still a relatively unknown disorder (although the NHIF is
working to spread information about it). In addition, the effects of a brain
injury vary so widely from person to person that it is impossible to predict
how any individual will be affected.
4. The survivor of severe head trauma will move through various stages of
recovery, from coma to semi-independence, and has differing needs for
treatment at each stage. To plan for appropriate treatment, the family must
acquire a great deal of information about programs, professionals, and
therapies. They must also become familiar with different types of financial
aid and must aggressively pursue admission to, and funding for, treatment
programs for the injured person.
5. Although some brain-injury survivors had problems prior to their accidents,
many were normally developed teenagers and young adults who were preparing to establish independent lives, or who had already established them.
Their parents had raised them to the point at which they were increasingly
on their own. After brain trauma, a survivor is thrown back into a position
of extreme dependency on his or her family. All of the family's dreams for
the person's future are also destroyed.
6. Severe brain injuries often bring significant personality changes along with
the physical and cognitive defects. The family must mourn the loss of the
person as he or she was before the trauma and must learn to adjust to these
posttrauma differences.
7. Traumatic brain injury is a lifelong disorder-and therefore a lifelong source
of stress for the family. The family will need to plan for the future care of
the brain-injured person and may face severe financial burdens, in addition
to the emotional stresses described above.
This combination of factors creates special needs for these families, needs
that must be met by the residential treatment facility.

8. Residential Treatment

203

6.2 The Family's Role in Treatment

Before entering residential treatment, most brain-injury survivors have been


through intensive care units, acute hospitalization, and acute rehabilitation
hospitals. They may be ready for discharge from the rehabilitation hospital
but may be unable to return home. Others may have gone home but have
been unable to resume their pretrauma activities. They may have already been
involved in outpatient treatment.
What has been happening to a patient before he or she enters a residential
treatment program often affects how the family feels about the program. For
example, the need for continued treatment may be viewed optimistically,
especially if the treatment follows right after the patient's discharge from the
rehabilitation hospital. Some families may have had to work very hard to prepare the injured member for a residential program (i. e., the patient had to meet
physical goals first). For families in which the injured member has been home
but has been unable to resume his or her former life, the residential program
may be viewed as a "second chance" (beyond the acute rehabilitation program)
for the patient to make a new posttrauma life.
On the other hand, the residential facility may represent a significant separation for family members after months of intense involvement with the patient.
They may also be depressed and pessimistic about the patient's failure to return
to pretrauma levels of functioning. Interactions with personnel at different
facilities may have left them distrustful of professionals. If dire predictions for
the injured person's future prove wrong, the family is less likely to believe any
prognosis about the level that he or she will reach with treatment. The staff
must be aware of each family's unique set of circumstances before admission
to understand the family's position and be ready to accept it and support them.
Although a head injury is amenable to treatment, it does bring permanent
changes. The survivor will not be restored to his or her pretrauma self. Therefore, the family must be involved in the treatment process, not only to learn
what tasks the brain-injured person can perform but also to come to accept the
person's condition. This requires frequent contacts with the patient and with
the staff over the course of treatment to keep abreast of changes and gain understanding about how the patient is different from his or her pretrauma self.
Such contact will also ensure consistency in treatment of the patient, if staff
and family interact with the person in the same way.
The need for family .members to be closely involved in the treatment program provides a particular challenge for residential facilities. Most often, a
residential treatment program is located at some distance from the family's
home. The family can rarely afford to visit more than once or twice a year;
some families may not even be able to afford to make long-distance phone calls
as often as they would like to, or to talk for very long. It is the responsibility
of the program to reach out to the family during each stage of treatment and
followup [34].
During the initial period of observation and assessment, the staff should use

204

II. Rehabilitation Programs

the family as a major resource. Having spent many hours observing and interacting with the patient, family members can provide a great deal of information.
Rather than waiting to make discoveries about the patient by observing him or
her (in a sense, "reinventing the wheel"), staff members should consider the
family as the first source to check with when they have questions about the
patient.
The staff will also need to be in contact with the family members to determine their expectations and goals for the patient's treatment. The treatment
plan should reflect the views of the family as well as the clinical team, and it
should not just be presented to the family without its input. At the same time,
the staff can also be assessing the family's level of acceptance of the patient's
injury, looking especially at the areas of cognition (has the family learned about
the problem?), emotion (has the family begun to accept the problem?), and
behavior (how does the family act toward the patient?) [35]. Understanding
how the family functions in these areas will enable the staff to formulate goals
for the family's development over the course of the patient's stay in the facility.
The treatment plan should be reviewed in a conference with the family.
Enough time should be allocated so that each facet of the program can be discussed in detail. Any discrepancies (between what the family expects and what
the staff has determined that the patient can achieve) must be openly discussed,
so that the family can decide whether to support the treatment or to seek help
elsewhere.
This is an appropriate point at which to mention that it is essential that family
support of the treatment plan be an ongoing clinical goal of residential treatment. This, support will provide a tremendous impetus to the staff. It is also a
means through which the family can express its expectation that the patient
will take an active role in his or her program and not just wait for the program
to make him or her better. This support is also crucial because it implies that
that the family will not "rescue" the patient (i.e., remove him or her from
treatment) whenever treatment becomes too stressful. When the staff and
family are working together, staff members can alert the family whenever they
feel that the patient may be in distress and will need encouragement to stay in
the program.
At the end of the very important initial meeting, there should be a sense of
agreement about long-term treatment goals and an understanding of the shortterm objectives and methods that will be used to reach those goals. In addition,
the staff will need to be very clear with the family about what can be expected
for the patient, in terms of both daily care and progress over the course of
treatment.
During the next few months of treatment, the family will need to know the
patient's schedule so that they know what he or she is doing on a daily basis.
One member of the staff can be designated to make regularly scheduled telephone calls to the family, weekly or every two weeks, to provide the family
with information about the patient's program and progress. This is helpful

8. Residential Treatment

205

because patients often have poor memories and cannot give even routine descriptions of their daily lives in the program. Family members should feel that
they can freely call the facility at any time, but if they are anticipating a scheduled call, they are more likely to hold their questions and concerns until then.
Telephone calls can also be a way for the staff to determine the family's level
of acceptance of the patient's deficits and to follow its progress in understanding, accepting, and behaving in accordance with the changing needs of the
patient. Moreover, the family's ability to follow a preset telephone schedule
indicates how much trust the family has in the facility. For example, when
family members continue to call frequently between scheduled phone calls
(several times daily or almost daily), the staff will infer that there is a lack of
support at home. A family that has received regular updates from staff about
the patient's status and program but that continues to call daily, or even several
times daily, with concerns has not received sufficient reassurances from staff to
be able to relax and believe that the facility is providing good care. Staff members will then have to reevaluate how to approach the family to establish a
communication system that will better serve the family's needs.
Regularly scheduled conferences, held as often as the family can afford to
travel to the facility, give family members an opportunity for contact with
other staff members besides their telephone liaison person and any staff they
may have met at the initial conference. It is probably ideal to have conferences
every three months, but almost all families will be able to visit every six months.
During visits, family members can also observe what the patient is doing and
can get some personal experience with the program. These conferences will
also be a forum in which to discuss how the family is dealing with having a
brain-injured member. If the staff has noticed that the family is having trouble
supporting the program, this must be openly discussed and concerns must be
resolved in order for treatment to continue and for it to be effective.
Home visits are an important component of treatment. During the patient's
visits to the home, the family can observe the progress that has been made as a
result of his or her treatment. The staff will need to help the family structure
the time at home; if necessary, the staff, the family, and the patient can work
together to create a daily schedule to be followed at home. The family and
patient can be given goals to accomplish together during the visit. It is extremely important that the patient be able to make enough visits home so that
he or she remains familiar with the home community. This is particularly
necessary if the patient will return there upon discharge. The staff will also
want to help the family feel comfortable interacting with the patient in the
home, especially as he or she improves and becomes more independent in the
course of treatment. Often this will involve helping family members do less
for the patient, and encouraging them to allow the patient to take care of
himself or herself. The family will also need to be involved in aspects of the
treatment program that address aggressive behavior and sexuality.
The family must take part in planning for the patient's discharge from the

206

II. Rehabilitation Programs

residential treatment program. From the time of admission, there should be a


discharge plan that patient, staff, and family follow and revise together as
necessary. Most often the patient will be returning to the home community.
This makes it vital that the family develop the plan with the staff, be able to
support it, and help implement it.
Contacts with the family should continue after the patient's discharge. By
the time of discharge, the staff has often become a strong support system for
the family and should continue to be involved while the family develops a new
support system. If the patient goes to a new program, the staff can provide
valuable information to ease this transition. Maintaining contact with the family
and the brain-injured person will also enable the residential treatment facility
to conduct followup studies.
6.3 Support for the Family

As noted before, brain-injured people often undergo a period of depression


when they face the permanence of some of their disabilities. Family members
also have many adjustments to make. There is a growing realization [12] that
reaction to the head injury of a family member differs from reaction to a death
[361. It is generally accepted that when a member of a family dies, the family
goes through various stages of adjustment, the final stage of which is acceptance. However, after a person has suffered a head injury, family members
experience these stages over and over again, particularly as the injured person
passes-but is not able to participate in-important life events [37].
Parents of persons who were injured as children or teenagers report that
they have strong emotional reactions to watching their son's or daughter's
friends graduate from high school and college, get married, become parents,
and so on. Older injured adults are often unable to participate effectively in
important aspects of their lives, such as parenting. Many brain-injured people
have been involved in relationships with boyfriends or girlfriends or spouses
before their injuries. These relationships are often devastated by the injury,
and the non-injured people will need support to determine the extent to which
they will remain involved with the brain-injured person. Unfortunately,
although there may be funding for the brain-injured individual's treatment,
there is usually little funding for the family to receive professional help in
dealing with such issues.
On the other hand, the staff members who work with the patient often have
the opportunity to provide support through the strong relationships that they
form with the family as well [38]. Staff members in a residential treatment
facility are also important sources of information. Many families have not yet
developed an understanding of why the brain-injured person is different from
his or her pretrauma self. It is often a great relief for them to know that a given
behavior can be attributed to head injury and is, in fact, a common sequela. It
is the changes in the patient's cognition and personality that are most devastating for family members [10, 39], whereas many of the services provided for

8. Residential Treatment

207

the patient during acute rehabilitation are focused on physical disabilities.


Family members need to understand these cognitive and personality changes.
They also need to alter their expectations for the brain-injured person. It is
especially challenging for the family to accept a reorientation of the patient's
future, for example, in a vocational rather than academic direction. The older
adult may need to change vocations, seek early retirement, or become involved in volunteer activities (as opposed to competitive employment). Recommendations may need to be repeated over numerous calls and conferences
before the family can accept them.
7 RESIDENTIAL TREATMENT STAFF

7.1 Staff Composition

A residential treatment facility for brain-injured patients should have a staff


whose members represent a variety of professions but who all have a common
understanding of traumatic brain injury. The staff should include rehabilitation nurses; physical, occupational, and speech/language therapists; teachers
with a background in special education who are able to provide cognitive retraining as well as academic instruction; vocational and recreational counselors;
and psychologists. Psychiatric and medical personnel should be available for
consultation.
Therapeutic-recreation staff members who are on duty evenings and weekends must be as well-educated about head trauma-and as committed to working with brain-injured patients and implementing programs-as the day staff
members arc. These therapeutic-recreation staff members should also receive
training that will enable them to help patients interact with the community.
They need to know specifically how to give patients feedback-in a discreet,
respectful manner-about their interpersonal interactions when they are out in
the community.
For patients in a residential treatment program, evening and weekend hours
are often just as meaningful as daytime hours, and staff schedules should reflect
that fact. Overlapping the schedules of professional staff and recreational counselors ensures program continuity and consistent interactions with patients.
It also facilitates training and communication among the staff: Therapists
share their professional knowledge, and recreational staff share their unique
perspectives of the patients.
7.2 Major StaffIssues

Staff members in a residential treatment facility are vulnerable to a number of


problems that interfere with effective treatment of patients [40]. Three problems that are particularly troubling are "burnout," anger at patients and their
families, and infighting for power or control. (Section 7.4 also discusses some
of the ways in which these and other staff problems can be alleviated.)
The issue of burnout has received a great deal of attention in the mental
health literature. Head trauma as a disorder is exhausting to treat, and staff are

208

II. Rehabilitation Programs

highly susceptible to burnout. Burnout is a state of exhaustion and diminished


morale among staff, resulting from overwork, overresponsibility, or lack of
support in recognizing and discussing staff members' emotional responses to
specific patients or patient populations. Staff members often react with irritability or anger to demands of the patients, the work setting, or their supervisors. They may feel guilty about their reactions and are often unaware of
their state of exhaustion or excessive stress. An enormous investment of time
and energy may be made, over months of treatment, for very small gains that
may only be apparent in one setting (i. e., the gains do not generalize to other
situations). It is common for staff members to feel that nothing they do is
worthwhile or meaningful. They may continue to go through the motions but
without any enthusiasm for their work.
Staff members sometimes feel a great deal of anger toward their patients or
their families. Although this anger has rarely been addressed, it is a very real
phenomenon in rehabilitation [41]. When anger is present, it colors all interactions among staff and between staff and patients. Anger arises particularly
whenever staff feel powerless to effect change in their patients; as a result, they
feel helpless, frustrated, and incompetent. In other words, the staff members
begin to feel that the patients' progress, or lack of progress, is a reflection of
their efforts and expertise. Brain-injured patients' lack of motivation and their
inability to generalize performance from one area to another are particularly
difficult characteristics for staff to deal with. It is difficult for staff to understand that these behaviors are the result of the anatomical changes in the brain,
rather than the patients' "will" -that is, it is easy for staff to feel that the
patients could behave in a particular way if they "really wanted to."
Tension among staff often becomes manifest in fighting about patients or
other issues of authority and power. Staff members fight about who should
perform which functions with patients. Each discipline becomes increasingly
rigid about what tasks it will or will not be responsible for, which results in
tense, suspicious interactions. Professionals become rivals, instead of working
together as a team to provide the best program for the client. Conflict among
staff is usually a sign that staff members are feeling helpless, overwhelmed,
and unsupported. (Similar feelings can lead to conflicts between staff and
patients, too.) These feelings need to be addressed. Leaders within the therapeutic community who have authority, as well as respect, will need to raise
these issues directly. They will also need to be straightforward in pointing out
problematic interactions among staff members and in addressing the underlying
feelings.

7.3 Staff Training

A major priority for a residential treatment facility must be the education of all
staff members in several specific areas: neuroanatomy, traumatic brain injury,
neuropsychiatric functioning, and different treatment modalities.

8. Residential Treatment

209

7.3.1 Neuroallatomy

The study of neuroanatomy provides staff members with a common base of


communication with each other and with other professionals. Although the
material can be at a fairly simple level, it should include information about the
major areas of the brain and their basic functions, as well as the mechanisms
for transmission ofinformation within the brain. This basic knowledge ofneuroanatomy is a critical building block for understanding what has happened to
the brain-injured individual.
7.3.2 Traumatic Braill Illjury

To understand why the patients are the way they are, staff must learn about
the initial impact and sequence of events that occur in traumatic brain injury.
They should learn to distinguish between problems that occur as a result of
damage to specific regions of the brain and those that result from widespread,
diffuse brain damage. It is extremely important not only that staff recognize
and understand the common problems associated with brain injury, but also,
that they understand the organic nature (i.e., resulting from the brain damage)
of these problems. Many staff who work with brain-injured patients have
worked previously with a psychiatric population. Often they will be predisposed to think that, if motivated, the clients can bring their symptoms under
control. However, given the organic nature of the symptoms, it is most likely
that the patients will be unable to act differently, even though they are highly
motivated. Understanding this dilemma keeps staff from "blaming" the patients
for their problems and from setting unrealistic expectations.
Staff members will also need to be familiar with the course of treatment of
brain injuries. They should know what the patients will be doing and experiencing in residential treatment-as well as what the experience will be like for
the staff. Only when staff from various areas have a common understanding of
residential treatment will they be able to develop a community with shared
knowledge and purpose with regard to the patients. The staff must know that
they can have an impact on the problems associated with head trauma. They,
in turn, will pass this understanding and sense of community on to new staff.
Staff members will also need training that will prepare them to deal with
other problems commonly associated with head injury, for example, patient's
seizure disorders, sleep disorders, and eating disorders. They should also learn
how to care for tracheostomies, if and when a patient with this special condition
is admitted. Most staff members should be certified in cardiopulmonary resuscitation (CPR) every year [8]. The physical problems commonly experienced
by brain-injured people (e. g., seizure disorders, dysphagia) make these safety
precautions particularly necessary.
Even though they are working with clients who do not require a closed facility, staff will need training about how to handle aggressive outbursts. Every
staff member needs to learn techniques of crisis intervention for those situations.
This will require that the stafflearn how to recognize a situation that may lead

210

II. Rehabilitation Programs

to an aggressive outburst, learn how to calm down the people involved whenever possible, and develop a safe way to handle an aggressive outburst if it does
occur. Other brain-injured patients, particularly those who are passive, are
terrified by these outbursts and need reassurance that the staff will be able to
protect them by preventing the aggressive individual from harming himself or herself and others. Every facility will need a plan for safely handling
aggressive outbursts that takes into account the individual patient's needs
and capabilities. Such a plan might involve developing a behavior-management program to help the individual control his or her behavior; neuropsychiatric consultation for medication; and identification of short-term hospital
placement.
7.3.3 NClIrnpsyciziatric FlIllctioning and Treatl/le/lt Modalities

The third major focus of training for staff is in the area of psychiatric diagnoses
and treatment. Brain-injured patients often present neuropsychiatric symptoms, including paranoia, depression, mania, anxiety, delusions, and hallucinations [9]. Staff need to be able to recognize these symptoms and implement
any recommended neuropsychiatric procedures that might be useful in treating the symptoms. Staff should be familiar with neuroleptic and psychotropic
medications and their side effects, so that they can observe any problems that
the patients are having as a result of medication and report those problems to
the neuropsychiatrist.
7.4 Staff Support

Working in residential treatment programs is extremely demanding, because


of the severity of the patients' problems, the length of the treatment, and the
slow gains that the patients make toward reaching their goals. Staff members
can rapidly become exhausted. For this reason, staff support is a particularly
critical issue. Certainly, education is one important method for supporting
staff and alleviating burnout. Education alleviates feelings of being helpless or
overwhelmed by providing staff members with tools with which to handle
difficult situations. It also provides time for staff members to interact with
each other, as well as being a respite from patient care. Educational activities
can be carried out at the facility through a regular in-service program, and staff
members should be given time off from their regular responsibilities to take
part in such programs. In addition, all staff should have financial support from
the facility to attend outside workshops and conferences. These provide staff
with up-to-date information as well as a respite from their usual daily activities.
Frequent educational programs also are good for staff morale because they create
a feeling of being at the forefront of knowledge in the field.
In residential treatment programs, many staff members work with each
patient. All of these people must have an organized method of communicating
regularly about the patient (e. g., through regular staff meetings, patient conferences, etc.) to ensure consistency in treatment and to share important observations and information. Unless they have effective ways of communicating,

8. Residential Treatment

211

staff members will feel that the patients' treatment is fragmented and haphazard. They will also need a forum in which to discuss their different professional
perspectives about the patients. This sharing of backgrounds will enable staff
members to coordinate their activities effectively when treating patients. It
will also encourage staff to respect each other's unique contributions to the
treatment program-instead of competing over the patients.
Finally, it is most essential that staff have a way of gaining emotional support
at work [29, 42]. To facilitate this, the clinical director (or some other person
in a position of responsibility and respect) should encourage open discussion
among the staff. In these discussions, staff can share the positive~ caring feelings
that they have for each other and the patients, as well as their frustrations about
treating such difficult patients. Staff need to be able to acknowledge the patients'
handicaps, rather than be frightened by them, in order to work with them.
Overall, staff must embrace a common philosophical viewpoint: Believe in the
patients; accept them as they are; respect their fight to live a more independent
life. It is only with this respect for the brain-injured patients and their struggle
that the staff will be able to help them achieve age-appropriate skills and the
degree of independence that their disabilities allow, rather than infantilizing
them and impeding their progress.
8 VOCATIONAL PROGRAMS IN RESIDENTIAL TREATMENT

8.1 Need for Specialized Programs

Vocational training for brain-injured patients requires a unique approach for


several reasons. First, there can be a great deal of resistance to vocational programs. For many patients, prospects for work are greatly diminished after their
injury. Often they are not capable of the same kind of training or employment
that they were before the trauma. However, it is very difficult for them to give
up these pretrauma expectations. The resulting denial that develops around
vocational interests is frustrating for staff who are working to help the braininjured patients set up "realistic" work experiences. Patients' rejection of those
realistic goals and refusal to participate in job trials while holding on to higher
expectations is a major impediment to progress.
Second, many programs target specific skills in areas of cognitive and social
functioning. However, as has been noted, brain-injured patients have difficulty
generalizing what they learn in one area to another area, even when the situations
are very similar. Working on problem-solving skills in a cognitive-retraining
classroom will not prepare the patient to solve problems in a work setting.
Prigatano et al. [16] found that, in the absence of specialized vocational programs to target their work skills, only slightly more of their treatment group
(50%) were employed or engaged in realistic training programs than were
members of a control group (36%).
Finally, the types of vocational experiences that have been made available to
brain-injured people are often inappropriate for this population. For example,
sheltered workshops can be helpful in providing supervision and work ex-

212

II. Rehabilitation Programs

perience with little or no pressure to produce, but they generally mix braininjured people with other mental-health populations, which can create a lot of
tension. Traditionally, staff in these workshops do not have the background
to understand and address problems that are specific to the brain-injured.
For all these reasons, ideally any program that provides postacute rehabilitation services to the brain-injured must design separate vocational programs
that will specifically address the special needs of those patients, instead of
attempting to incorporate the brain-injured patients into programs geared
toward other patient groups.
8.2 Vocational Assessment and Counseling

In the past, initial evaluations of patients' vocational interests and abilities have
been somewhat limited by the available measures, which tend to be inappropriate for this population. Often the tests are timed, and brain-injured patients,
with their slow motor-response time and poor information processing, are
unable to demonstrate their true level of functioning. Interest tests are often
better at indicating patients' denial of their deficits than they are at identifying
realistic interests. That is, patients often answer with their pretrauma interests
but do not take into account their posttrauma strengths and limits. The vocational counselor will need to be creative in using traditional measures and
finding new ways of determining these patients' interests and abilities.
Pretrauma interests will continue to be very important for almost all patients.
The goal of vocational counseling will be to apply these interests in a job setting that realistically matches the patients' posttrauma abilities. In some cases,
patients may need to be steered in a completely different direction. This can be
an extremely difficult task if the patients are set on a particular vocation. In
fact, patients may need to experience failure in the desired vocation before
allowing themselves to be redirected [43].
8.3 Graduated Work Trials

Initial retraining in cognitive and social skills will undoubtedly address areas
vital to a vocational program. Improvement in such basic areas as memory,
frustration tolerance, and attention and concentration is necessary before patients
are capable of participating in a work trial. However, training in such higher
level functions as problem solving, organization, interpersonal interactions,
and the like may be irrelevant unless those skills are practiced at the work site
itself. This is where lack of generalization becomes obvious. For example, no
matter how many paper-and-pencil problems dealing with interpersonal conflict they solve in the classroom, brain-injured patients will still need a great
deal of help and support to handle such problems at the work site.
There has been a growing realization that job trials are cost-effective and
beneficial for preparing any disabled population for work [44]. This is especially
relevant for brain-injured patients.

8. Residential Treatment

213

As brain-injured persons progress through the work program, they should


receive different levels of support and supervision. They should begin with a
position in which they have very little pressure. For example, the first work
trial can be a volunteer position within the treatment program itself or in an
equally sympathetic setting. Again, the type of work selected should be related
as much as possible to the patients' ultimate vocational goals. It is likely that
they will have to go through a period of adjustment with each new job, even
though that is a time-consuming and inefficient process.
During the first work trial, staff should concentrate on determining whether
thejob is suited to a patient's strengths [32]. On-site supervisors can determine
which are the best aids for the patient to use in order to make the job run
smoothly. For example, printed checklists of tasks can serve as memory aids;
pictures or tape recordings can be used in place of written material. These and
similar aids that are developed for the current job can also be used in future
positions. Experimenting at this stage with changes in the work environment
itself will also save unnecessary expenditures of time and money later on [44].
Staff who are familiar with the patients and their needs are best able to help
with these efforts; in conjunction with the patients, they can determine which
changes may have improved job performance and which should therefore be
maintained as the patients move through the job trials.
After successfully completing the work trial, a patient will move from this
highly supervised and supportive environment to a more competitive position.
This transition will be most successful if a staff member is able to accompany
the patient. In this way, the staff member and patient, together with the job
supervisor, can plan strategies for handling assignments. After working along
with the patient (and actually experiencing the job), the staff member can
suggest stress-management techniques that will help the patient tolerate the
frustrating aspects of the job. Employers may also be more willing to hire
brain-injured people if they know that they will not have to allocate their own
staff to teach the job and supervise performance.
Staff supervision at the work site should be removed very gradually. The
first step in this process may be to move from total supervision to staff assistance only at the beginning and end of each shift. Later, a staff member can
move to once-daily checks and then to weekly checks. The brain-injured
person may also be able to increase time at work-from a few hours a week
to half-time or even full-time, depending upon his or her fatigue level. It is
often better for brain-injured patients to work part-time rather than full-time,
both because they may continue to tire easily and because they may need
more time to accomplish tasks of daily living.
These preliminary stages in a vocational program should occur while the
brain-injured patients are still in residential treatment. At that point, other
therapists will still be available to help the vocational counselor plan the best
strategies for use at work.
When they are making plans for patients' discharge, staff members should

214

II. Rehabilitation Programs

attempt to find jobs for the patients that are as similar as possible to the worktrial jobs. Contacts should be made with job supervisors so that working conditions can meet the patients' needs wherever possible. Even so, it should be
assumed that this is an entirely new situation for the patients and that they will
encounter many of the same problems that had already been resolved in thejob
trials. It is hoped that these problems will be resolved more quickly at this stage,
but they should still be anticipated. It will be necessary for someone to accompany patients to their new jobs and provide support and supervision. This
supervision can gradually be removed as the patients become more comfortable. They may want to start working only a few hours a week, particularly
since there will be many other changes in their lives after discharge. Work
hours can be increased as other life areas become manageable.
8.4 Sheltered Workshops

Some severly brain-injured patients will be unable to participate in the types of


vocational programs described above and will require a sheltered-workshop
setting. However, as was noted earlier, it is difficult-and sometimes impossible-to incorporate these patients into workshops that have been set up for
other patient populations. However, if their special needs can be met, braininjured patients will benefit from the opportunity to engage in productive
work.
Workshop staff members often find that patients' cognitive deficits, such as
impaired memory and short attention span, are particularly challenging. They
must understand that these (and other) deficits are specific to brain injury and
must be able to help the patients compensate for them. They must also be prepared to deal with any verbal and physical outbursts in a calming, consistent
way. The types of work that are available at the workshop must be appropriate
for varying degrees of fine motor coordination and other physical limitations.
9 COMMUNITY REINTEGRATION

9.1 Activities of Daily Living (ADL)

Depending upon their age at the time of injury, some patients may never have
developed independent life skills. They may also have a great deal of resistance
to learning housekeeping skills. For example, male patients who are striving to
assert their masculinity in spite of their feelings of being physically damaged
can be particularly averse to doing anything that is typically viewed as
"women's work." Head trauma patients-both men and women-who have
been cared for continuously since their accidents also resist these tasks.
However, it is not helpful to these patients, in the long run, to do any work
for them that they can do for themselves.
There are a tremendous number of steps to master in each area ofhousework,
and it will take time for patients to develop and practice strategies. For example,
"cooking" includes all of the following m~or steps, anyone of which can be
overwhelming: making a menu of balanced meals, finding recipes (if needed)

8. Residential Treatment

215

for each menu item, making a shopping list, shopping, storing food appropriately, planning for the time required for meal preparation, actual meal
preparation, cleanup, and storing leftovers.
An ADL program should begin at the time of the patient's admission. Tasks
in the two major areas-housecleaning and cooking-should be ordered in a
hierarchy of difficulty and complexity. On admission, each patient should start
with the first item on the hierarchy and move up as each step is mastered.
Whether they are organized into teams to accomplish each set of chores or are
given individual responsibilities, all patients should be involved in the care of
their house and the preparation of meals.
The actual daily work can be supervised by the recreational staff. However,
occupational therapists will need to be involved with each patient as he or she
works on chores. There may be a need for physical deviees or other strategies
to make a job safer or less time-consuming. Most brain-injured people will
also need to allocate more time for each chore, if they have slower informationprocessing or motor planning. This ean be a source of frustration that can best
be dealt with in a supportive environment.
9.2 Social Skills Training

Social skills must be a focus of any treatment program. These skills, which
are central to a person's quality of life, are often impaired as a result of head
trauma. For example, even simple housekeeping chores like shopping cannot be performed without some interaction with the community. However,
appropriate behavior at the grocery store is only one of many situations the
brain-injured person must be able to manage.
Major impairments of social functioning that are associated with head trauma
include loss of learned responses (social knowledge), insensitivity to subtle
verbal or nonverbal social cues, and the inability to take another point of view.
These deficits, often complicated by impulsivity or passivity, interfere with
patients' ability to function at an appropriate age level. Training patients in
social skills can be very difficult, because the skills to be taught are both abstract
and subtle. Social interactions must be broken down into discrete (and concrete) steps and tangible skills before they can be taught, and this is a challenging proposition for staff. Such details as facial expressions, tone infleetion, and
figures of speech need to be addressed.
Only one problem should be the focus at anyone time. It may help to begin
with the most obvious problem first, since there is more of a chance that a
patient will be aware of that problem, and also that interactions will improve
significantly if it is resolved.
Work on social skills can begin in a group setting. Group members need to
be very straightforward about the problems they notice in each other. Videotapes can also be used to present problems as well as to confront denial. In
addition, videotapes can be used to show patients how they present themselves
to others (which also serves to confront their denial of specific behaviors).

216

II. Rehabilitation Programs

However, social skills trammg should not be restricted to formal sessions;


social objectives should be worked into all other areas of the program, including
the residence. Because all social interactions involve communication, speech
therapists must be involved in the training program, both in pragmatic communication groups and individual sessions. They can work within the groups
to identify problems and formulate solutions and can then work with patients
at a more intensive level in individual sessions. Special attention to interactions
in the community must be a priority. It is essential that patients be provided
with immediate feedback in the community, if they are to develop appropriate
behaviors.
9.3 Leisure Activities

The ability to occupy their free time with activities that are enriching as well as
pleasant is another significant aspect of people's lives. In fact, even making
decisions about leisure activities is a very creative process. People must imagine
different pursuits, judge their ability to engage in them (physically, mentally,
and financially), and be able to get to them at the right time with the appropriate equipment. Brain-injured individuals may find this process difficult
from the very beginning, if they are unable to spontaneously think of a list of
activities. If they do think of an activity, it may be something that they enjoyed
pretrauma but are no longer able to do, or they may have one idea for an
activity (say, at the YMCA) but have no suggestions for filling their free time
at home. Transportation can also be a problem. Calling to find out dates,
times, and costs of activities confuses them, or they may forget to write down
the information when they do call.
All of these planning and problem-solving skills can be emphasized as staff
members work with the patients to set up a full recreational program. The
program should include some exercise and some sedentary pursuits, and group
activities as well as individual activities. These activities should be planned to
encompass time at home as well as trips into the community. A recreational
program will need to take into account the patients' attention span and should
also ensure that enough things have been planned to fill the available time.
Rather than being a reward for patients' participation in therapies, leisure
activities constitute a program in their own right, one that is of equal
importance to other therapies. Once developed, a viable program can be
implemented in other communities after the patients have been discharged.
10 CONCLUSIONS AND FUTURE DIRECTIONS

This chapter has outlined some of the major issues that are involved in serving
brain-injured patients within a residential treatment facility. Although this
topic has not yet received a great amount of attention in the professional literature, the available information does appear to indicate that residential treatment for severely brain-injured patients increases their level of independence
and improves their functioning.

8. Residential Treatment

217

It is vital that programs be practical and oriented towards teaching the


patients age-appropriate life skills. The treatment plan should focus on only a
few major goals at a time. It should also be directed towards finding the best
possible alternate routes to the patient's goals, by circumventing their deficits
and utilizing their skills. The key to promoting functional adaptation is to
focus on the patients' strengths. Skills training should occur in situations
similar to those in which the skills will be used-with continuous and
immediate feedback about performance.
Brain-injured people should, as much as possible, have a primary role in
determining their goals and charting their progress. Both the treatment plan
and the discharge/aftercare plans must be formulated and supported, not only
by the staff, but by the patients and their families as well.
Residential treatment facilities need to establish regular programs for communication with their patients' families. Through these programs, the families
can learn not only about their family members' programs and progress in
particular, but also about head injury in general (information that may help
them temper their expectations). Such contact will also enable the staff to
assess and encourage the families' acceptance of the brain injuries and their
support for the program.
Education and training for staff should discuss the areas of basic neuroanatomy, traumatic brain injury (with its related medical and behavior problems),
and neuropsychiatric complications and treatment. In addition, communication
among staff members can be an effective vehicle to share professional backgrounds, observations about the patients, and support for one another.
In the future, residential treatment programs must continue to address
practical treatment issues, as well as provide opportunities for research. Standards for care and treatment will continue to be developed and revised. It should
be expected that programs will adopt these standards and demonstrate compliance.
Treatment for brain-injured patients will continue to rely on an interdisciplinary team to plan and implement programs. As more brain-injured people
are discharged from intensive programs, greater emphasis will need to be
placed on follow-up. The follow-up process must assess how well patients are
maintaining the skills that they have learned and the degree to which they have
been reintegrated into the community.
So little research has been done in the area of residential treatment that the
field is wide open. In the future, evaluations of the relative efficacy of different
types of programs for different types of problems must be conducted. Within
programs, staff members from each discipline should be taught-and should
use-the most effective methods available.
The most important facet of treatment, however, is that staff members,
families, and patients alike be committed to a philosophy in which patients are
accepted (whatever their problems) and are helped to use their strengths to
become as independent as possible.

218

II. Rehabilitation Programs

REFERENCES
1. Jennett, B. (1983). Scale and scope of the problem. In Rehabilitatiol1 of Head Injured Adult,
Rosenthal, M., Griffith, E., Bond M. and Miller,]. eds., F.A. Daivs, Philadelphia; pp. 3-8.
2. Weddell, R., Oddy, M. and Jenkins, D. (1980). Social adjustment after rehabilitation: A two
year follow-up of patients with severe head injury. Psycho!. Med. 10, 257-263.
3. Johnston, M. and Keith, R. (1983). Cost-benefits of medical rehabilitation: Review and
critiques. Arch. Phys. Med. Rehabi!. 64, 147-154.
4. Cope, D. and Hall, K. (1982). Head injury rehabilitation: Benefit of early intervention. Arch.
Phys. Med. Rehabi!. 63, 433-437.
5. Gloag, D. (1985). Rehabilitation after head injury. Br. Med.]. 290, 913-916.
6. Miller, E. (1984). Recovery and Mallagelllent of Neuropsychological Impairmellts. John Wiley and
Sons, Inc., New York.
7. Luria, A. (1963). Recovery of FUllctioll After Braill ifljury. Macmillan, New York.
S. Jennett, B. and Teasdale, G. (1981). Mallagemellt of Head IlIjuries. F.A. Davis, Philadelphia.
9. Leftoff, S. (1983). Psychopathology in the light of brain injury: A case study. J. Clin.
Neuropsycho!. 5, 51-63.
10. Lezak, M. (1978). Living with the characterologically altered brain injured patient.]. Clin.
Psychiatry 39, 592-598.
11. Rafferty, F., Hawley, L., Citron, c., Ducker, C. and Berry, V. (1982). Tertiary care of the
post-head trauma patient. Psychiatr. Hospital 15, 1893-197.
12. Rosenthal, M. and Muir, C. (1983). Methods of family intervention. In Rehabilitatiol1 of the
Head Injured Adult, Rosenthal, M., Griffith, E. Bond M. and Miller,]. eds., F.A. Davis,
Philadelphia, pp. 407-420.
13. Lynch, W. (1984). A rehabilitation program for brain injured adults. In Behavioral AsseSSlllelit
a/1d Rehabilitatiol1 of the Traumatically Bmin Dall1a,~ed, Edelstein, B. and Couture, E. eds.,
Plenum Press, New York, pp. 273-312.
14. Wilmot, c., Cope, D., Hall, K. and Beckin, C. (1982). Head Injury Rehabilitation Research
Project Final Report, Vo!' I, Santa Clara Valley Medical Center, San Jose, CA.
15. Task Force on Head Injury, American Congress of Rehabilitation Medicine. (1986). Report,
April 4.
16. Prigatano, G., Fordyce, D.]., Zeiner, H.K., Roueche, J.R., Pepping, M. and Wood, B.C.
(1985). Neuropsycholo,~ical Rehabilitation Afta Bmill Injury, The Johns Hopkins University
Press, Baltimore.
17. Cohen, L., Sargent, M. and Sechrest, L. (1986). Use of psychotherapy research by professional psychologists. Am. Psycho!. 41, 198-206.
18. Kazdin, A.E. (1980). Research Desigll in Clinical Psychology, Harper & Row Publishers, New
York.
19. Campbell, D. and Stanley,]. (1963). Experimelltal and Quasi-Experimental Desig11Sfor Research,
Rand McNally College Publishing Co., Chicago.
20. Sargent, M. and Cohen, L. (1983). Influence of psychotherapy research on clinical practice:
An experimental survey.]' Consult. Clin. Psycho!. 51. 718-720.
21. Parsons, O. and Prigatano, G. (1978) Methodological considerations in clinical neuropsychological research. ]. Consult. Clin. Psycho!. 46, 608-619.
22. Barlow, D.H., Hayes, S.c. and Nelson, R.O. (1984). The Scientific Practitioner, Pergamon
Press, New York.
23. Barlow, D. (1981). The relation of clinical research to clinical practice: Current issues, new
directions.]. Consult. Clin. Psycho!' 49, 147-155.
.
24. Diller, L. and Gordon, W. (1981). Interventions for cognitive deficits in brain injured adults.
]. Consult. Clin. Psycho!. 49, 822-834.
25. Lezak, M. (1979). Recovery of memory and learning functions following traumatic brain
injury. Cortex 15, 63-72.
26. Levin, H., Benton, A. and Grossman, R. (1982). Neurobehavioml Consequel1ces of Closed Head
Injury. Oxford University Press, New York.
27. Cole, J., Cope, D. and Cervelli, L. (1985). Rehabilitation of the severely brain injured patient.
Arch. Phys. Med. Rehabi!. 66, 38-40.
28. McLaughlin, A. and Schaffer, V. (1985). Rehabilitate or remold? Family involvement in head
trauma recovery. Cognitive Rehabi!. Gan/Feb), 14-17.
29. Hamilton, J., ed. (1985). Psychiatric Peer Review: Prelude atld Promise. American Psychiatric
Press, Washington, DC.

8. Residential Treatment

219

30. Bear, D. (1983). Hemispheric specialization and the neurology of emotion. Arch. Neurol.
40,195-202.
31. Tyennan, A. & Humphrey, M. (1984). Changes in self-concept following severe head injury.
Int.J. Rehabil. Res. 7, 11-23.
32. Long, c., Gouvier, W. and Cole, J. (1984). A model of recovery for the total rehabilitation of
individuals with head trauma. J. Rehabil. (Jan/Feb/Mar), 39-45.
33. Smith, R. (1983). Prevocational programming in the rehabilitation of the head injured patient.
Phys. Ther. 63, 2026-2029.
34. Sargent, M. and Littman, S. (1985). Development of a family program for the long-term
residential treatment of head trauma patients. Paper presented at the Fourth World Congress
of Biological Psychiatry, Philadelphia.
35. Sargent, J. (1983). The sick child: Family complications. J. Dev. Behav. Pediatr. 9, 50.
36. Popper, A. (1984). The profoundly injured child. Trial Ouly), 28-32.
37. Muir, C. R. and Haffey, W.J. (1984). Psychological and neuropsychological interventions in
the mobile mourning process. In Bevhavioral Assessmel1t al1d Rehabilitatiol1 of the Traumatically
Brain Damaged, Edelstein, B.A. and Couture, E.T., cds., Plenum Press, New York.
pp. 247-272.
38. Sachs, P. (1985). Beyond support: Traumatic head injury as a growth experience for families.
Rehabil. Nursing (Jan/Feb), 21-23.
39. Oddy, M., Humphrey, M. and Uttley, D. (1978). Stresses upon the relatives of head injured
patients. Br. J. Psychiatr. 133, 507-513.
40. Emener, W. G., Jr. (1979). Professional burnout: Rehabilitation's hidden handicap. J. Rehabil.
45 (Jan./Feb./Mar.), 55-58.
41. Gans, J. (1983). Hate in the rehabilitation setting. Arch. Phys. Med. Rehabil. 64, 176-179.
42. Rossman, P. (1979). A model for staff training in the psychiatric hospital treatment of
adolescents. Am. Acad. Child Psychiatr. 18, 559-570.
43. Deaton, A. (1986). Denial in the aftermath of traumatic head injury: Its manifestations,
measurement and treatment. Rehabil. Psychol. 31, 231-240.
44. Bond, G. & Dincin, J. (1986). Accelerating entry into transitional employment in a psychosocial rehabilitation agency. Rehabil. Psychol. 31, 143-146.
45. Hart, T., Hayden, M. and Helfenstein, D. (1984). Vocational components of project re-entry.
Paper presented at the Fifth Annual Traumatic Head Injury Conference, Braintree, MA.

9. CONCEPTS IN DAY PROGRAMMING

NATHANIEL H. MAYER AND DANIELJ. KEATING

1 WHAT IS DAY PROGRAMMING?

A day program is a comprehensive set of therapeutic interventions provided


to a brain-injury survivor who resides in a permanent (i.e., nontransitional)
community setting-at home with his or her family, in an apartment, or in a
group living arrangement. A day program has benefits for the survivor's
caregivers (e.g., the family) as well: By providing additional resources to meet
some of the survivor's needs, it reduces his or her dependence on the family.
Typical goals of a day program center on stabilizing the survivor's living
arrangements and developing and implementing a complete schedule of daily
activities or an activity pattern of living. Day programs arc community-based
and hence they support survivors and their caretakers in the local environment.
Day programs are therapeutic, not merely for "maintenance" of the patient's
current status. They provide comprehensive therapies that are managed by an
integrated team process, as opposed to single or multiple services that are
delivered in a coordinated but weakly programmatic-or even nonprogrammatic-way.
"Home base" for a day program may be a hospital facility or some other
community setting. We make a distinction between a day program and a
day hospital program. In our view, a day hospital program implies that patients
retain inpatient hospital status because they still require medical and nursing
monitoring on a daily basis-or at least Monday through Friday. The day
hospital program may be characterized as a "medical-model" program. On
the other hand, a day program is not necessarily characterized as a medical221

222

II. Rehabilitation Programs

model rehabilitation program. In fact, our impression has been that most day
programs are organized as educational, vocational, or social models whose
fundamental approaches are very different from a medical rehabilitation model.
A day program may be distinguished from a transitional-living program by
the transitional setting of the latter. Both day programs and transitional-living
programs typically engage in community and destination skills training. In
theory, transitional-living training is designed to prepare brain-injured people
to return to their own community destinations. In practice, however, many of
the adaptations that brain-injury survivors learn to make during their training
in the transitional-living setting do not apply to the community setting to
which they return. For example, a person could learn to take the subway in
Philadelphia, but this will not guarantee that he or she will be able to use bus
transportation back in a hometown that does not have a subway. A more
egregious example was seen by one of the authors who visited a facility in
a remote rural setting. He overheard a client who was working in a barn
exclaim, ''I'm a city kid from Philadelphia. What am I doing shoveling manure
[sic1 in the country?" When patients return to their permanent destinations,
new bridging services may be necessary. Many such patients could benefit
from a community-based day program that would provide the bridging
adaptations they and their families, or other caregivers, need to effectively
adapt and become a part of their local community. It would be unusual for a
community-based day program to accept a client whose permanent destination is located in a distant location unless the day program had some mechanism
for providing key destination and community training in that client's home
community.
2 ADVANTAGES OF DAY PROGRAMMING

The old rehabilitation maxim, "Good families make good rehabilitation,"


seems true for brain injury as well. Day programming is of particular benefit
to brain-injured individuals who are living at home with a good, supportive
family. Furthermore, day programming is particularly effective for patients
who live in their own community and have access to resources of a comprehensive brain-injury program.
Two major objectives for such patients include stabilization of their community destination and stabilization of their activity pattern of living. By
"community destination," we refer to a family residence or residential setting
that has permanent, identified caregivers and managers and is located in a
mainstream community. These caregivers and managers may be family members, legal guardians, or residential supervisors. "Activity pattern of living"
refers to a stable set of activities and responsibilities that a brain-injured person
goes through in his or her daily, weekly, and monthly schedule. When roots
and potential community support are available, a day program is in a good
position to mobilize these resources on behalf of brain-injured people and can
produce a stable plan of living for them for many years to come. Because the

9. Concepts in Day Programming

223

day program is present in a person's community, personnel from the program


are available, even after the client's discharge, to handle episodic crises and
provide "booster" support over the years.
There are, of course, other options to day programming. For example,
many facilities in the United States offer postacute rehabilitation at locations
far distant from the brain-injured person's home community. Because many
people with brain injury come from small communities in which there are no
comprehensive services, these geographically distant facilities do provide a
needed service. However, in our opinion, a major drawback for such facilities
is that they are not located in, and therefore do not know, the communities of
origin of the brain-injured clients they serve. As a result, transition back to
home communities is often problematic. In fact, some geographically distant
facilities have set up permanent residences-complete with work options-for
their clients with brain injury, because they are unable to integrate the patients
back into their own communities.
Perhaps services provided in a geographically distant facility may be a
better option for clients who eame from small communities that have few
resources. In addition, some clients served by these facilities may have no
permanent, identifiable destination to return to after acute hospitalization or
inpatient rehabilitation has ended. However, in larger communities in which
there is a sufficient number of clients with brain injury, day programs that serve
the needs of clients in those specific communities would seem to be a more
reasonable alternative. To be truly effective, the day-programming unit needs
to coordinate its efforts with the natural supports that are available to people
in their own communities. Community-based retraining, coupled with a
supportive family, makes an adaptive outcome more likely for the client with
severe brain injury.
3 WHEN TO ENTER A DAY PROGRAM

In our experience, clients who have been at home and in the communicy for a
variable period of time are better candidates for a day program than are clients
who are referred directly from an acute rehabilitation program. A period of
time within the community allows patients and families to "settle in" with the
effects that brain injury has on life's daily routines. Within the context of the
home or other community-living settings, clients' behaviors and deficits become more apparent to their families and caregivers. This enables the families
or caregivers-as well as the clients-to formulate their own ideas about what
they hope can be accomplished as a result of the day program.
Patients who come from an acute rehabilitation setting without an intervening period of being at home or in the community often have unrealistic
expectations about future recovery. Families tend to perpetuate these unrealistic expectations: Since the inpatient hospitalization has not produced the
"100% cure" that they expected, their hopes are that the comprehensive dayprogramming unit will somehow produce it. Clients and families who have

224

II. Rehabilitation Programs

had a period of time to live together posttrauma can more easily identify, and
can more easily accept, particular goals that they think will be a real contribution to their family system. They may defer their expectations for recovery to
"normal" in favor of adaptive, practical gains.
Although we would like to encourage application to our day program
within three months of discharge from an acute-care rehabilitation facility, we
have accepted clients five or more years after they had been injured. We have
also accepted many clients who have been told that the only thing they had to
do was to go home and wait for recovery. These people, typically defined in
the literature as suffering from "minor head injury," often have been told by
their family physicians and others that, over time, they would improve. When
they did not improve, and became maladaptive in everyday living activities,
they were eventually referred to our day program, and we were able to help
them formulate plans for the future.
4 EVALUATION AND TREATMENT PLANNING

After patients enter our day program, we not only analyze their actIvIty
patterns but also evaluate the stability of their living arrangements. Even
though a patient may have been discharged from an acute rehabilitation center
to his or her family home, the stability of the living arrangements that were
initially constructed must still be tested over time. Many families accept
patients back into their homes, not because they have been well-prepared by
the rehabilitation staff of the inpatient unit, but rather because they expect
change and are willing to wait for this change to occur. When they find that
change does not occur (or does not occur fast enough) and that their family
member has permanent residual effects from the injury, the family may ultimately reject the disabled person with brain injury. This swells the ranks of
long-term placements. Therefore, one goal in our evaluation is to assess the
stability of the current living arrangements of the client and, in particular,
determine how secure that living arrangement really is. A careful evaluation of
the expectations of the family members/caregivers is clearly important.
Many times, the patient who is discharged from an inpatient unit has
regained individual skills such as feeding, washing, grooming, or communicating, or some elementary social skills, but has not learned to integrate these
skills into functional, daily routines. Such people require a great deal of
supervision and structure. Often, a day program can take such a patient and
develop the integration of skills to the point where the family burden is kept
within tolerable limits; when this occurs, the family is more likely to provide a
long-term, stable destination for the patient.
A client's living arrangements are more likely to be stable if his or her
activity patterns are established and regular. In our view, the goal of braininjury rehabilitation is to produce an individual who will adaptively (not
necessarily independently) participate in any activities of daily life that he or
she is capable of. Our functional approach to brain-injury rehabilitation em-

9. Concepts in Day Programming

225

phasizes the adaptive, context-specific nature of the skills, routines, and activity patterns of daily life that we target for training. To help brain-injured
patients adapt to their actual or prospective living environments, we need to acquire a good deal of information about their premorbid activities, their current
activity patterns, and any future activity patterns that appear promising,
potentially stable, and feasible.
To obtain such context-based information, we have extensively modified an
instrument that was initially developed by Diller et al., called Activity Pattern
Indicators [1]. We call our version the Activity Pattern Analysis (APA) [2].
The AP A interview asks more than 100 questions about 32 areas of daily
life, categorized under 8 headings: 1) self-care; 2) housekeeping; 3) leisure;
4) personal finances; 5) personal business and consumerism; 6) community
mobility; 7) work-related activities; and 8) academic-related activities.
Caregivers and living arrangements have an important i~pact on the eight
categories of the activity pattern. When patients are evaluated for admission to
the day program, these living arrangements are usually an established fact and
are incorporated within the interview. Frequently, the interview is conducted
within the destination setting itself, and family members are asked to help
provide information about the client's activity patterns. Specific information is
sought within each of the eight categories. For example, the third area of daily
life that falls under the category of "personal business and consumerism"
identifies the client's activities with regard to obtaining household goods and
services. Does the client shop for groceries, read labels, and prepare written
lists when needed? When in a store, does the client actually use a shopping list
that he or she previously prepared at home? Does the client understand the
need for obtaining personal services (e. g., a haircut, eyeglasses, clothes, and
pharmaceuticals), and can he or she prepare for and evaluate these services?
Current client activities are noted, as well as whether current activities
represent a departure from premorbid activity patterns. When it has been
established that a client currently performs a specific activity, information is
sought about 1) the client's understanding about the need to do the activity;
2) any preparatory actions; 3) his or her actual performance of the activity
and evaluation of its outcome; 4) the ability to terminate the activity; and
5) whether and how the client cleans up following the activity or sets up
for future activities.
The APA interview includes gathering information about the premorbid
period and the postmorbid activity pattern before the client's entry into the
day program. The premorbid APA is ascertained from a reliable informant for
the period immediately prior to the brain injury. The premorbid activity pattern helps establish a contextual understanding about the patient, about what
was relevant in his or her life, and what is likely to be relevant to the patient
and the family system in the future. Many clients applying for admission
to a day program have already been living at home or in a stable, nonhospital
environment. For such patients, a current AP A can be useful in defining the
adaptive difficulties experienced by the patient and family at home.

226

II. Rehabilitation Programs

A description of a typical "day in the life" of the patient is included in the


APA structured interview. The interview not only extracts information about
outcome but also seeks to identify process information. For example, is the
client experiencing problems with time management? Does the client initiate
activities? Does the client effectively organize skills into routines, and routines
into a daily activity pattern?
Both the premorbid and the current AP A information contribute to the decision to admit the client into the day program and enable the staff to set initial
goals. In conjunction with available medical data and previous test results, the
AP A information points out areas of relevance to the life of the client and his
or her family, and helps generate initial hypotheses about impaired skills and
routines. These impairments are verified by the day-program staff through
objective assessment of the client in a variety of situations and evaluation of the
client's functioning.
The AP A is the first step in a three- to five-day comprehensive outpatient
rehabilitation evaluation (CORE). The CORE is conducted to identify an accurate description of skills, strengths, and weaknesses of a client. This evaluation consists of traditional measures of the presence or absence of brain damage,
traditional measures of vocational readiness, and more recently developed
individual and job-skill inventories. During the evaluation phase, previous
records are examined and pertinent data are recorded. During the CORE, staff
will verify and classify the diagnosis of brain damage secondary to traumatic
brain injury and will help determine the client's learning potential.
Once the AP A has been completed, the program staff will identify areas in
which further, more in-depth evaluation needs to occur. For example, the
AP A may reveal that the brain-injured person does not use public transportation. The client or his or her family maybe afraid of trying this for a variety of
reasons. Because the use of public transportation may be a necessary prerequisite for the client's work or the establishment of some other activity
pattern, a staff member will actually take the client on public transportation to
evaluate strengths and weaknesses in this process. This includes not only
evaluating concrete steps in using public tranportation but also evaluating
how the client handles frustrations brought on by unpredictable occurrences
or weaknesses in the transportation system itself.
A second example would be the performance of a housekeeping chore-for
instance, doing laundry. Perhaps the client was responsible for his or her own
laundry or the family's, or perhaps it would be helpful to the family if the
client assumed responsibility for doing laundry now. If taking care of the
laundry would be an activity likely to have meaning in the ultimate activity
pattern, professional staff would ask the client to do a load of laundry. The
client would, if possible, bring a load of laundry from home and go through
the process-from sorting clothes before placing them into the washer, to
folding clothes after removing them from the dryer. Laundry is a complicated
routine and requires significant cognitive processing. The steps must be proper-

9. Concepts in Day Programming

227

ly sequenced. Clothes must be sorted, dials must be set, detergent and bleach
must be measured, directions on containers must be read. During the process,
staff members observe the client, determine where the process breaks down,
and try various intervention strategies to learn the most effective ways of
assisting the person to overcome apparent obstacles.
If the client demonstrates poor attention to details, the staff person may
attempt to highlight pertinent details. For example, if the client misses information while reading the directions on a detergent box, the staff member
may use a "highlighter" pen to accentuate the necessary steps. If this is not
sufficient, the evaluator may print the directions in short, concise commands
in order by number on an index card. A check-off system may also be added.
The intention is to determine if there appears to be potential for the client to
learn strategies to overcome apparent deficits.
Similarly, routines in all areas of pertinence to the client are evaluated. A
client may cook a meal (self-care), operate a vacuum cleaner (housekeeping),
complete a craft project (leisure), balance a checkbook (personal finance),
shop for groceries (personal business and consumerism), obtain information
necessary for an outing (community mobility), perform a work task (workrelated activities), or read a magazine article (academic-related activities).
To sum up: The CORE identifies available learning strategies, identifies
activities that might go into an activity pattern, and gives staff evidence for projecting an ultimate activity pattern outcome (e.g., competitive employment,
supported work, volunteer work, school, leisure pursuits, or a combination
of these).
4.1 Case Illustration: Perry

Perry was a 32-year-old high school graduate and photofinisher. In 1982, he


sustained a brain injury with a depressed skull fracture and an epidural hemorrhage in the right parietal area. A craniotomy was performed on the right side
to remove the large epidural hematoma. After being in a prolonged coma, he
recovered consciousness but was left with a residual left hemiparesis, which
partially improved. Five months after the injury, he developed recurrent
seizures that were complex, sensorimotor, and stereotypical.
Four years after his brain injury, Perry was referred to our day program. In
the intervening interval, he had only been seeing a clinical psychologist for
treatment of depression. Ironically, a psychiatrist who evaluated him several
years after the brain injury thought that Perry would do well in a day rehabilitation program with structured day activities but that he was not a candidate for individual psychotherapy unless this was part of a multidisciplinary
program. Perry then began seeing the clinical psychologist, who, in fact,
began psychotherapy to help Perry with his sense of inadequacy, low selfesteem, and feelings of hopelessness. After a long period of psychotherapy, he
was tested neuropsychologically. Findings indicated "organic brain pathology
affecting predominantly the right nondominant hemisphere with particular

228

II. Rehabilitation Programs

dysfunction of the right parietal region." The neuropsychologist also commented on Perry's tangential thinking and noted that he was unable to "get
to the point" during a conversation. He suggested "some form of sheltered
workshop environment in which he can receive significant and ongoing
supervisory support." He did not recommend that Perry attempt vocational
placement or training until his seizures could be brought under better control.
Because of vocational placement issues, Perry was referred to our day program four months later. During the AP A interview at intake, it was determined that Perry lived at home with his parents and his 27-year-old brothel",
and that he had lived at home with his parents prior to the injury as well.
Throughout the interview, Perry's conversations were tangential, and he
had difficulty answering questions directly and getting to the point. Although
he frequently strayed from the topic, he was usually able to catch himself and
get back to the topic. The staff was aware that, because of Perry's memory
problems and cognitive limitations, the accuracy and reliability of his selfreport would have to be corroborated with other family members. The staff
also suspected that the "typical" day that Perry described might actually
contain a number of activities that were performed over several days, rather
than on just one day.
4.2 A Typcial "Day in the Life of Perry"

Perry arises at 9:30 a. m., smokes a cigarette, and turns on the stereo. Then he
showers, dresses, and prepares his own breakfast. He reads the newspaper
while he eats breakfast. After he finishes eating, Perry goes to the post office
to check his box. He might visit a friend's record store near the post office.
Perry "hangs out" there for approximately an hour. Twice a week, he sees
his psychologist. Afterwards, he works out at a health club. On other days,
he rides a stationary bike. (When this was further explored, it was discovered
that he only rides the bike for six minutes.)
On a typical afternoon, Perry indicated, he does "nothing." He stated that
he feels like he is retired, and often he is very bored and depressed. At five or
six o'clock in the evening, Perry eats dinner (which is prepared by his mother).
He helps clear the table and washes dishes. His parents are "too nice" and do
not require much of him in the way of other chores.
After dinner, Perry takes a nap, and then may go to shoot pool at a friend's
house. Upon further questioning, he admitted that more typically he stays at
home, watches TV, and makes phone calls. He goes to bed about 12:30 a.m.
and sometimes has difficulty sleeping.
4.3 Summary of Daily Life Categories of the APA
4.3.1 Self Care

Perry has no problems performing basic self-care actIvitIes such as eating,


showering, dressing, toileting, and so on. However, he indicates that when he
walks, sometimes his "equilibrium is off" and he may stumble and fall.

9. Concepts in Day Programming

229

4.3.2 Housekeepin,r;

No significant changes were noted in Perry's level of participation compared


to his premorbid activity pattern. His parents had always assumed the responsibility for the majority of housekeeping activities. Chores for which Perry is
responsible include occasional dishwashing, laundry, cleaning up his room,
and preparation of his own breakfast. Perry says that he can prepare simple
hot meals but that he does not do so because his mother cooks for him.
Since his injury, Perry's participation in general home maintenance and
repairs has been significantly reduced.
4.3.3 Leisure

Before his injury, Perry's interests included dancing, bike riding, photography,
and drinking with friends. Currently, he spends more time alone and engages in
sedentary activities. He says that it is "hard to be around people and sometimes
I want to fall into a hole and disappear." Perry's leisure activities include
reading, taking walks, working out, and talking on the phone. His social
interactions are less frequent than before, and he reports that he gets along
better with strangers than with old friends. He visits a bar or nightclub several
times a week.
4.3.4 Personal Finances

Currently, Perry has little responsibility in this area. He has no savings or


checking account and does no banking. He cashes his Social Security Disability
Income (SSDI) check at a local pawn shop and purchases money orders at the
shop to pay his bills. Each month, he pays rent and pays for his phone,
transportation, and medical insurance. Perry has no formal budgeting system.
However, prior to receiving his monthly check, he tries to figure out how
much money he owes and what he will have left to spend.
4.3.5 Personal Business mId Consumerism

Perry sees a psychologist twice a week. He is responsible for obtaining and


arranging for his own personal services. Perry's mother does the majority of
the household and grocery shopping. However, Perry indicates that he may
pick up a few things that they need. Perry occasionally dines out by himself,
less frequently than he did prior to his injury. Basically, there has been no
change in his participation in community activities. He occasionally goes to
church and reports that he is a registered voter.
4.3.6 Community Mobility

Perry's primary means of transportation is the bus. Occasionally, he may ride


a train. He says that he has no difficulty using schedules or riding public
transportation. He reports that he does not purchase tokens but instead pays
cash for his bus fare.

230

II. Rehabilitation Programs

4.3.7 Work-Related Activities

Before his head injury, Perry worked as an assistant printer. He is currently


not working. He left his last job at a portrait gallery in a shopping mall because
he was not bringing in the "required quota." Perry has had several other jobs
since his brain injury, most of which have been terminated due to his difficulty
in getting along with others on the job.
4.3.8 Academic Activities

Perry is a high school graduate and was an average student. (When he was
questioned further, he indicated that he received mostly C and D grades.) In
1974, he enrolled in a photography course at a technical school. Currently, he
is not participating in any educational activities.
4.4 Recommendations

Even a cursory analysis of this information indicates that Perry's actIvIty


pattern is currently diminished and less routinized in comparison with his
premorbid activity pattern. Perry appears to be dissatisfied with his current
situation, but it is not clear how he would like to improve it. Perry's inability
to retain stable work suggests the need for guidance and training for him to
accomplish any work goal that he might be agreeable to. He might also benefit
from increasing his leisure activities within the home and the community,
with special emphasis on socialization.
After they completed their evaluation, the staff of the day program met with
Perry and his parents to present their findings. Perry and his parents were
informed that it would be important to refer Perry to a neurologist to reinvestigate the nature of his seizures. He had been taking phenytoin (Dilantin)
for about 3 1/2 years; given the symptoms he described, another drug might
be more appropriate. Our staff felt that, once the neurological picture was
clarified, Perry might have vocational potential that could be developed through
a situational assessment and treatment approach. Leisure skills and socialization strategies were to be included in the day program as well, to help Perry
improve his low self-esteem.
5 IMPLEMENTATION OF THE DAY TREATMENT PROGRAM

Classic rehabilitation aims at physical restoration through functional-skills


retraining. Once skills are retrained, it is expected that the patient will use
every restored skill as circumstances require. This approach assumes that the
patient will be able to integrate each restored skill within the context of the
clusters of skills and routines needed for daily life.
Brain-injury rehabilitation differs from this classic conception. The rehabilitation program not only must focus on training individual skills but must also
concentrate on restoring the patient's ability to integrate these skills in problem-specific, context-specific ways. For this reason, "independence" in skill
performance (e. g., "independence" in ambulation, "independence" in activi-

9. Concepts in Day Programming

231

ties of daily living, etc.) is not an acceptable standard of outcome when


rehabilitating brain-injured patients. Generally speaking, standards of independent skill performance assume that a patient can successfully integrate
newly retrained skills within the context of the patient's activity patterns of
life-an assumption that does not stand up well to clinical experience with
brain-injured clients.
Therefore, in our program, we take a "nested-skills" approach to retraining
the competencies of daily life. We emphasize what skills the patients need to
adapt to their home and community environment and train them in specific
adaptive behaviors, rather than in general abilities. For example, we train
walking to the bathroom rather than gait training, preparing a sandwich rather
than performing eye-hand coordination exercises, making a shopping list
rather than memorizing computer-generated lists. Component skills of these
activities are identified, evaluated for each patient, and then taught. In our
view, a functional deficit exists only in the context of a specific situation.
Therefore, individuals must be examined in context to determine how their
residual deficits will affect them.
Skills are the smallest fundamental units of "real life" function. When groups
of skills are linked together by a common goal, we call them a "functional
routine." Clusters of routines that fill the day in the life of an individual
constitute that person's activity pattern of living.
Assessment and implementation of treatment goals and objectives really
revolve around context-specific situational assessments and training. In Perry's
case, the staff discerned that he had had difficulty getting along with others on
a variety of jobs that he had tried after his head injury. Given his history, the
staff felt that Perry's future success at acquiring and retaining a job would
depend not only on the content of the job but even more on his ability to get
along with others. Therefore, Perry's training program within the day program was to include a series of job trials in which getting along with others
would playa prominent role. Socialization was a great need, and social skills
were felt to be important for Perry's future success in a job role, for acquiring
and retaining friends, and for shoring up his personal self-esteem and selfImage.
6 ACTIVITY PATTERN

A stable activity pattern can be made up of a combination of activities. The


number of activities will depend on the needs of the individual. For most
nondisabled people, work activity is responsible for filling most of their days.
When possible, work is an ideal activity to place into the activity pattern of a
person with a brain injury.
6.1 Competitive Employment

In some cases, if a brain-injured individual has had a good work history, and if
his or her employer is interested in doing whatever is necessary for the valued

232

II. Rehabilitation Programs

employee, there may be an opportunity to return the client to his or her


previous job (provided, of course, that the client is capable of doing the work).
Andrew's case is an illustration of this. Andrew was struck by a car in 1984.
He had a period of altered consciousness and posttraumatic amnesia that
extended for at least several weeks. He had a right hemiparesis, and at the time
of entry into the program he demonstrated residual weakness and stiffness
in his right arm, leg, and face. Memory difficulties, emotional flatness, poor
organization and planning, and lack of initiation were also evident.
Prior to his injury, Andrew worked on an assembly line, fitting together
pieces of automobile carburetors. He was on a break from work when he was
struck by the car. The employer was very involved and interested in Andrew's
rehabilitation. Project staff spent days observing other employees perform the
job, videotaped the performance of the job, simulated the job for Andrew, and
supervised him on thejob. Ultimately, Andrew returned to his former job. He
started back part-time but has been working full-time for over a year.
Situations in which employers are sympathetic and in which clients can
return to their previous positions (or similar positions) in the company are the
easiest to deal with. It is somewhat more difficult to place clients who have the
potential for competitive employment but who have a poor work history and
have no job to return to, or those who are unable to perform their previous
jobs. In these situations, clients must be placed into competitive positions
commensurate with their abilities. Program staff must create simulated work
trials to establish a client's ability to perform real work.
Our program uses work trials in various departments of the hospital (e.g.,
maintenance, housekeeping, dietary, gift shop) or volunteer positions in other
settings (e.g., the zoo, nursing homes, nursery schools). Clients are given
volunteer jobs and are trained to perform the necessary job tasks. Once the
clients demonstrate an ability to work, the vocational counselor looks for
competitive jobs in the community that might be suitable for the clients. When
a job is identified, the client is placed. However, rehabilitation is only partially
completed with job placement. After a client has been placed, project staff
members need to work with the client on the job to be sure that the job is
learned and completed satisfactorily.
This second scenario was followed for Perry. Perry was not working just
prior to his entrance into our program; he had held several jobs and was unable
to retain any of them. As part of his program, he was placed into a work trial
in the hospital gift shop. This work trial was selected for a number of reasons.
For example, Perry had demonstrated difficulty with social interaction in the
past, yet he liked people and had some intact relating skills. The gift shop
allowed for constant opportunities for social interaction. His job included
waiting on customers, operating the cash register, making change, stocking
shelves, and closing the store.
Staff members were on the job with him to teach him the various job tasks;

9. Concepts in Day Programming

233

they wrote out all the procedures in short, concise sentences, with written cues
about what steps to follow. He was also given demonstrations of the procedures. Gradually, supervision was reduced. Initially, when Perry was left
alone, some problems arose. For example, during the closing procedure, he
would leave drawers unlocked. Because this was a security problem, program
staff had to develop a procedure to help Perry remember this step. It was put
into a closing checklist that had to be completed.
Perry also left the job earlier than the agreed-upon time. This indicated a
problem with taking responsibility. Not only did Perry have to answer to
program staff, he had to answer to the operator of the shop, who would not
tolerate having the shop closed early. Once again, staff had to intervene,
through job coaching, to ensure successful job performance.
The vocational counselor had been seeking a competitive job for Perry in the
community. After Perry had been successful for two months in his part-time
work trial, the counselor started to take him on job interviews. Perry needed
assistance in developing good interview skills. For example, he tended to start
his interviews by telling the prospective employer what he could not do and
by focusing on the negative effects of his accident. He needed to be instructed
in how to focus upon his strengths and how to convey what he had to offer the
employer. Videotapes of work interviews were useful in this process.
Perry was eventually placed in the stockroom of a company that packages
refrigerated products. His job duties include locating stock items on shelves,
gathering stock items for distribution to various warehouses, and taking inventory. Cognitively, Perry is aided by color codes and by having listings of
all items attached to warehouse shelves.
6.2 Volunteer Work

For some clients, a volunteer activity pattern may be the most desirable and
suitable. Many nonprofit organizations rely upon volunteers to provide vital
services. Volunteer work is real work.
Jerry's case is a good example of this. Jerry had been a lieutenant in a fire
company. He fell from a ladder, landed on the left side of his body, and
sustained a head injury and multiple musculoskeletal fractures. Before he
entered the day program, he required two inpatient rehabilitation stays, the
second one for surgery to correct heterotopic ossification in the region of the
left hip. The ossification process had prevented him from walking for more
than three years. At the time of the second discharge from acute rehabilitation
in November, 1983, Jerry's condition was summarized in this way:
The craniocerebral trauma has left him with wcakness on the left side of his body. In
addition, he has multiple fractures of his left femur and has undergone surgery for
heterotopic ossification of the left hip. He has moderate cognitive deficits, including
decreased organizational abilities, memory deficits, difficulties integrating and sequenc-

234

II. Rehabilitation Programs

ing information, and visual-spatial deficits. He also has ambulation deficits (including
problems with balance and posturing), difficulties in dressing himself, and mild deficits
in transferring himself between his wheelchair and his bed.

At first, the day program worked with Jerry to establish a home routine
whereby he became more responsible for self-care and for housekeeping routines. (He lived with a roommate, who became a vital member of the rehabilitation team.) Emphasis was also placed on leisure and personal business
pursuits in Jerry's home community. A system of index-card files was developed to help Jerry manage many of his own affairs. Still, Jerry was left with
residual deficits that continued to make competitive employment unlikely.
Ultimately, his activity pattern was developed to primarily consist of a fiveday-a-week volunteer job as a "patient representative" in the hospital. The
patient representative visits all newly admitted patients, informs them about
the services available to them, answers any questions, and attempts to find
solutions to any problems. With the aid of a psychologist who acted as a job
coach, Jerry was trained to perform this job. The psychologist had to actually
perform the job to learn the intricacies. She also had to observe others doing
the job. Once this was done, Jerry had to be taught the various components
of the job. This was accomplished through role-playing, with the assistance
of videotaped feedback.
Perhaps more important, an organizational system had to be developed to
ensure Jerry's success on the job. For example, because Jerry is unable to write,
he uses a memo writer (a compact, calculator-like typewriter) to write notes
about the visit. He uses a pencil to keep his place as he goes from item to item
in his script. He needs a stapler to staple his notes to the interview form.
(Others who perform this role write their responses directly onto the interview form; Jerry is not able to do this.) Jerry could not carry the necessary
materials from room to room. Originally, a lapboard was designed to fit onto
Jerry's wheelchair; the stapler, pencil, and memowriter were fastened to the
board with Velcro. In addition, a series of instruction sheets was developed to
guide Jerry through the interview process and take him step-by-step to the
conclusion. These instruction sheets were put into a binder with specially
marked pockets that identified the information to be kept in each section.
Jerry eventually learned the job and has been successful in his position for
about a year. However, he requires intermittent assistance. For example, a
major revision in Jerry's system became necessary when his ambulation improved and he no longer needed to use a wheelchair. Although he is now able
to walk with a cane, he still cannot carry the materials necessary to perform the
job. A creative solution was needed. A small shopping cart had been donated
to the program by a large department store. The cart was fitted with a cover to
hold the necessary equipment. Jerry now pushes his cart from room to room
as he conducts his interviews. When items fall from the cart, he has a
mechanical "grabber" that he can use to pick them up without having to
bend over.

9. Concepts in Day Programming

235

6.3 Quality of Daily Living

Some clients may not even be able to maintain an activity pattern of volunteer
work-for example, Beth. In 1979, when she was 21, she was injured in an
automobile accident. After several inpatient rehabilitation stays, she continued
to demonstrate residual quadriparesis, inability to speak, and moderate organic
brain syndrome. After her final discharge from the inpatient program, Beth
was dependent on others for most activities of daily living, although she could
feed herself if food was placed in front of her, and she could communicate
using an electronic nonvocal communication system if it was set up for her.
Beth entered the day program four years after her injury. The major goal of
the program was to establish a routine daily activity pattern that would enable
her to have as active and productive a life as possible.
Beth was a cheerful and pleasant person who enjoyed contact with other
people. (During rehabilitation, she had also undergone considerable behavioral
programming designed to assist her in being less demanding of immediate
attention.) Staff members made considerable efforts to establish activities in
which Beth could participate. For example, prior to her injury, Beth had
enjoyed art and music. Staff members explored activities in these areas to find
some aspects that she could take part in. She also developed a desire for
information about motivation, perseverance, and the meaning of life. These
areas-art, music, psychology, and philosophy-could form the basis for
the development of an activity pattern. In addition, Beth needed physically
oriented activities designed to prevent atrophy and to help her maintain her
weight.
The ultimate plan for Beth was a schedule that included activities in her areas
of interest as well as self-care routines. An activity notebook was developed
that would eventually include all information that Beth needed to manage her
own affairs. As a result of therapeutic involvement, Beth's cognitive awareness developed nicely, and even though she was very limited physically (and
was physically dependent on others), she could direct most of her own affairs.
Beth's activity notebook included sections for schedules (and blank schedules), trip-planning procedures, school activities, art, games, self-care, a
journal, access guides (information about accessibility of public buildings,
recreational facilities, and cultural institutions), instructions about how to use
her voice synthesizer, and clothing she wants.
Under the "school" heading is information about a local community college
where Beth attends classes. She does not take the courses for credit, but audits
them. The college's coordinator for students with disabilities worked closely
with staff from our program to establish a procedure whereby Beth can make
appropriate course selections, establish necessary contacts with professors, and
arrange transportation. Because of her interests, Beth initially selected courses
relating to psychology. Guidelines for classroom behavior were established by
Beth and program staff to promote a successful experience for Beth and her
classmates. There is also a report sheet for Beth to complete after each class to

236

II. Rehabilitation Programs

help herself and others evaluate her experience and to allow staff and the
instructor to make necessary modifications. Similar procedure lists have been
established for other activities in Beth's activity pattern.
One last word about Beth's activity pattern: Arrangements were made to
have her write a column for the community college newspaper. This was seen
to be a productive, meaningful activity that would enhance Beth's self-image
and provide her with even more opportunities for socialization.
6.4 Follow-up

In all cases, continued follow-up is needed to ensure continued stability of the


activity pattern. It cannot be assumed that after clients have been discharged
from day programming, they will no longer need assistance. Activities that
make up the activity pattern may change. Changes may also occur because the
availability of the necessary resources is altered or because the client's interests
may change. Changes in conditions on the job, or in the client's life circumstances, may create situations that might unwittingly alter the client's
ability to perform the job successfully. For instance, an employer may want
to promote an employee with brain injury, without realizing that a change in
responsibilities would markedly affect the employee's performance.
It may be necessary to maintain regular contact, even if only once a year,
with clients who have been placed into competitive employment. For example,
Andrew initially required monthly visits from a staff member to troubleshoot
problems that developed; contact with him now has become less formal.
Perry has just been placed in a job; to ensure his continued successful
performance, it is expected that regular (at least weekly) contact will be
necessary until he has learned the job.
Jerry frequently asks for varied activities and changes in jobs. Although he
enjoys his duties as a patient representative, he has expressed boredom and
would like to learn new jobs. Unfortunately, the work that goes into training
him for new duties is costly in time and financial resources. In addition, he
does not always realize the severity of the residual deficits he exhibits. Other
people who want to help may inadvertently accede to Jerry's wishes without
fun consideration of the consequences. Counseling with Jerry and others
involved in his volunteer supervision is necessary on a regular basis.
Beth continues. to attend the day treatment program one afternoon a month
for monitoring of her activity pattern. Program staff members assist her by
making recommendations for changes or for more effective ways of accomplishing her objectives.
7 SUMMARY

In our view of rehabilitation, adaptation of the survivor to his or her social and
environmental context is essential, meaningful, and practical. We believe that
issues of destination and activity pattern should form the basis of rehabilitation programming. A day program located in the community setting of the
survivor is best suited to deal with these issues in a realistic way.

9. Concepts in Day Programming

237

REFERENCES
1. Diller, L., Fordyce, W., Jacobs, O. and Brown, M. (1981). Rehabilitatioll l"diwtors Pro;ect.
Institute of Rehabilitative Medicine, New York.
2. Copies of the AP A can be obtained from the Drucker Brain Injury Center, Moss Rehabilitation
Hospital, 12th St. and Tabor Rd., Philadelphia, PA 19141.

III. NEUROPSYCHOLOGICAL
REHABILITATION TECHNIQUES

10. NEUROPSYCHOTHERAPY

DAVID W. ELLIS

1 INTRODUCTION

The individual neuropsychological treatment of the survivor of brain injury


is a controversial subject. Many professionals question whether individual
treatment is beneficial [1]. Often, traditional procedures must be altered
substantially to be applicable to a brain-injury survivor.
In this chapter, the history of the treatment of the brain-injury survivor will
be outlined, as well as some current perspectives. After the personality and
behavioral characteristics of survivors are described, a model of individual
treatment will be proposed by outlining 1) a theoretical view of personality,
2) the phases of the ncuropsychotherapy process, and 3) the techniques and
strategies of neuropsychotherapy.
2 REVIEW OF INDIVIDUAL TREATMENT AFTER BRAIN INJURY

In brain-injury rehabilitation, the understanding and clinical applications of


brain/behavior relationships must be used to provide appropriate treatment.
Of special significance for the early phases of this brain/behavior tradition is
the research by Broca and Wernicke in the 1860s and 1870s [2, 3]. These initial
discoveries led many neurologists to believe that the key to understanding
behavior and language was through delineating brain functioning. Although
there had been progress in understanding brain function before this time, the
interconnection of psychological and neurologic perspectives was a new
approach.
241

242

III. Rehabilitation Techniques

2.1 Historical Perspective

Following the work in neuroanatomy and clinical practice, neurologists such


as Charcot [4] and Janet [5] attempted to understand individuals by means
of their responses to treatment. However, it was a practicing neurologist,
Sigmund Freud, who provided one of the richest and most carefully described
theoretical bases for understanding personality and treatment. Freud was not
content simply to describe human behavior. He began to search for explanations, which led to what he termed "the royal road of the unconscious" [6], or

psychoanalysis [7].
Freud's writings supported the centuries-old tradition of a "talking cure,"
whose roots are described in Judeo-Christian thought [8]. The "talking cure"
tradition is based on the epistemological premise that "the truth shall set you
free." Although this is the basis for psychological treatment, there is still a
controversy in brain-injury rehabilitation circles as to whether the "talking
cure" is efficacious for people after brain injury.
Kurt Goldstein [9], a German neurologist who treated brain-injured patients
during and after World War I, argued that the best treatment for brain-injury
survivors should be conducted in a protective and supervised environment,
since the individual's ability to think independently was lost. However, this
rehabilitation process consisted of accepting one's disability and remaining in a
supervised, yet dependent position.
Given this historic perspective, it was inevitable that practitioners of traditional treatment procedures considered psychotherapy to be an inappropriate
treatment for brain-injury survivors. In addition, the psychodynamic psychotherapists coming out of Germany, France, and Switzerland at the turn of the
century also considered brain injury as a condition that was incompatible with
their particular type of psychological intervention. The theory behind psychodynamic psychotherapy or psychoanalysis is based on the assumption that the
person has an intact central nervous system [10].
For example, one underlying premise of psychodynamic treatment is the
concept of interpretation of unconscious material into consciousness. Interpretation requires a higher level of concept formation than was considered possible for the brain-injury survivor. Therefore, it has been generally accepted
that individuals who sustained brain injury could not be helped by traditional
psychotherapy [11].
However, Alexander Luria [12], a Russian neurologist-neuropsychologist,
proposed a more optimistic view of the rehabilitation process. According to
his theory of neuropsychology, recovery after brain injury is possible through
specific methods of intervention, which are based on an understanding of
brain-beha vior functioning.
2.2 Current Perspectives

Over the last 20 years, advances in modern medical technology have enabled
approximately 50% of victims of severe brain trauma to survive. The increasing

10. Neuropsychotherapy

243

incidence of traumatic brain injury (TBI) compels us to understand brain-injury


rehabilitation and attempt to extend our current knowledge. As recently as 15
years ago, the psychological treatment and rehabilitation of brain-injured patients was considered a limited and unfulfilling area by many professionals.
However, psychological and rehabilitation research suggests new optimism
regarding the potential for recovery. As a result, many centers for rehabilitation
of these patients have opened in this country and throughout the world [13].
The use of individual psychotherapeutic treatment after brain injury remains
a rare subject for study and practice. Prigatano [14] proposed that psychotherapy could be of value to brain-injury survivors in several areas. For
example, it could help brain-injury survivors 1) develop a model of the brain
injury; 2) understand the implications of the brain injury; 3) accept the injury
and develop forgiveness for themselves and others; 4) make a commitment to
relationships and work; 5) improve their ability to handle social situations;
6) develop strategies of behavior to compensate for deficits; and 7) feel hopeful about the future. These application areas can serve as guidelines for the
treatment process.
3 CHARACTERISTICS OF THE BRAIN-INJURY SURVIVOR

For a therapist to competently treat a brain-injury survivor, he or she needs to


thoroughly comprehend the neuropsychological functioning of the survivor.
A therapist needs to individualize treatment based on the personality characteristics of each brain-injury survivor. These characteristics vary, depending
on the severity of the brain injury and the patient's premorbid functioning.
In the next section, some frequently encountered characteristics of braininjury survivors will be described, in order to illustrate typical posttraumatic
personality variables and behavioral functioning.
3.1 Personality and Behavioral Changes

Lezak [15] described alterations in personality functioning after brain injury in


relation to a social learning context. She described five broad areas of impairment that influence the brain-injury survivor's adaptive behavior: 1) social
perception; 2) self-control and regulation; 3) planning, organization and initiation; 4) emotional control; and 5) ability to learn from experience. Benton [16]
categorized the behavioral functioning after closed-head iUury. He reported a
complex of symptoms, including 1) concentration and attentional deficits;
2) fatigue; 3) memory deficits; 4) dyscontrol of emotional modulation difficulties (with lowered thresholds for noise and frustration); 5) alterations in
personality; 6) aphasias of various forms; and 7) sensory deficits (both basic
and higher level abilities). Brooks [17] studied subjects across a five-year span
after brain injury and reported that approximately 70% of them had major
personality changes, including depression, anxiety, mood swings, and childish
behavior.
One of the most comprehensive studies of recovery patterns has been

244

Ill. Rehabilitation Techniques

carried out by Goethe and Levin [18]. They examined not only the general
characteristics of brain-injury survivors in relation to the severity of the injury
but also the early and long-term stages of posttrauma recovery. They found
certain common characteristics, and rated patients as either Grade 1 (i.e., conscious at hospital admission, no neurological deficits); Grade 2 (i.e., unconscious for less than 24 hours with some neurologic deficits); or Grade 3 (i.e.,
unconscious for 24 hours or longer and manifesting neurological deficits).
They concluded that the severity of the personality alteration paralleled the
severity of the trauma to the brain. Severely injured persons (e.g., Grade 3)
manifested signs of emotional withdrawal, poor modulation of emotions,
conceptual disorganization, hostility and suspiciousness, motor difficulties,
unusual thought content, disorientation, and memory difficulties.
Ellis and Zahn [19] evaluated the psychological functioning of 35 young
adults with severe closed-head injury, at least one year posttrauma. They
found that the primary psychological difficulty evidenced was that of an
apperceptive disturbance, which appeared to be a result of a detached associative process. Perceptions were vague and impressionistic. The modulation of
affect was not controlled. In addition, disruptive emotions produced painful
experiences. Interpersonal relatedness was inadequate. For example, patients
were often not able to handle confrontations, even over small or inconsequential events, without first exploding into verbal and possible physical violence.

3.2 Effects of Catastrophic Brain Injury on Personality

Integrating the severe brain injury into the life of the survivor presents issues
similar to those described after other catastrophic events. A "catastrophic stress
reaction" [20] is a response by the human organism to a situation that is totally
overwhelming and that cannot be understood or mastered by the individual.
The catastrophic stress reactions of brain-injury survivors appear similar to the
stress reactions observed in survivors of wars.
Van Der Kolk [21] reported that after a catastrophic stress, a person has the
potential to react to highly emotionally stimulating experiences as if the experiences were a total resurgence of the original traumatic stress. Therefore, any
emotional stimulation might elicit an affective discharge of primitive, rigid,
and strong emotion. There are a number of reactions; however, the "fight or
flight" reaction is most common. Although the survivor often docs not have
any memory of the traumatic event, the unconscious psychological defense for
the survivor appears to be an attempt either to ignore the posttrauma changes
or to experience the trauma as a totally destructive event that has left him or
her with no future. Constant affective discharge interferes with a person's
ability to integrate a catastrophic experience into his or her life, to master
ordinary conflict-laden situations, and to accumulate experiences that are considered positive and comforting.
Bartemeir et al. [22] have described the psychological and biological re-

10. Neuropsychotherapy

245

sponses to catastrophic stress as a series of emotional reactions, with an initial


inability to master the catastrophic event. An inability to handle stress within
a tolerable limit may lead to adaptive failure. Such failure to adapt may be reflected by some or all of the following: 1) irritable reactions to others; 2) hypersensitivity to perceived threats; 3) sleep disturbance; 4) startle response or
reaction; 5) responses of the motor system (i.e., involuntary flight reactions);
6) withdrawal of emotions from awareness (e. g., through psychological
mechanisms; isolation; drugs; loss of interest in events, eating, and communicating with others; loss of interest in the welfare of others); 7) indecision, confusion, and problems in judgment; 8) psychosomatic responses to awareness
(including vomiting, tremors, and diarrhea); 9) attempts at passivity, with
subsequent behavior that is incoherent, wild, and reckless; and, finally, 10)
aggressive behavior, with rage, depressions, and fight and/or flight behaviors.
As part of the examination of ongoing stress reactions and patterns of response after trauma, Grinker and Spiegel [23] studied war survivors who had
undergone catastrophic experiences and reported on the development of a
posttraumatic stress disorder of an acute panic (i.e., fight or flight) type. They
found that those survivors with the highest incidence of severe panic reaction
had a premorbid history of adjustment problems and personality disturbances.
Although this description of psychic stress reaction was originally formulated
to describe those survivors of stress who have an intact central nervous system,
the characteristics described are strikingly similar to those exhibited by survivors of catastrophic brain injury, and the model may thus be quite useful for
understanding TBI survivors' behavior and personality changes.
3.3 Anxiety and Learned Helplessness

Distress mobilizes painful emotion, and Schafer [10] referred to physiological


distress states as the forerunner of an overwhelming anxiety. Klein [24] spoke
of this overwhelming anxiety as an unbearable feeling of "falling to bits."
This type of "falling to bits" is the experience spoken of by brain-injury
survivors and described by Piotrowski [25] as the "catastrophic panic" of the
brain-injured. Seligman [26] spoke of a pattern of helplessness or submission
after devastating events that is similar to depression, masochism, and a general
feeling of hopelessness. The degree oflearned helplessness will vary, depending upon the support system and whether the situation seems unmodifiable or
completely unbearable.
Seligman reported that affective states of helplessness are characterized by
dysphoria, affective arousal, and a period of physiological hyperactivity, with
an impulsive attempt at flight or fight until the emotion is contained. Van Der
Kolk [21] wrote that the state of emotional surrender (i. e., helplessness) after a
major trauma-after physical pain ends, cognition is regained, and the sense of
tragedy emerges-can be completely overwhelming. The patient's sense of
pleasure and gratification is diminished; indeed, the entire experience of success
and pleasure in life is altered.

246

III. Rehabilitation Techniques

Learned helplessness and posttraumatic stress appear to create feelings of


depression and hopelessness. These alterations in the self-regulatory affectual
systems may include severe emotional changes as well as physiological changes.
However, the "posttraumatic stress disorder" seen in non-brain-injured survivors of catastrophes is qualitatively and clinically different from that seen
in brain-injury survivors and should only be used as a rough model for the
emotional reactions to the brain injury.
3.4 Summary

To conceptualize and plan treatment for a brain-injury survivor, the therapist


must understand not only the recovery patterns but also the process behind the
brain injury. For example, the survivor's ability to organize his or her own
thoughts in a meaningful way, as well as the formation of memory, are critical
to successful treatment. The personality characteristics of brain-injury survivors have been described by many researchers and appear dependent on premorbid personality and social characteristics, as well as on the severity and
type of brain trauma [27]. The types of interventions necessary depend on the
characteristics of the survivor.
4 A WORKING MODEL OF PERSONALITY

Schafer [10] explained the importance of having a theoretical model for the
analysis of patients. He suggested that a therapist's ability to empathize with a
patient depends on the therapist's ability to construct a mental model of the
patient. For this reason, a model of the brain-injury survivor that outlines the
general areas of neurodevelopmental theory, personality structure, and posttrauma psychopathology appears to be a useful tool for developing treatment
strategies.
When treating children or adolescents who have survived brain injury, it is
important to remember that the injury will dramatically alter the developing
central nervous system-and will thereby alter their personality development.
In this context, the concept of habilitation replaces rehabilitation. If the brain
injury occurs at a very young age (i.e., in childhood), the integration and development of the selfhas not been completed, especially since at approximately
16 years of age there is another shift and expansion in personality development

[28, 29].

Freud's f7] general model of the structure of personality has been reformulated and described by Weiner [11] as consisting of three major areas: 1) the
unconscious; 2) conflict/defense; and 3) experiencing-selflobserving-self. After
a brain injury, personality functioning would also be complicated by an additional area: 4) the neurological deficits. The person's pretrauma personality
characteristics, which may be exaggerated or altered by the trauma as well as
by compensation for the deficits [30]. Depending on the extent and site of
damage, the brain-injured patient may have serious cognitive deficits. The injuries may affect the survivor in the areas of experiencing-self and observing-

10. Neuropsychotherapy

247

self, as well as his or her ability to process information and therefore to


adaptively cope with conflict.
4.1 Personality Structure

The psychodynamic theoretical model of personality functioning appears to be


a comprehensive and helpful one for this type of work. This model is still quite
useful in understanding personality functioning, especially when it is considered
with the posttrauma deficits. The psychodynamic concepts will be briefly described, in order to present a comprehensive and rationally defined model of
personality structure.
4. 1.1 Unconscious

The unconscious is viewed as all those feelings and thoughts that people have,
but which are not in their awareness. Neuroscientists have provided evidence
that the brain processes contribute towards keeping certain information out of
consciousness as well as transforming original events into symbols [31, 32].
Although neuroscientists have sometimes questioned the concept of the unconscious, we assume that the unconscious continues to exist after a person has
sustained a closed-head injury. Questions have been raised about the presence
and nature of the posttraumatic unconscious material. Professionals sometimes
confuse the two concepts of unconscious material and physiologically altered
attention, memory and cognition [33]. This area of neuroscience is highly controversial and professional discussion quickly deteriorates into the historical
debate over the concepts of mind versus brain [32]. With these issues in mind,
the therapist needs to scrutinize his or her view of the unconscious and must
consider the possibility of having erroneously interpreted a particular behavior
as "unconscious conflict." A cognitive or physiological description may
handle the event just as well.
4.1.2 Conflict and Defense

The term conflict is generally used to refer to an individual's response to knowledge, wishes, or desires that are not acceptable to the conscious self. The
person experiences anxiety surrounding unacceptable thoughts or feelings, and
then defends against the conflict through behavior and thoughts that guard
against possible anxiety-producing experiences. These behaviors or thoughts
that guard against anxiety are termed difenses.
Conflict and defense are commonly present after head injury. For example, a
majorconflictual area for the survivor is the total physical and emotional dependency upon hospital staff, which produces an institutionally induced learned
helplessness. Other patients may defend themselves through running away or
withdrawal. The defense mechanism is evident in the patient's hostility that is
directed at family members and helping professionals in the later phases of recovery. This response appears to be a defensive yet necessary response to the
helplessness and stress. As a result of major catastrophe that produces stress,
conflict is heightened-and defenses are strengthened.

248

III. Rehabilitation Techniques

4.1.3 Experiellcillg-Self alld ObserviI1i;-Se!f

The term self [34] is used to refer to the physiological and psychological being.
The "experiencing-self" is that aspect of the self that is phenomenological and
which is therefore felt by the person. The "observing-self" is that aspect of the
self that enables a person to "step outside" of himself or herself and look at his
or her behavior. Usually, people are capable of certain degrees of both experiential and observational behavior. Both of these behaviors exist on a continuum, and vary depending upon age, life experiences, and cognitive abilities.
These last three variables are primary factors in understanding the treatment of
children, adolescents, and adults.
When the observing-self is severely impaired, the ability of the brain-injury
survivor to observe his or her own behavior may collapse. This collapse of the
observing-self is viewed by others as a "denial" of problems, whereas in reality
it represents a loss of the observing-self, either through fragmentation of personality or destruction of memory and other neurobiological structures. As
the observing-self is rebuilt through compensatory mechanisms, the "denial"
is usually resolved.
After brain injury, the experiencing-self and observing-self are usually drastically altered. The experiencing-self changes as a result of the loss of inhibition,
increased frustration, poor affective control, and depression. The observingself changes because of information-processing difficulties and the general
apperceptive disturbance. Because both the experiencing-self and observingself can be altered through the inability to feel and recognize experiences, the
personality changes can be dramatic.
While the brain-injury survivor is working to improve the specific cognitive
or behavioral areas, the issue of what it means to be "damaged" and a different
person than before needs to be addressed. How has the person's narrative [35]
about himself or herself been disrupted? Reconnecting the threads of the narrative and integrating the last traumatic experience (the catastrophic brain injury)
are major challenges for both the survivor and the therapist.
4.1.4 Behavioral Sequelae

To fully understand a brain-injury survivor's personality functioning, the


therapist must consider the neurological damage to the brain, along with the
survivor's behavioral characteristics and methods of compensation. In many
survivors, the contusions after brain injury are usually diffuse; however, the
damage may be more or less severe, depending upon the type of injury. For
example, beca~se of their location, the most vulnerable areas of the brain are
the occipital, temporal, and frontal regions of the cerebral cortex. In conjunction with this type of trauma, there is usually a shearing of the entire brain,
frequently involving the limbic system and other deep brain structures; this
shearing, combined with other damage, may lead to subsequent personality
changes [36]. The patterns of damage that typically results from accelerationdeceleration brain injuries (e. g., after motor-vehicle or other high-impact acci-

10. Neuropsychotherapy

249

dents) is that of primary-trauma injuries (e. g., prefrontal 10 be damage) with


secondary complicating consequences (e. g., increased intracranial pressure).
Although patterns exist, the effects of brain damage are generally unique to
each individual, and it is crucial to have a complete neurological and neuropsychological evaluation to understand the survivor's specific strengths and
weaknesses. For example, Heilman, Watson, and Valenstein [37] described
some characteristic behavioral sequelae of injury to the right hemisphere of the
brain. These general characteristics included left-sided visual neglect, left-sided
hemiparesis and other physical manifestations, emotional dyscontrol, and
denial of problems (especially those that are not immediately apparent). Predominantly left-hemisphere damage has been related to speech and language
deficits, feelings of hopelessness and depression, and poor self-esteem.
However, many researchers have discussed and agreed upon the usual residual problems after brain injury [17]. These include deficits offunctioning in
the following areas [27]: 1) cognitive organization; 2) memory; 3) fatigue;
4) irritability; 5) unrealistic self-appraisal; 6) poor planning; 7) poor interpersonal relationships; 8) poor family relations; 9) reduced social contact;
10) loneliness; 11) impatience; 12) verbal and physical aggression; 13) depression; 14) anxiety; and 15) labile emotions. Although the localized injury may
produce specific symptoms, severe diffuse brain injury results in a generalized
pattern of problems in adaptation.
4.2 Personality System

In the psychological literature [38], professionals have reported on methods for


understanding the self-system of an individual. This is not done to artificially
break apart the gestalt of the person or to reify the ego structure [39], but rather
to make sense of the behavior and personality that are observed. As Weiner
[11] has stated, any model of personality is simply a collection of constructs
aimed at simplifying conceptualization. However, conceptualization of personality functioning is important when the result of the conceptualizing is a
model that attempts to provide the therapist with a comprehensive clinical
picture of the self-system.
For purposes of description, the personality or self-system can be artificially
separated into the following concepts: 1) cognitive functioning, 2) affect,
3) interpersonal relationships, 4) identity, and 5) behavior [40]. This structure
can be used to attempt to understand the total person, no matter what the extent of the brain injury. Each of these concepts will be explained individually.
4.2.1 Cognitive Functioning

Within the context of personality structure, the term cognitive functioning will
be used to indicate thought processes and reality testing, as well as executive
functions.
Weiner has described thought processes as consisting primarily of "cognitive
focusing, reasoning and concept formation" (p. 16) [39]. He asserts that the

250

III. Rehabilitation Techniques

hallmarks of thinking include 1) scanning and selecting important information,


2) drawing logical inferences about the relationships between events and objects, and 3) being able to interpret these inferences at the appropriate level of
abstraction.
In addition, he considers the perceptual process essential to the patient's
"reality testing" and "reality sense." Reality testing is viewed as the "accurate
perception of the environment" (p. 16) and reality sense is the accurate sense
and perception of the person's body [39].
Freud [40] and Holt [41] have described other aspects of thinking, consisting
of the person's cognitive style, creative abilities, resilience, fantasy life, and
memory. For example, the cognitive and intellectual changes due to frontal
lobe impairment generally include 1) problems with the motor completion and
output of symbols, language, calculations and logic; 2) abstract informationprocessing and reasoning; 3) judgment; and 4) diminished capacity for concentration, focusing, and orienting place and time.
The executive Junctions are certain aspects of reasoning, judgment, initiation,
and motivation that affect the carrying-out of everyday life [42]. In this context, they can be thought of as a complex cognitive steering mechanism. Lezak
[43] described executive functions as the ability to conduct daily living and to
exercise interpersonal and intellectual skills in such a manner that supervision
is not needed. A major problem after brain injury is that the executive functions
are often damaged.
4.2.2 Affect

Affect is a term that is us.ed to describe the emotional life of the person. The
feelings, expression, and modulation of emotion (as well as general mood
states) are all a part of the overarching concept of affect. After brain injury and
possible damage to the prefrontal cortex and the limbic system (as well as
after seizures and interictal behaviors), the expression of affect may be strong,
unmodulated, and periodically accompanied by violent behavior.
Mackinnon and Yudofsky [44] summarized the affective alterations that can
result from damage to the frontal cortex. These include: 1) emotional shallowness, Jpathy, and indifference; 2) irritability, panic behavior, and lability of
affect; and 3) rage, violence, and general dyscontrol of behavior.
4.2.3 Interpersonal Relationships

The psychologically intact survivor is able to meet, form, and continue relationships with others at a satisfactory level. After trauma, however, this ability or
capacity for interpersonal relatedness may be altered. An inability to observe
onself and others appears to disturb interpersonal functioning through poor
social skills and an inability to empathize with others. At first, the survivor
often does not perceive interpersonal problems. The disturbed apperceptive
process is manifested by a childlike self-focus, which results in a restricted
world view.

10. Neuropsychotherapy

251

The changes in cognitive functioning may result in interpersonal relations


that are characterized by denial, poor planning, impulsive behavior, poor interpersonal skills, and behavioral dysfunction. The survivor who is unable to
engage in social relationships usually begins to withdraw from others, becomes
increasingly alone, restricts contact, and eventually becomes secluded, with a
diminishing interest in the outside environment. Over time, other problems
appear, including loneliness and impatience, as well as verbal and physical
aggression. This series of behaviors may occur due to the person's extreme
loneliness or isolation.
The interpersonal deficit is often manifested through changes in sexual functioning. Sexual expression is complicated and subtle and is frequently affected
by brain injury. The dysfunctional behavior may take various forms. As Boller
and Frank [45] have noted, some patients experience a reduction in sexual relations after brain injury. Blackerby [46] has descrbed the neuropsychology of
sexual behavior and the tremendous effect that a brain injury has on sexual
functioning. Patients often exhibit either hypersexuality or hyposexuality,
both of which appear to be a result of a complex interaction of the deficits of
cortical functioning and damage to the limbic system, medication, seizures,
and endocrine levels, as well as reduced interpersonal relations.
One couple's experiences [47] illustrate many of the concerns couples have
about the impact of brain injury on sexual relations. The husband, in his mid30s, reported an increase in sexual desire, with obsessive and inappropriate
behavior. His verbal sexual behavior was similar to that of an early adolescent,
and he had feelings oflowered self-esteem, dependency, and self-doubt about
his masculinity. His wife said that she kept comparing her husband to the way
he had been before the injury and also found that there had been a change from
a "wife-husband" relationship to a "mother-son" relationship. She felt pity for
herself and her husband, had been faking sexual satisfaction, and was feeling
attracted to other men. Similar experiences are reported by many couples after
one of the partners has been brain-injured [46].
4.2.4 Idelltity

Identity is defined as the integration of the experiencing-self and the 0 bservingself, resulting in an integrated self-system [48]. An area of conflict for the survivor is the feeling of being a victim. The survivor feels, "Why me? Why have
I been singled out for this 'vengeance of God'?" The passive helplessness of
this "victim mentality" must be changed into assertiveness; that is, the person
needs to recognize that he or she is a survivor, rather than a victim. An example
of this assertiveness is the emphasis that many survivors place on physical
exercise; they attempt to strengthen their bodies (even though their minds are
not functioning properly) to compensate for the cognitive deficits they have
experienced and to alleviate the conflicts that have been aroused. In addition,
the involvement of their physical condition gives them a rapid and visual display
of positive changes.

252

III. Rehabilitation Techniques

The interplay of a person's pretrauma personal narrative of himself or herself and the type and location of brain injury can produce disturbed behavioral
patterns. These behaviors are usually part of the reason that the survivors have
come into contact with the professionals who are treating them. Individuals'
narratives about themselves contain the story of their lives and their own
views of what their reasonable behavior would be. Disturbed behaviors need
to be brought into awareness in order to bring them under control. Bruner
[35] described a person's narrative as that story which the person uses to tell
the story of his or her life. The past contains the person's childhood, loves,
losses, work, and education. The catastrophic accident may-or, more usually, may not-be a part of the person's past. The present contains the difficulty
of the current rehabilitation program, and the future is an idealized "cured"
goal.
An inability of survivors to observe and monitor their behavior appears
to be a prime reason for the disruption of the individual's narrative that combines both pretrauma and posttrauma selves. The high level of abstraction
needed to observe oneself is shattered, and the person is caught up in the ongoing experience of day-to-day life, losing touch with overall reality because
of an inability to grasp the unique differences between events (i. e., the knowledge that two events were actually different in structure and content). Recognizing difference is crucial in understanding interpersonal relatedness; for
example, the difference between a smile and a frown depends on the interpretation of the structure of the lips, cheeks, and eyes. An understanding of the
difference between what a person feels and what he or she observes is critical
for the self-monitoring process. A person's observational level of awareness of
his or her own narrative reflects the degree of functioning; that is, the more
aware of self the person is, the better he or she is functioning.
4.2.5 Behavior

Behavior is altered by damage to the central nervous system. We have a fairly


accurate view of these behavioral problems following brain injury. For
example, damage to the perfrontal lobes usually results in the most severe
examples of behavior change posttrauma. In face, behaviors usually seen after
damage to the prefrontal lobes represent the most pervasive emotional and
behavioral difficulties seen in brain-injury survivors. Damage to other areas of
the brain, although devastating, do not usually result in the type of behavioral
syndromes that create the need for total supervision.
The behavioral changes due solely to frontal lobe damage may include [44]:

1. Exacerbation of pretrauma behavioral characteristics such as obsessiveness, suspiciousness, anxiety, and oppositionalism
2. Apathy and loss of concern about others, as well as lack of interest in
appropriate social behavior
3. Poor hygiene, lewd behavior, and loss of social grace

10. Neuropsychothcrapy

253

4. Increased use of profane language, intrusiveness, and unmodulated conversation


5. High risk-taking, gluttony, unrestrained use of alcohol or drugs
6. Distractibility, poor attention, and poor concentration

5 NEUROPSYCHOTHERAPY: A MODEL OF TREATMENT

The primary goal of neuropsychotherapy is to help the brain-injured person


form an integrated sense of identity, based in part on posttrauma elements of
personality, as well as the person's personality before the trauma. An understanding of the survivor's neuropsychological functioning and methods of
compensation for deficits is essential in order to understand his or her behavior
patterns. The model was derived from dynamic psychotherapy, as well as an
understanding of clinical treatment of brain injury and deficits due to trauma.
During the initial phase of treatment, the therapist assesses the brain-injury
survivor's neuropsychological functioning and forms hypotheses about his or
her strengths and weaknesses. Information is gathered from the medical
history, neurological evaluation, previous or present observations by staff and
therapist, family observations, and the client's self-report. The therapist then
integrates the information and designs specific intervention techniques that
will take advantage of the survivor's strengths and compensate for his or her
weaknesses.
After the patient's medical condition is stable, treatment can begin. The first
priority is to help the patient stabilize his or her sensorimotor and physical
functions; the next goal is to enable the patient to achieve higher cortical
functioning. However, individual neuropsychotherapy after brain irUury is
usually not enough to compensate for the extensive damage that has been
done. In such cases, individual treatment must be done in conjunction with
other modalities of rehabilitation. For example, one patient had a headon motor vehicle accident that left him with left-side weakness, seizures, poor
reasoning skills, and inappropriate displays of emotions. His therapeutic
regimen included antiseizure medication, physical and occupational therapy,
cognitive remediation, and group and individual psychotherapy.
In the initial phase of neuropsychotherapy, the therapist attempts to build a
relationship with the client through communication. A variety of techniques
may be used to assist the survivor to communicate. The techniques can be verbal (e.g., speaking to one another), visual (e.g., using pictures or actual displays of items to help someone whose verbal skills have been damaged), or
tactile (e.g., writing words on the person's hands). After that, the therapist
uses these communication techniques to explore the personality dynamics of
the survivor. The therapist attempts to help with identity refinement and tries
to understand and resolve the patient's conflicts. The consequences of the
injury upon the survivor's future are explored, as well as his or her ability to
resolve these conflicts.

254

III. Rehabilitation Techniques

5.1 Initial Phase of Treatment

The initial phase of individual treatment is crucial in any therapeutic relationship, but especially with a brain-injured person. The therapist must convince
the skeptical survivor of the value of the treatment. The first visits are used as a
time to become acquainted and to go through the initial behaviors (or rituals)
that a particular culture views as acceptable. These initial rituals include introductions and preliminary conversations about likes and dislikes, the reason the
patient is there (i. e., in neuropsychotherapy) and the role that the professional
will play in this relationship. There may also be a discussion about how this
relationship differs from a traditional friendship and about why the survivor
should even bother making the effort to be there.
The initial phase is composed of contract setting, building the relationship,
and introducing techniques for communication (some of which were mentioned above). During this process, the therapist may have to employ unusual
techniques for the survivor in a manner that is not accepted in the usual practice
of dynamic psychotherapy. For example, a therapist may bring in a third
person to role-playa particularly difficult situation, over and over, so that the
client not only learns the behavior, but "overlearns" it. "Overlearning" helps
the person overcome the slow information-processing that has resulted from
the brain injury.
As was mentioned before, one of the differences that sets communication
with survivors apart from communication in traditional psychotherapy is that
the therapist uses many other techniques in addition to strictly verbal intervention. The techniques may be verbal or visual and may consist of symbolic display, videotape, or any other modality the therapist may consider. Whatever
the technique, it is used to question, clarify, explain, exclaim, and interpret, as
well as recall.
The structure of personality posttrauma makes a difference in technique
and practice necessary; this is one way in which neuropsychotherapy differs
from traditional therapy. For example, a brain injury may result in perseveration, loss of abstract thinking, and emotional dyscontrol. The therapist
would intervene into the problems of the survivor by making repetitious comments and successively introducing more complex concepts, as well as by
taking a cognitive approach to the emotional dyscontrol (e.g., the usc of a
cognitive therapy technique such as "stop, think, and plan"). However, the
overall goal remains the same-that of aiding the patient in adjustment to
everyday living.
5.1.2 Therapeutic Contract

Research [49, 50] suggests that therapists who appear to be most helpful to
patients are those who are empathetic, genuine, and warm. However, these
qualities are not enough to ensure that relationships with patients will develop
smoothly. To forestall complications that might arise due to miscommunication between therapist and patient, a therapeutic contract should be established

10. Neuropsychotherapy

255

at the outset of treatment. The therapeutic contract is the agreement set up by


the therapist and the patient regarding the professional relationship. In essence,
the contract sets up the "ground rules" for the treatment process.
A therapeutic contract is a reciprocal arrangement. A survivor needs to be
able to depend on the therapist to meet the conditions of the contract. (This is
extremely important to someone who has difficulty trusting his or her own
apperceptive sense of the world.) Likewise, the therapist needs to know that
the patient can fulfill his or her part of the bargain. The contract should be
explicit about what the consequences will be if this does not happen.
In the contract, the therapist must spell out the underlying conditions for
treatment, such as confidentiality, the frequency, time, and location of meetings, and what action the therapist will take if the client does not keep the
appointment. This last point is a good illustration of how the therapist may
actively intervene in the life of the survivor by imposing structure. Patients
often miss their appointments because of their posttraumatic problems with
memory, orientation, and information processing. Until it is clear that the
client knows the time and place of the meeting, it becomes the therapist's
responsibility to ensure that the client attends the sessions on time. This may
mean that the therapist must completely take over these executive functions,
for either short or long periods of time. If the client forgets a session and does
not show up, the therapist needs to attempt to locate him or her and make
contact.
The usual frequency of therapy sessions should be three to five sessions (of
varying lengths) per week, depending on the level of intensity of treatment
needed. This is determined by the treating therapist, based upon the severity
of the posttraumatic sequelae. This schedule is geared to overcome two of the
primary difficulties that occur after brain injury: deficits in memory and deficits of attention. Without such frequent contact, the development of a supportive relationship involving the manifestations of empathy, warmth, and
genuineness has little chance of occurring. However, after weeks or even possibly months, some clients may not feel as if treatment is important to their
lives. This is due to their lack of awareness of their condition. It is not clear to
the survivors why they should continue to come to treatment. They do not
realize that, as a result of treatment, their lives have stabilized and they have
begun to develop relationships with people outside the family. This awareness
can be created through an ongoing log or diary of their social interactions over
time.
5.1.3 The Therapeutic Relatiollship
Survivors' emotions and thoughts are usually disorganized, and in the initial
phase this disorganization must be contained. A therapist attempts to do this
through 1) communication, 2) manifesting an empathy for and understanding
of the client's confusion, and 3) providing structure for reality. This is done
through active listening and attempting to understand the themes that are

256

III. Rehabilit,1ction Techniques

emerging in the communication patterns (i. e., a therapeutic relationship).


A therapist must carefully monitor the communication between the client
and himself or herself. Therapists may find themselves interacting with their
own theoretical system rather than with the survivor. This may happen when
spoken language is used as the functional communication system, while, in
reality, spoken language may be more for the therapist's benefit, since he or
she is more comfortable with this mode. A one-way therapeutic communication system is one of the most frequent therapeutic errors made by therapists
who work with brain-injury survivors. This style of communication, in which
the initial sessions are recreated over and over again, can be avoided by using
these preliminary sessions as an assessment period during which the therapist
attempts to understand the client's problems with 1) limitations of memory,
2) difficulties in interpersonal relationships, 3) modulation of emotions, 4) reality-testing, 5) general intelligence, and 6) cognitive integration. This information should be carefully collected over the first five or ten sessions. Once
the therapist has this information, then he or she can begin to plan ways of
avoiding one-way communication.
One of the major concerns for brain-injury survivors is that their perceptual
world has been altered [19]. The processes underlying perception (i.e., basic
sensation) have also been disrupted. The integration of the perception of the
world into a sensible place must be encouraged and taught.
In an attempt to understand the survivor, the therapist should obtain a history of the person's pretrauma personality; however, the pretrauma personality
mayor may not explain the problems that the survivor experiences posttrauma. In essence, the person who was injured was (and remains) part of a
family, as a son, or daughter, or sibling, perhaps even as a spouse or parent. In
many cases, the person was a productive member of society. The possibility of
brain injury had not entered into the narratives that survivors had had for themselves. However, after a serious head injury, people must learn to accept that
they may not be able to function within their previous roles and that their life
stories may become very different.
In the initial therapeutic interactions, survivors may frequently deny having
any problems or, conversely, may display exaggerated self-importance. The
characteristics of these themes appear to be related to the loss experienced as
well as to the apperceptive disturbances secondary to trauma. Initially, the
struggle is one of regaining control over one's body and then attempting to
make sense of a fragmented and shattered life. The issues of competency (i. e.,
legal, emotional, and intellectual) and the contrast between pretrauma functioning and posttrauma functioning appear to have a large effect on the degree of
denial or grandiosity seen. For example, difficulties in functioning in these
areas appear to affect the narratives told by and about the person, and judgments about the quality of those narratives often influence decisions about
competency. Furthermore, after a person has been judged to be competent, he
or she may feel justified in thinking grandiosly or denying problems. During
the initial phase, the grandiosity and denial themes need to be explored. The

10. Neuropsychotherapy

257

therapeutic exploration should have the objective of understanding self-esteem,


interpersonal relatedness, and the defensive structure underlying the narrative
descriptions.
Another behavioral disturbance that deserves comment is that of the suspicious stance. Suspiciousness, or even paranoid ideation, may be a result of
the pathophysiological alteration in brain functioning. The full paranoid ideation is usually secondary to altered sensory perception (i.e., visual/auditory
input), which increases vigilance and suspiciousness without necessarily involving severe cortical damage. Suspicious content needs to be explored for themes
and subject matter, which will enable the therapist to assess levels and types of
concerns. Therefore, during the initial two or three times the suspicious material is presented, it is important to explore and understand the content and
degree of the suspiciousness. If the suspiciousness is at a paranoid/delusional
level, then medication may be needed. (For more information on this topic,
see Chapters 3 and 4.)
After the information has been obtained, the themes need not be explored
further. As in schizophrenia, suspiciousness does not appear to be based in the
rational world, and therefore should not be a point to be debated and challenged
in the initial phase. Instead, the topic of communication should be changed
(rather than explored), when exploratory information is complete. This can be
accomplished by asking a question about a related or nonrelated topic.
Seizures may be a consequence of severe closed-head injury [51]. The loss of
control, the fear of embarrassment, and the possibility of death during status
epilepticus (i.e., a series of rapidly repeated, incessant seizures) are consequences that produce tremendous fear in the brain-injury survivor. Sleep disorders are also a sequela of closed-head injury, since the circadian system is
disrupted after brain injury. Researchers have found that the deeper levels of
sleep are not reached [52]. Faced with the possibility of apperceptive difficulties,
epilepsy, and sleep disorders, the survivor may feel as ifhe or she has awakened
from the injury into an intolerable, hopeless situation.
5.1.4 Techniques

Certain techniques can be used during the initial phase of rehabilitation to help
the survivor communicate. The therapist must reach an understanding of what
the survivor comprehends and must determine the best way to communicate
with him or her. One way that appears helpful to the person who has information-processing difficulties is to use the techniques that will fit with the
communication pattern available to the survivor. For instance, someone may
not be able to effectively communicate verbally but can point to written or
printed words or pictures to communicate needs, concepts, and emotions.
In the individual treatment of a survivor, the typical talking intervention
may not be extremely helpful. In such cases, some other method of conflict resolution is helpful-for example, the use of an external supervisory network to
direct the behavior of a patient. The environmental behavior-modification systems may be used in conjunction with the interpersonal-relationship building.

258

III. Rehabilitation Techniques

If the goal of therapy is adjustment to the disability, then interpersonal


communication that helps to resolve conflict and enables adaptation needs to
be established. Since memory is the key to making communication meaningful, memory for events becomes the forward and backward limits of clients'
narratives about themselves.
Memory deficits provide one of the greatest challenges to the use of individual treatment of the brain-injured. Amnesia after injury may be anterograde
(i.e., loss of memory for events occurring after the trauma) or retrograde (i.e.,
loss of memory of events that occurred before the trauma) or both. When
memory is not functioning properly, and the perceptions of the survivor are
disturbed, confabulation can occur. This can happen in two ways-when one
event is confused with another so that the actual event is not correctly understood, or when thoughts are invented in the abse: ce of memory. Perseveration
denotes a person's inability to change behavior that is occurring. This behavioral pattern is frequently seen with prefrontal lobe injury. In treatment,
the combination of cognitive deficits and memory deficits makes treatment
complicated. However, memory deficits seem most apparent when issues of
conflict arise, and the attribution of disturbed behavior to either memory
deficits or denial has continued to be a hotly debated issue.
Confrontation of denial is a specific technique that is used in all phases of
treatment. A major question to be answered is whether the resistance to acceptance of limitations is due to a conflict related to the injury or to a cognitive
deficit [53] that prevents the person from understanding the extent of posttraumatic changes. Interventions at this level become quite subtle, as the therapist attempts to maintain a supportive position while confronting the survivor
with cognitive difficulties. With this type of reaction to brain injury, a detailed
history or account of the therapy needs to be documented to present patterns
or themes to the survivor at a later juncture. Since some of the posttrauma difficulties include organization of conceptual material, attention, concentration,
memory deficits, and learning new information, it is essential to develop some
method to establish meaningful communication. For example, drawing or
writing out the issue on a pad of paper enables the survivor to see the issues at
hand as well as to hear about them; the paper can also be photocopied and
given to the client to use as a memory aid.
Although videotape, audiotape, and pictures are useful as communication
and memory aids, they provide only a certain amount of information about
the points that the therapist is attempting to make. Because memory deficits
create information-processing difficulties, the ongoing use of a therapy log
may aid in the therapeutic process. This should be separate from a patient's
day-to-day memory log, but should be used in conjunction with it.
5.2 Middle Phase of Treatment

In the middle phase of treatment, the major goal is to develop and communicate
an in-depth understanding of the patient's new identity and self-narrative. The

10. Neuropsychothcrapy

259

midphase can be thought of as focusing on three different areas: 1) the method


of communication, including the embellishment and deepening of the content and process of the communication; 2) the resistance to communication,
which includes all of the different aspects of resistance found in typical psychotherapy, as well as resistance due to and because of the deficits; 3) the therapeutic
relationship, including the actual relationship, transference, countertransference, and the working alliance. These concepts are based upon the works of
Freud (1904) [7], Federn (1926) [54], Langs (1974) [55], and Weiner (1976) [11].
5.2.1 Communication

Neuropsychotherapists draw upon many therapeutic techniques to communicate with brain-injury survivors; these techniques may include questions, clarifications, explanations, exclamations, confrontations, interpretations, and
nonverbal communications (e.g., drawings). The content and process oftherapy must always be held in the mind of the therapist. Content refers to the
subject matter of the communication. For example, the content of a communication could be that a person was injured in a certain manner and plans to go
back to college. The process of the communication refers to why the client is
talking about a certain subject (i.e., the injury), how the client states an issue
(i.e., going to college), and why the client is not talking about another subject.
In this example, the client's attending college may be a realistic goal if the injury was not a diffuse or serious one, but it may not be realistic if the injury
was severe. Communication of reality should always be balanced with hope
for the future.
Communication of reality through the use of goals, objectives, and strategies
may help change the client's behavior. These interventions provide a base for
the patient and the therapist and impart a sense of cohesive and continuous
treatment. In the treatment of brain-injury survivors, the onus of responsibility
for helping the patient achieve behavioral change rests on the shoulders of the
therapist. This shift in responsibility for action is directly due to the deficit that
has been introduced into the theoretical model. This new frame of reference
enables the therapist to have some method of understanding singular events
in an array of events-for example, a client's missing a therapy session every
Monday because the weekend disturbs the usual course of events.
Furthermore, goals, strategies, and objectives are especially helpful in the
face of strong attacks on the therapist's empathy, in which a survivor periodically displays anger at the therapeutic world that has not cured him or her. In
addition, the client may verbally or emotionally attack the therapist when his
or her homeostasis is questioned. The therapist's observing-self must be fortified against such attacks.
5.2.1.1 Content. In the midphase of treatment, the predominant issues after
catastrophic brain injury appear to be conflicts concerning annihilation, dependency or counterdependency, ideal self, pretrauma "real" self, and the reality
of existence. These realities include changes in 1) intellectual and personal func-

260

III. Rehabilitation Techniques

tioning, 2) family relatedness, and 3) work and home relationships. These conflicts need to be worked through and resolved in some meaningful manner.
The possible issues of dependency, physical changes, and loss of a future must
be explored. Comprehending the efforts of the brain injury is essential in order
for survivors to create their own self-narrative posttrauma.
These variables point to why a comprehension of who the person is, both
before and after the injury, is essential for the treatment process. It is essential
to place the trauma in its rightful place in the survivor's life. The traumatic
event can be used as an effective link to tie pretrauma and posttrauma events
and relationships together.
Sometimes, as a result of day-to-day difficulties, a survivor may develop a
global and undifferentiated form of anxiety. The survivor may dissociate and
ignore the anxiety as either danger not related to him or her or as overwhelming
anxiety that must be defended against. In an attempt to remove or dissipate the
anxiety, the survivor may immediately reach out to loved ones for comfort
and reassurance or to addictive behaviors such as drugs or alcohol to lower the
anxiety. As with most anxiety disturbances of differing origins, drugs and
alcohol are often used to mask the emotional dyscontrol reactions of anxiety,
irritation, or sadness.
The memory loss that results from closed-head injury also presents a massive pattern of problems. It is difficult for a therapist to treat a client who does
not remember something from a previous session, since the content and process of communication in the therapy session are crucial. When the therapist
states that he or she happens to have material from previous sessions, the
client often vigorously resists the injection of the information into the present.
For example, one patient who viewed a videotape of himself denied that it was
his voice or that he was the person who had been taped. Heilman et al. [53]
report that such tremendous denial usually has a pathophysiological basis.
During the midphase of treatment, the therapist may give the client a tape of
the neuropsychotherapy session. The client will be instructed to replay the
tape two or three times, in order to aid the memory process. Over time, a tape
library is built up to reinforce the client's memory of the therapeutic process.
5.2.1.2 Process. Schafer [10] has addressed the need for an appreciation of
the analytic attitude in any type of psychodynamic treatment. Because a braininjury survivor is not clear about what reality is, the many interactions used
and viewpoints expressed by helping professionals may actually result in confusing the survivor. In this context, resistance to change may not be an attempt
to block therapeutic intervention but may represent fear of annihilation.
Empathy is essential to the therapeutic relationship. This is especially true if
suspiciousness or even paranoia exists. The paranoid stance may be approached
in a manner similar to the treatment of paranoid schizophrenics. It is important
that the therapist avoid becoming enmeshed in the paranoid system.
When clients are overwhelmed by feelings and are fearful of events, the
initial attempts to modify their behavior may break down, leaving clients

10. Neuropsychotherapy

261

"frozen" in a particular stance. Psychological paralysis is a freezing of mental


activity, an inability to mobilize resources, automatic obedience, and a
blockage of affective response. These behaviors are also similar to those of a
person who has sustained diffuse prefrontal damage. If a therapist attempts to
intervene into this frozen psychological stance too quickly, personality breakdown may occur, resulting in paranoid and psychotic ideation.
5.2.1.3 Objectives and Strategies. The survivor of a severe injury may face a
number of types of personality deterioration. Typical ones are paranoid
ideation, depressed helplessness, and the loss of self through fragmentation or
the inability to differentiate between self and others.
In therapy sessions, the therapist and client first map out the objectives,
emphasizing the client's strengths and weaknesses. The client's own opinion
about these strengths and weakenesses is solicited, and the client is also asked
whether he or she accepts others' views of his or her performance. These neurological and psychosocial strengths and weaknesses are guideposts that indicate
what strategies should be employed.
After the initial therapeutic contract has been set up, the therapist uses a
multisensory approach to help develop the observing-self and experiencingself of the client. Videotape, audiotape, and therapist-client notebooks are
examples of tools that enable the therapist to assist the client in reconstructing
a self that off~ets the distortions of the disturbed perceptive process. The client
is taught to practice, rehearse, role-play, and finally use the learned material
in real life. This is accomplished through capitalizing on the functioning
brain system (e. g., visual, auditory and/or tactile) and using the cognitive,
intellectual, affective, and behavioral strengths of the person.
5.2.2 Resistance to Communication

The resistance to the communication process in ongoing treatment is welldocumented [7, 11,55]. For example, Weiner [11] pointed out: "The patient
who is resisting becomes temporarily unwilling or unable to fulfill the terms of
the treatment contract, even though he continues to want help and to believe
in the potential helpfulness of the therapist's efforts" (p. 160). He described
resistance as a client's unconscious and seemingly paradoxical effort not to
participate effectively in treatment with the therapist.
Resistances in treatment are usually classified as resistances to change and to
content of the therapy sessions (as well as both character and transference resistances). Character resistance refers to the particular style in which the individual interacts with his or her environment, and how that interaction occurs
in treatment. Weiner [11] outlined three areas to be considered in character
resistance: 1) the particular set of defenses a person uses; 2) the broad cognitive
style a person uses; 3) a particularly difficult characterological problem (e. g.,
masochism) that impairs the ability to communicate the issues.
The transference resistance is of particular importance after brain injury.
The entire process of neuropsychotherapy engenders both negative and po-

262

III. Rehabilitation Techniques

sitive feelings from the client towards the therapist. When an isolated braininjury survivor is presented with warmth, empathy, and genuineness on the
part of the therapist, such sensitivity and concern initially elicits a strong positive reaction from the survivor, and he or she begins to wish for a different
type of relationship from the therapist. To the survivor, the idealized therapist
may appear as filling the roles of "miracle worker," parent, child, mate, lover,
or special friend. Related to these roles are specific feelings of dependency,
love, and sexuality, as well as general dimensions of attachment and bonding
on both a real and idealistic level. As treatment continues, however, and the
therapist maintains, a professional stance, the survivor begins to feel frustrated
and disappointed. The survivor's positive, caring feelings may turn into
negative and angry ones. The therapist may be viewed as uncaring. Since,
obviously, the professional is neither "all good" or "all bad," the survivor and
therapist must work toward integrating a real relationship.
These extremes in emotional reactions are examples of the transference
resistance. The survivor's loss of frustration tolerance and emotional control
makes it necessary for the therapist to attempt to balance the transference with
a real relationship. Because of the survivor's fragile personality structure, a
full-blown negative transference reaction may damage the therapeutic alliance.
The integrated person can address the frustration and disappointment oflosing
an idealized partner/parent/child; however, the brain-injury survivor whose
personality structure has been altered by deficit may not accept such frustration. The resistance must be understood in relation to the survivor's posttraumatic deficits, as well as his or her personality style before the injury. In
addition, the posttrauma coping style must be integrated into the therapy and
used to affect the degree of resistance displayed in treatment.
One difference in the resistance displayed after brain injury is that the survivor may not believe or accept clearly organized and <;upportive information
about his or her resistance to change. It is commonplace for the therapist to see
or detect emotions through the survivor's actions, gestures, and tone, while at
the same time the survivor is denying the experience of the emotions. During
this period, the survivor may report the experience of being uncomfortable
with the therapist's line of questioning or clarifications. This discrepancy
between the therapist's experience and the survivor's appears to be a differentiation in experience of the event. The survivor no longer experiences the
world or observes himself or herself in the same way as before the injury.
Emotions are not processed-or rather, not experienced-in the same fashion.
The apperceptive world of the survivor, as understood through the experiencing-self and the observing-self, has become damaged.
Often, as a result of brain injury, the survivor's ability to use emotions as
cues to understand personal interactions is also damaged. Anxiety appears
to be triggered quickly and unexpectedly. For example, Piotrowski [25] has
demonstrated the catastrophi<; anxiety reactions of people with organic brain
syndrome. The resulting damage to interpersonal skills and intimate relationships is enormous.

10. Neuropsychotherapy

263

An inadequate apperceptive system and failure in interactive communication


may cause the survivor to feel as ifhe or she is "going crazy," which may lead
to violent or hostile behavior. No matter whether the behavior is violent, or
psychotic, or some variation of these, the behavior must be carefully evaluated
in an attempt to understand its underlying causes. The undifferentiated affects may appear to be vague and uncontrolled emotional reactions. However,
these affective dyscontrol states are usually triggered when the person's usual
methods for controlling overwhelming affect are not functioning adequately
(as a result of the injury). Therefore, in tandem with exploring resistance and
content, the therapist must help the survivor strengthen his or her observingself through rebuilding adequate defense strategies. Fear of feelings, lack of an
adequate affective response, and inability to process information indicate why
emotion may be experienced in such complex ways.
The verbal message that the head-injury survivor reports to others is that he
or she is fine, and that the problems experienced are a result of the external
environment. The question of whether "someone else is to blame" may be
accurate, but it may also reflect the brain-injury survivor's projection of rage,
anger, and feelings of inadequacy onto the environment. However, this projection of feelings is usually accompanied by a feeling that the environment is
going to hurt or punish the person in some fashion.
The concept of "acting-out" as a primary maneuver of resistance must be redefined after brain injury. The alteration in cognitive status must be considered
when the therapist is presented with behavior from the client which is disruptive
or hostile. The act of forgetting or memory loss must be scrutinized carefully.
For example, Sandel and her colleagues reported on a patient who displayed
amnesias secondary to a dissociative disorder and brain injury (Sandel, M. E.,
Weiss, B. and Ivker, B., unpublished data, 1987). Further examination indicated that the patient appeared to develop a splitting of personality, similar to
the condition of multiple personalities.
5.2.3 The Therapeutic Relatiollship: Transference

From the view of the survivor, the therapeutic relationship can be understood
as a combination of a transferential relationship, a real relationship, and a working alliance [11]. The real relationship is what is reasonable and to be expected,
and the working alliance is based on the nature of the therapeutic work. After
brain injury, these facets of the therapeutic relationship are not as clear-cut as
they would be in traditional psychotherapy.
The brain-injury survivor's initial transf$:Tential reaction to the therapist is
that of a passive yet hopeful patient waiting for a cure. If nothing is done (i.e.,
if the therapist takes no action), the survivor will just wait for the problems
to go away. The longer this passive stance is allowed to exist, the more difficulty the survivor has in confronting the posttraumatic sequelae that cause
the problem.
When the survivor feels that he or she is not getting well quickly, anger,
denial, and mistrust are directed toward the therapist. The survivor feels as if

264

III. Rehabilitation Techniques

he or she has been taken advantage of (at least financially) and the therapist is
discredited as having nothing to offer. This movement from a positive transferential relationship to a negative one has been seen as a natural progression in
psychodynamic treatment [10]. Since frustration tolerance is lower after brain
injury, however, the therapist cannot allow the negative transference to build
to too high a degree. If this occurs, the patient may drop out of treatment directly or lose complete confidence in the process and stop attending.
The survivor's feelings of dependency and vulnerability after brain injury
continue to be a major transferential issue. Coming so soon after the extreme
dependency and learned helplessness of the hospitalization, the process of
neuropsychotherapy may create conflict. The conflict may be manifested by
1) a hostile relationship, 2) a dependent relationship, or 3) a hostile-dependent
relationship with his or her therapist. The hostile-dependent relationship that
the survivor may develop with the family and primary caregivers can be viewed
as the survivor's attempt to set up a relationship that he or she has learned in
the hospital and rehabilitation area, and which has apparently succeeded in the
acute and postacute hospital phase. If the therapist accepts the role of the authority who will "cure" the patient, then the beginning of the hostile-dependent
relationship has been set in motion. Since the therapist cannot cure brain injury, the trauma must be explored with the survivor to evaluate what the effects of the brain injury have been, are, and will be. The effect on the family
must also be explored. Often, the conflicts that occur in the family after the
traumatic event are transferred into the therapeutic, dyadic relationship. The
therapist must be aware that he or she automatically becomes a part of the survivor's system of interactions. This system has been created to "fix" the patient (i.e., restore the patient to the way he or she was before). When the
family becomes disappointed (because the therapist cannot cure the survivor),
they may seek a new therapist or rehabilitation program. The therapist needs
to explain to the family the possible reasons for their dissatisfaction, even
though the explanation may not alter the outcome.
Consciously or unconsciously, the survivor resists empathy and then initially appears to experience it as a violent or hostile invasion of privacy, or as too
intimate, or as some combination of these. In addition, the survivor may not
be able to sense the therapist's empathy. The type of reaction varies, depending upon the primary conflict that has been experienced, the client's pretrauma
personality, and the therapist's characteristics. The theme that appears to surface is that the survivor feels vulnerable in the face of a therapeutic relationship
and therefore appears to resist the therapeutic relationship. When this occurs,
the therapist is faced with a number of options, yet if the resistance is strong,
then the therapist must "go with" the resistance to avoid destroying the therapeutic alliance. This may mean going into a supportive mode of interaction
until the client's personality structure has been reaffirmed.
5.2.3.1 Countertran~feren(e. From the viewpoint of the therapist, the therapeutic relationship can be understood as a combination of countertransference

10. Neuropsychotherapy

265

plus a real relationship, as well as the working alliance of the therapeutic relationship [11]. Countertransference is the active projection of the therapist's
feelings and thoughts (arising from the therapist's past) towards the patient.
These are irrational and inappropriate reactions to the behavior of a patient.
The awareness of one's own countertransference material may be therapeutically useful in guiding treatment.
The therapist must recognize the profound impact that the catastrophic
injury has had on the survivor. Therefore, the empathic therapist must understand and accept the patient's emotional conflict without attempting to prematurely confront the patient about the denial of the impact of his or her injury,
as well as the emotions aroused by the event and consequences. In this way,
the therapist contains as much of the conflict as possible for the client, similar
to Winnicott's [56] "good enough mothering" concept.
Countertransference has its origins in the structure of the therapist's needs
and personality, yet the countertransference is usually activated by the survivor's condition, behavior, and experiencing-self. Most therapists have certain emotional reactions towards brain-injury survivors-indeed, these are the
usual patterns of emotional reaction to any person who has experienced a catastrophic injury that has long-term consequences. A therapist may emotionally
react to a survivor's catastrophic injury with feelings of depression, hopelessness, and a general sense of loss of the ideal. An unconscious form of response
to this is to attempt to transform the patient from the person that he or she is
(after the trauma) into what he or she was like before the accident. Because
of this type of natural reaction, the therapist may find it difficult to accept the
survivor as he or she is, rather than as some idealized, non-brain-injured
person.
The therapist should monitor his or her therapeutic ability as an index of
possible countertransference issues. The active therapist must be careful not to
avoid or act out a projected depression of the survivor. Therapists frequently
speak of their feelings of hopelessness and depression when working with the
brain-injured. The therapists become depressed because the survivors are
"brain-damaged" and therefore "cannot change." What this really means is
that the therapists have an unrealistic expectation about the amount of change
that is possible. In fact, they may consciously or unconsciously believe that the
survivors may be able to return to pre morbid levels of functioning.
In addition, professionals sometimes state that the survivor has cognitive
deficits and that the survivor's problems are merely cognitive ones. Therefore,
they claim, the answer is to administer cognitive remediation and therefore fix
the problem. Another version of this scenario is to help the survivor make a
transition back into his or her community and take part in activities of daily
living (ADL). Individual therapy is viewed as superfluous by the therapist who
delegates intervention to cognitive remediation and ADL training or pharmacology. These general reactions appear to be a method by which therapists will
not have to grapple with the emotional life of the survivor.

266

III. Rehabilitation Techniques

Other typical general reactions seen in therapist behavior include 1) shortening or not holding the regularly scheduled sessions; 2) fearing that the survivor
will go out of control and hurt himself or herself or others; and 3) feeling hopeless about making any real difference and believing that the deficits are somehow the therapist's fault. For example, the client may blame the therapist for
keeping him or her from going back to work or to school (i.e., from becoming
a whole person).
Many therapists do not continue to work directly with brain-injury survivors for any length of time. The reasons appear to be not only the countertransferential feelings of sadness, but also sorrow about what has happened
and frustration about the amount of change observed. Searles [57] wrote
about the same type of countertransference and real frustrations encountered
in treating schizophrenic patients.
The therapist experiences a fragmentation of his or her therapeutic framework toward the survivor and, as a result, may have difficulty with his or her
empathy toward the survivor. Since empathy is a major tool for the therapist,
any attacks on empathy by the survivor's unconscious resistance to change and
the resulting emotional reaction by the therapist should be explored. Schafer
[10] has described some of the difficulties and complexities that arc associated
with therapeutic empathy.
Empathy appears to affect the survivor's structure of defense. If denial of the
injury (or even denial of hope) was a major part of the survivor's defense structure, then empathy can produce painful emotions. Defensively, the survivor
may not appear to want hope, but rather wants to be left alone. Empathy is
defended against because it is overstimulating; it stirs emotions that feel
overwhelming and painful, because it may offer hope.
Unfortunately, because of the brain injury, the survivor's progress may be
quite minimal and slow. The therapist may develop blind spots in the treatment process, especially considering minimal gains. Blind spots are those
points at which the therapist loses insight into the pattern or process of treatment-for instance, if the therapist begins to view certain behaviors of the
survivor as oppositional (rather than as resistance or cognitive defects) and
becomes angry. These signals enable the therapist to note that he or she may
have lost track of the direction of treatment-for example, when the survivor
acts in ways that the therapist does not like, and the therapist begins to dismiss
the survivor as "too impaired" or "too angry" (or "too violent") for treatment.
The therapist needs to explore why the behavior of patients is irritating to such
a degree.
5.3 Final Phase

By the final phase of treatment, the therapist should have helped the survivor
learn to use different information-processing aids, support personnel, and
community services to better enjoy his or her life. In addition, because a major
goal of treatment has been to help the client come to an acceptable resolution

10. Neuropsychotherapy

267

of the experienced loss, another part of the final phase of treatment is to again
work through this resolution (to make certain it has been achieved).
During treatment, the idealized pretrauma self of the brain-injured individual has been allowed to die. This includes the initial denial and isolation of
emotion, the different facets of acceptance, and movement toward the future.
The termination phase occurs after the survivor, therapist, and therapeutic
team agree that the survivor has reached the highest level of functioning currently possible. Other factors that may enter into this decision are reports from
the survivor's family and significant others.
At this point, the survivor has come to accept some, if not most, of the
deficits and the many profound changes that have resulted from the trauma.
Part of the primary goal of the treatment has been to help the survivor integrate
this acceptance of reality into his or her life in some meaningful way.
In practice, the therapist is hard-pressed to predict the final outcome for a
survivor. The final phase of therapy frequently consists of a long period of
ongoing support for the survivor, which, in fact, may continue as long as the
person lives [12].
6 CONCLUSION

Severe brain injury with subsequent emotional and behavioral disturbances


does not preclude a survivor from taking part in individual treatment. However, the individual therapy must take into account the survivor's neurological
status and his or her posttraumatic personality structure.
In working with this population, however, therapists need to revise their
usual strategies, tactics, and goals. For instance, verbal therapy alone may not
be effective because of the survivors' tendency to misperceive language and abstract strategies, to unrealistically neglect social and emotional priorities, and
to be deficient in memory, attention, concentration, and information-scanning
and information-processing abilities. Throughout all phases of neuropsychotherapy, the therapist must also be much more active than he or she would be
in traditional psychotherapy with non-brain-injured persons.
REFERENCES
1. Ball, J. (1988). Psychotherapy with head injured patients. Mcd. Psychothcr. 1, 15-22.
2. Broca, P. (1864). Deux cas d'aphemie tramatique, produite par des lesions de la troisieme
circonvolutiou frontale gauche. Bull. Soc. Chir. 5, 51-54.
3. Wernicke, C. (1874). Del' aphasichc Symptomcl1complex. Frank & Weigert, Breslau.
4. Charcot, J.M. (1887). Lectures 011 CC1/tral Nervous System Diseases, Made at the Salpitriere (Paris
Hospital for the Needy), Vol. III. Lecrosnier et Bobe, Libraires-Editeurs, Paris.
5. Janet, P. (1888). Les acts inconscients et ]a memoirc. Rev. Phil. 13, 238.
6. Freud, S. (1953). The interpretation of dreams. In The Stalldard Edition of the Complete PsycholoJ<ical Works of S(~mulld Freud, VoIs. 4 and 5, Strachey, J., cd. and trans., Hogarth Press,
London. (original work published 1900)
7. Freud, S. (1953). Freud's psycho-analytic procedure. In The Stalldard Editioll of the Complete
Psychological Works of Sigmulld Freud, Vol. 7, Strachey, J., ed. and trans., Hogarth Press,
London. (original work published 1904)

268

III. Rehabilitation Techniques

8. Jaynes, J. (1976). The Origin ofCol/sciousness in the Breakdown of the Bicameral Mind. Houghton
Miffiin, Boston.
9. Goldstein, K. (1942). Aftereffects of Brain I/~iuries. Grune & Stratton, New York.
10. Schafer, R. (1983). The Analytic Attitude. Basic Books, New York.
11. Weiner, I.B. (1975). Pril/ciples ~f Psychotherapy. John 'Wiley & Sons, New York.
12. Luria, A.R. (1966). H(f?her Cortical Functions in Man, 2nd ed. Basic Books, New York.
13. Diller, L. (1985). Neuropsychological rehabilitation. In Neuropsychological Rehabilitation, Meir,
M., Benton A., and Diller, L., eds., Guilford, New York, pp. 3-17.
14. Prigatano, G.P. (1986). NRliopsychological RehabilitatiOlI After Brain Injury, Johns Hopkins
University Press, Baltimore.
15. Lezak, M.D. (1982). The problem of assessing executive functions. Int. J. Psychol. 17,
281-297.
16. Benton, A. (1979). Behavior consequences of closed head injury. In Central Nervous System
Trauma Research Status Report, National Institute of Neurological and Communicative Disorders and Stroke, Bethesda, MD.
17. Brooks, N., ed. (1984). Closed Head b!iury: Psychological, Social and Family COl/sequences.
Oxford University Press, Oxford, England.
18. Goethe, K.E. and Levin, H.S. (1984). Behavior manifestations during the early and long-term
stages of recovery after closed head injury. Psychiatr. Ann. 14, 540-546.
19. Ellis, D.W. and Zahn, B.S. (1985). Psychological functions after severe closed head injury.
J. Pers. Assess. 49, 125-128.
20. Krystal, H. (1984). Psychoanalytic views on human emotional damages. In Post-Traumatic
Stress Disorder: Psychological and Biological Sequelae, Van Der Kolk, B.A., ed., APA Press,
Washington, DC, pp. 1-28.
21. Van Der Kolk, B.A., ed. (1984). Post- Traumatic Stress Disorder: Psychological and Biological
Sequelae. APA Press, Washington, DC.
22. Bartemeir, L.H., Kubie, L.S., Menninger, K.A. et al. (1946). Combat exhaustion. J. Nerv.
Ment. Dis. 104, 358-389.
23. Grinker, R.R. and Spiegel, J.P. (1945). Men Under Stress, Blackiston, New York.
24. Klein, M. (1946). Notes on some schizoid mechanisms. Int. J. Psycho anal. 27, 99-110.
25. Piotrowski, Z.A. (1974). Perceptionalwlysis, 3rd ed. Ex Libris, Philadelphia.
26. Seligman, M.E.P. (1975). Helplessl/ess: all Depressioll, Developmellt and Death. W.H. Freeman
& Co., San Francisco.
27. Levin, H.S., Grafman, J. and Eisenberg, H.M. (1987). Neurobehavioral Recovery from Head
/tljury. Oxfurd University Press, New York.
28. Ellis, D.W., Gehman, W.S. and Katzenmeyer, W.G. (1980). The boundary organizatiun of
self-concept across the 13 through 18 year age span. Educ. Psychol. Meas. 40, 9-18.
29. Ellis, D.W. and Davis, L.T. (1982). The development of self-concept boundaries across the
adolescent years. Adolescence 17, 695-710.
30. Lezak, M.D. (1978). Living with the characterologically altered brain injured patient. J. Clin.
Psychiatry 39, 592-598.
31. Winson, J. (1985). Brain and Psyche: The Biology of the Unconscious. Anchor Press/Doubleday,
Garden City, NY.
32. Libet, B. (1985). Unconscious cerebral initiative and the role of conscious will in voluntary
action. Behav. Brain Sci. 8, 529-566.
33. Bear, D.M. (1983). Hemispheric specialization and the neurology of emotion. Arch. Neurol.
40, 195-202.
34. Jacobson, E. (1964). The Self and the Object World. International Universities Press, New
York.
35. Bruner, J. (1986). Actual Minds, Possible Worlds. Harvard University Press, Cambridge, MA.
36. Lezak, M.D. (1983). Neuropsychological Assessment, 2nd ed. Oxford University Press, New
York.
37. Heilman, K.M., Watson, R.T. and Valenstein, E. (1985). Neglect and related disorders. In
Clinical Neuropsychology, 2nd ed., Heilman K.M. and E. Valenstein, E., eds., Oxford University Press, New York, pp. 243-293.
38. Rappaport, D., Gill, M.M. and Schaeffer, R. (1975). Diagnostic Psychological Testing, rev. ed.,
International Universities Press, New York.
39. Weiner, I.B. (1966). Psychodiagnosis ill Schizophrenia. John Wiley & Sons, New York.

10. Neuropsychotherapy

269

40. Freud, A. (1946). The Ego and the Mechanisms of Defence, International Universities Press, New
York. (original work published in 1936)
41. Holt, R.R. (1956). Gauging primary and secondary Rorschach responses. J. Proj. Tech. 20,
14-25.
42. Goldberg, E. and Bilder, R.M., Jr. (1987). The frontal lobes and hierarchical organization of
cognitive contro!' In The Frontal Lobes Revisited, Perecman, E., ed., IRBN Press, New York,
pp. 159-187.
43. Lezak, M.D. (1982). The problem of assessing executive functions. Int. J. Psycho!. 17,
281-297.
44. Mackinnon, R.A. and Yudofsky, S.c. (1986). Psychiatric Evaluation in Clinical Practice. J.B.
Lippincott, New York.
45. Boller, F. and Frank, E., eds. (1982). Sexual Dysjimctiorl in Neurological Disorders: Diagnosis,
Mana/tement and Rehabilitation. Raven Press, New York.
46. Blackerby, W.F. (1987, December). Sexual dysfunction and adjustment in head injury. In
Head Injury Frontiers: Research, Rehabilitation, Re-entry. Symposium conducted at the Sixth
Annual Symposium of the National Head Injury Foundation, San Diego, CA.
47. Lally, K.M., Collins, D.B. and Collins, V.V. (1987, December). Working together toward
sexual compatibility, post head injury: A couple's perspective. In Head In;ury Frontiers: Research, Rehabilitation, Re-entry. Symposium conducted at the Sixth Annual Symposium of the
National Head Injury Foundation, San Diego, CA.
48. Erikson, E.H. (1968). Idel/tily: Youth and Crisis. W.W. Norton, New York.
49. Truax, C.B. and Carkhuff, R.R. (1967). Toward Effective COlll1selill/t and Psychotherapy, Aldine
Publishing, Chicago.
50. Garfield, S.L. (1981). Psychotherapy, a 40-year appraisa!. Am. Psycho!. 36, 174-183.
51. Guidice, M.A. and Berchov, R.C. (1987). Post-traumatic epilepsy following head injury.
Brain Injury 1, 61-64.
52. Riley, T.L., ed. (1986). Clinical Aspects & Sleep Disturbance. Butterworths, Boston.
53. Heilman, K.M., Bowers, D. and Valenstcin, E. (1985). Emotional disorders associated with
neurological diseases. In Clinical Neuropsycholo/ty, 2nd ed., Heilman, K.M. and Valenstein, E.,
eds., Oxford University Press, New York, pp. 377-402.
54. Federn, P. (1952). Ego Psycholo/ty and the Psychoses. Basic Books, New York.
55. Langs, R. (1973-1974). The Technique of Psychoanalytic Psychotherapy, Vols. 1 and 2. Jason
Aronson, New York.
56. Winnicott, D. W. (1953). Transitional objects and transitional phenomena. Int. J. Psychoana!.
24, 89-97.
57. Searles, H. (1960). The Nonhuman Environl/1ent. International Universities Press, New York.

11. STRUCTURED GROUP TREATMENT


FOR BRAIN-INJURY SURVIVORS

YEHUDA BEN-YISHAY AND PHYLLIS LAKIN

1 INTRODUCTION

This chapter articulates the conceptual underpinnings, rationale, and principal


clinical and methodological features of a systematic, structured group therapy
approach that was especially developed for survivors of traumatic brain injury.
This group therapy methodology has been successfully applied, for well
over a decade, in Israel [1] and by the Head Trauma Program of the New York
University Medical Center, Rusk Institute of Rehabilitation Medicine [2-8],
as an integral part of a multifaceted and holistic program of neuropsychological
rehabilitation for traumatically brain-injured young adults.
2 CLINICAL FOUNDATIONS

2.1 Organic Bases of Emotional Disturbances Following Brain Injuries

It has long been recognized that emotional difficulties, personality changes,


and various types of psychiatric complications are frequently part of the posttraumatic sequelae in head injuries. Indeed, at times, such problems may be the
only consequences of brain injury [9-10].
Lishman [10] pointed out that a "rigid distinction between psychogenic and
physiogenic factors cannot of course be maintained, since both may operate
together and each may aggravate the other" (p. 373). He further stated:
It cannot be doubted that certain psychiatric symptoms derive in a more or less direct
manner from organic lesions of the brain. This is firmly established where cognitive
271

J,7Z

III. Rehabilitation Techniques

functions are concerned, and any intellectual impairment after head injury is readily
attributed to neuronal damage. (p. 373)

Essentially the same point was made by Brosin [11], who pointed out that
in the case of persons with traumatic brain injury and subsequent personality
disorders,
The assumption underlying ... is that the physical injury to the brain substance causes
a chronic derangement resulting in altered behavior, usually with symptoms which can
be identified as directly related to this organic damage, be it anatomic, biochemical, or
electronic, and not due primarily to neurosis. (p. 1182)

But in spite of the above, as Boll [12] has pointed out, many of the psychiatric symptoms exhibited by brain-injured patients are frequently passed off
as "functional" in nature, and the patient is referred for mental-health counseling to help deal with what are presumed to be purely functional problems.
Several recent studies by Rimel et al. [13], Jane et al. [14], and Levine (H.
Levine, personal communication, 1982) suggest that there is a definite underlying organic component to the behavioral, cognitive, and emotional sequelae
of even minor head injuries. Such studies appear to be fully supported by
Oppenheimer's [15] postmortem findings in humans and Jane and Rimel's [16]
findings in monkeys, which indicate that even minor head injuries resulted in
multiple microscopic lesions and degeneration of white brain matter, despite
the absence of clinical neurological or neurometric findings while the humans
or monkeys were alive.
That even minor head injuries can result in the physiological disruption of
brain function, as well as in morphological alterations, has significant implications for differential diagnosis and therapy. It is now necessary to assume
that any emotional and personality changes following a traumatic head injury
involve an organic component. The organic component of the emotional
and/or personality disturbances following a brain injury achieves unique
behavioral-symptomatic expression within the following contexts: 1) the
mental constitution of the individual, 2) factors of premorbid personality, 3)
situational conflicts that are unrelated to the head injury, 4) the patient's reactions to the fact of being injured, and 5) the particular pattern of the patient's
intellectual impairments.
2.2 The Nature of the Association Between Lower Level and
Higher Level Impairments of Mental Functions

Goldstein [17], commenting on the relationships between lower and higher


level impairments in the mental functions of brain-injury survivors, stated:
There is no doubt that the process in the higher and lower levels of the brain are, to
some extent, associated; both belong to the mental apparatus. [When one examines the
patient's higher level intellectual functions,l it is not ... appropriate ... to omit the

11. Structured Group Treatment

273

symptoms belonging to defects in the lower level functions [the so-called motor and
sensory instrumentalities which are necessary for the higher level mental functions].
When the brain injury is severe enough to cause an impairment of the abstract attitude,
neither the automatic nor the emotional reactions of patients with impairment of the
abstract attitude appear[ s] ... normal. (p. 776)

Considering the various implications-for diagnosis as well as for rehabilitation-of the interdependence between the functioning of the newer and
older components of the nervous system, Moore [18] points out two consequences of brain injury that echo Goldstein's point of view:
The neosystems are extremely dependent upon the older components of the nervous
system for setting the background tone or postural adjustment, timing and coordination of stereotyped patterns of movement. Likewise, the reverse of this is that the
older systems, in order to carry out these functions, are dependent upon certain inputs
from the neocortex and neocerebellum for inhibiting "unnecessary" activity in the
older subcortical systems, including areas in the brain stem and spinal cord .... This
enables the phylogenetically newest components of the neosystems to carry out higher
cortical functions without having to expend excess energy regulating and coordinating
the lower systems. (p. 38)

Hence, brain damage alters both the functions that are directly affected by the
lesion and those functions that are determined, in part, by areas of the brain
not directly affected by the lesion. The net effect is that the patient's lower
level "automatic" functions, cognitive and other mental functions, and
emotional responses are all subject to alteration to a greater or lesser extent,
following a brain injury, At least some aspects of the emotional and personality disturbances observed in brain-injury survivors are directly attributable
to the organic deficits. They are not-and should not be viewed as beingmerely functional (i.e., "neurotic") responses to the brain injury.
2.3 Empirical Evidence for the Association Between Lower Level and
Higher Level Cognitive Functions and Interpersonal functions

A five-year comprehensive study by Ben-Yishay et al. [19] on the effects


of a systematic program of neuropsychological rehabilitation of chronic traumatically brain-injured young adults provides strong support for validity of
the statements in Section 2.2, concerning the interdependence of the lower
level ("automatic") functions, the cognitive and daily life functions, and the
emotional (intrapersonal and interpersonal) functions in persons who have
sustained moderate to severe traumatic brain injury.
Before beginning remedial treatment, subjects were administered an extensive battery of tests to measure baselines of basic psychomotor and attentional
functions, lower and higher level cognitive functions, intactness of academic
skills, and intactness in 19 areas of functional, daily life skills. In addition, five
situational intrapersonal and interpersonal functions were measured in vivo

Basic psychomotor measures


Visual Reaction Time
Finger Tapping
NYU, IRM, Basic attelltiol1l1IfclSUreS
Visual Discrimination (VDC)
Time Estimates (TE)
Rhythm Synchrony (RSC)
Fil1ger dexterity measures
Purdue R hand
Purduc L hand
Purdue Assem bl y
NYU, IRM Threading
NYU, IRM Nuts/Bolts
Mmloryjill1ctiol1 measures
Sentence Repetition
WMS, Logical Content
Benton VRT, Recall
COl1strtlctiollal praxis measures
W AIS, Object Assembly
W AIS, Block Design
Visual il1{ol'lltatiol1 pl'Occssillg measures
NYU, lRM Figure Recognition
NYU, IRM Picture Description
NYU, IRM Navigation
NYU, lRM Spatial Relations

Other domains

.32
(.29)

(.26)
.32

Self-esteem

.42
.31

Self-appraisal

Intrapersonal

.38

.53
.39
.46
.30
.43

.38
(.26)
.41
.37
.33
.46

(.27)

.52
.35
.49

.53

.4()

(.26)

.38
.34

interaction

.47

.38

.36

Empathy

Small
group

(.28)

(.29)

Social
cooperation
and leadership

Interpersonal

Table 11-1. Correlations" between intrapersonal and interpersonal measures and lower and higher level cognitive and daily life functions b

AO

.33
.52

.59

A8

(.29)

.33
.28

(.26)

.32

.35
.59
.54

A6
A3
A2

.32

Al
A9

.31
.38
(.29)

.30
.33

.37

.30

A6

.34

A3

.52
.50
.37
(.29)
.30
(.28)

A9

A3

Al

.34
.37
(.26)

.53

.60
.63
.38

A2

(.27)

(.26)
.31

.35
.34

(.26)

A5

.32

(.29)
.65

A4

A7

.32

.39
.37
.38

A9
A9

.37

.37
.52
.64

A9

.54

.58

.37

Al

A5

.56
.53
.61

.36
(.29)
.32
.33

(.26)
.33
(.28)

.36
.42
(.29)

.31
.52
.39
.39

A4

.38
.33

A7

(.27)

A4

.39

A4

.33
.36

(.26)

A5

.36
.32

., r" crit. p < .05 = .30; P < .01 = .39; ( )a, r, approaches signiticance.
h 11 = 40; patients had severe traumatic brain injuries, measures were obtained before renledial intervention.

NYU, IRM.{llllcliollal competellce illdex


Self care
Chores
Initiative/activity level
Orient, familiar environment
Orient, unfamiliar environment
Memory
Adaptive skills
Appropriateness, signif. others
Appropriateness, strangers
Organizational ability, routine
Organizational ability, new
Auto-regulation of affect
Safe expression of emotions
Intimacy
Sexuality
A wareness of disability
Acceptance of disability
Cooperation
Academic skills
NYU, IRM Reading of Directory
MATSpclling
MA T Vocabulary
MA T Comprehension
NYU, IRM Arithmetic
Hili/lei' level illtelleclllal aptitude
WAIS, VIQ
WAIS, PIQ
Categorical thinking
Flexibility of verbal thinking
Flexibility of nonverbal thinking
Generating "telegrams"
Selecting telegrams

276

III. Rehabilitation Techniques

(i.e., in small groups). Table 11-1 presents the results of correlations between
each of the five measures of intrapersonal and interpersonal functions and the
other domains.
As can be seen from the pattern of correlations in Table 11-1, the five intrapersonal and interpersonal measures interrelated to various degrees with at
least some of the variables of everyone of the domains of functioning sample.
The number of significant correlations was modest and accounted for only 9%
to 40% of the variance, leaving much of the variance to be accounted for by
other factors. These findings appear to confirm the hypothesis that, in traumatically brain-injured individuals, defective functioning in the interpersonal
areas is at least partially determined by the deficiencies in other areas of the
patient's mental apparatus, including the lower level psychomotor functions.
Since, however, only a relatively small amount of the variance was accounted
for by these correlations, the findings also appear to support the notion that the
specific clinical picture, in any given patient, will vary greatly according to the
other mediating influences, including the premorbid mental constitution,
temperamental aspects, personality factors, special family or situational problems at the time of the injury, and the patient's subjective responses to being
injured.
2.4 Implications for Psychotherapeutic Interventions

It is generally accepted [20, 21] that for a person to benefit from any type of
insight-producing psychotherapy, it is necessary that the person possess 1) a
degree of awareness (i.e., acknowledgement of the existence of problems), 2) a
general readiness to seek and accept help from another person, 3) the ability to
carry out resolutions to change undesirable or maladaptive behaviors, and
4) intact "ego-functions."
There are probably as many definitions of ego functions as there are extant
theories. For the purposes of this discussion, we will confine ourselves to an
examination of Goldstein's [17, 22, 23] notion of the abstract-concrete attitude,
since it is the most relevant to psychotherapeutic considerations in people with
brain injury, and to Bellak's [24] definition of ego functions.
Goldstein [17] has identified one of the core (i.e., generic) deficits in head
injury as being the impairment of the ability to assume an abstract mode of
thinking. He described the abstract attitude as consisting of ten interrelated
ability-components. These consist of the ability to 1) assume a definite mental
set; 2) give an account to oneself for acts and for thoughts; 3) shift reflectively
(i.e., at will) from one aspect of a situation to another; 4) keep in mind simultaneously various aspects of a task or of any presentation; 5) grasp the essentials
of a given whole (i. e., mentally break the whole into separate isolated pieces)
and synthesize them back into the whole; 6) abstract common properties of
different things; 7) mentally perform concepts, think in symbols, and be able
to understand them; 8) voluntarily evoke in one's mind previous experiences
(e.g., images); 9) be able to assume the "merely possible" (i.e., be capable of

11. Structured Group Treatment

277

imagining "as if" something has occurred); and 10) be able to detach oneself
from one's other experiences of the outer world or from one's inner experiences.
In considering the question of whether brain-injured individuals' emotional
and personality disturbances lend themselves to treatment by conventional
psychotherapeutic methods, Goldstein [22] called attention to the need to
differentiate symptoms that are due to "disturbance of inborn or learned
patterns" from the "catastrophic conditions" that produce "the expressions of
the protective mechanisms" of the organism (p. 65). Goldstein cautioned that
because of the similarity between the "defense mechanisms" against anxiety in
non-brain-injured neurotic patients and the organismic "protective mechanisms" of the brain-injured patients, it is easy to confuse the two and "consider
them as being determined by the same underlying mechanism. " This is not so,
Goldstein asserted, since the term" 'defense mechanisms' [in neurotic patients]
refers to a more voluntary process in the mind of the neurotic person, whereas
the protective mechanism in the brain-impaired person occurs passively
through organismic adjustment" (p. 65).
Bellak [24], a psychoanalytic therapist and theoretician, recognized the
problems of doing conventional psychotherapy with people who have organic
dysfunctions underlying their apparently psychiatric symptoms. He defined
"ego functions" as consisting of 12 components, which are identified and
briefly defined in Table 11-2.
A closer examination of Goldstein's [17] listing of the components of the
abstract attitude and Bellak's [24] definition of ego functions reveals that these
Table 11-2. BelIak's list of components of ego functions
Component
1. Reality testing
2. Judgment
3. Sense of reality

4. Regulation and
control of drives,
emotions, and
impulses
5. Object relations
6. Thought processes
7. Adaptive regression
in the service of the
ego
8. Defensive
functioning
9. Stimulus barrier
10. Autonomous
functioning
11. Synthetic-integrative
12. Mastery-competence

Brief definition
Refers to ability to make proper distinctions between inner and
outer experiences, stimuli
Refers to ability to express emotion appropriately
Refers to self-esteem, self-identity
Refers to effectiveness of ability to delay or express impulses,
emotions, drives
Refers to degree and kind of relatedness to/with others
Refers to cognitive functions such as memory, attention,
reasoning
Refers to temporary states of diminution of cognitive acuity
(until person recovers from stress)
Refers to strength and kinds of defenses exhibited to ward off
anxiety
Refers to sensory perceptual thresholds
Refers to degree of independence and autonomy exhibited in
personal relations
Refers to higher lever abstract (e.g., "executive") functions
Refers to the degree of discrepancy between one's actual mastery
of tasks ofliving and the subjective feelings of mastery

Figure 11-1. Components of ego-fun ctions (based on Bellak [24]),

~ ,

I
iF

it / 1:

( DEFE N SES )

AWARENESS

IMPULSE REGULATION

ADAPTIVEN ESS

- THOUGHT PROCESSES

AUTONOMY

RELATEDNESS

EMOTIVE PROCESSES

.E '
c

:l

(1)
()

,..,

:l

0'

~~

"

(1)

0Cl

11. Structured Group Treatment

279

two conceptions are identical in some respects (though differing in the use of
semantics) and complementary in other respects. It is clear from both that the
concept of intact ego functions implies the intactness of several capacity levels
in an individual's mental functions. These capacity levels can be described by a
highly simplified metaphor that is presented in Figure 11-1. The concept of
intactness of ego function assumes a degree of basic intactness of at least some
or all of the component functions identified in Figure 11-1.
To date, no one has presented an objective and reliable method of identifying which brain-injured individual is sufficiently intact to permit conventional
forms of psychotherapy and which one is ego-impaired. Yet the implication is
clear: Brain-injured individuals who do have impaired ego functions are not
good candidates for conventional forms of psychotherapy.
3 A MODEL FOR GROUP THERAPY FOR BRAIN-INJURY SURVIVORS

A number of clinical investigators and practitioners have called for development of techniques especially suited to the needs of brain-injury survivors
[1,2,25-27]. The most explicit analysis was presented by Ben-Yishay et aI.,
[7] in a parsimonious model distinguishing the essential properties of conventional psychotherapies from the one proposed for use with brain-injured individuals. We present this model here with a brief explanation of its
underlying assumptions and rationale.

THERAPEUTIC
INTERVENTION
SETTING

~-~.,
"CHANGE ATTITUDE". __
I

COGNITIVE

NON BRD..----.
INTACT
EGO ...

DIALOGUE
COMMITMENT

"CHANGE BEHAVIOR" .. !
FOLLOW EXAMPLE

BRD
IMPAIRED
EGO ...

ENDORSE PUBLICLY

I
I
I

DEM(tIJSTRATE SINCERITY
HABITUATE

REASSERT CONVICTION

PERSUASION

Figure 11-2. A parsimonious model of psychotherapy.

OUTCOMES

280

Ill. Rehabilitation Techniques

Figure 11-2 represents a juxtaposition between insight-producing therapies


in non-brain-injured individuals (i.e., people with intact ego functions) and the
type of therapeutic intervention suited for brain-injured (or ego-impaired)
individuals.
In essence, the therapeutic endeavor in people with intact ego functions
could be described as consisting of the setting, the therapeutic transaction
itself, and the outcomes of the therapeutic process. The setting consists of certain "structural" elements (e. g., outpatient clinic; individual or group; number
of sessions per week; the particulars involving seating arrangements) and
various "contractual" elements of the therapeutic undertaking (e. g., amount
of payment for services; manner of payment; policies concerning notification
of absences or vacation time; "rituals" of greeting; understandings concerning
what are and what are not "relevant" topics for discussion or mutual expectations between therapist and "client").
Considering the therapeutic process (i. e., the transactions taking place in the
therapeutic intervention), most therapists would agree that the objective of
psychotherapy is 1) to assist the client, during the process of self-examination,
to understand the cognitive basis of characteristic ways of behaving (i. e.,
thoughts, beliefs, convictions, intentions); 2) to assist him or her in confronting ("re-living") real feelings and emotions; and 3) to help him or her understand unique action dispositions (i. e., characteristic and/or habitual ways of
responding in given circumstances). Some would describe this process as
involving the analysis of one's "personality," while others (such as Allport
[28]) would refer to this process as one of examining and modifying one's
"generic attitudes." Thus, the core of the therapeutic transaction is selfexamination and self-understanding, followed by voluntary resolutions to
change or modify maladaptive attitudes and behaviors.
The main mode of exerting therapeutic leverage on the client is the therapeutic dialogue. It is essentially an exchange of verbal (and also some nonverbal) communications between the therapist and client, conducted according
to definite rules of the game. The rules are interpreted by the therapist. For
best effects, the client must endorse the rules and develop a therapeutic alliance
with the therapist. Following a period of internalization and assimilation of
the various messages conveyed during the therapeutic sessions, the client will
gain the necessary insights or understandings that lead to specific commitments
to change. A key factor, however, is that the client is autonomous in making
commitments to change, and acts voluntarily when implementing resolutions.
When sufficient changes have taken place in the client's central behavior
patterns (or habits), the result will be a change in the overall life-style of the
individual. This accounts for the dramatic "transformations" that can often be
observed in people following successful therapy.
Juxtaposed with this model for insight therapy is the model for therapeutic
intervention in the traumatically brain-injured, which is fundamentally different in several respects. The following is a brief explanation of the differences:

11. Structured Group Treatment

281

1. As Figure 11-2 indicates, one major difference is that the "contractual"


understandings in the brain-injury model are often achieved with the active
asistamce of the "significant others" of the brain-injured patient. Due to his
or her impairments, the patient cannot act as an autonomous "agent" on his
or her own behalf. In addition, brain-injured individuals often lack the sufficient understanding, will, and ability to exercise sound judgment and to
voluntarily accept therapeutic help, necessitating cajoling and even "pressuring" by their significant others.
2. The most fundamental difference between the two models of psychotherapy
clearly emerges in the therapeutic transaction phase. The main emphasis in
the therapeutic transaction in the brain-injury model is not an examination
of the underlying cognitive, affective/emotional, and behavioral components of the patient's generic attitudes, or his or her personality features;
rather, the objectives are:
To induce the patient to follow the example of others who demonstrate how
to perform certain specified therapeutic tasks
To induce the patient to act in certain ways explicitly
To induce the patient, by various direct means, to accept being corrected or
"coached"
To induce the patient to follow the example of others who demonstrate how
to perform certain specified therapeutic tasks
To induce the patient to act in certain ways explicitly prescribed by the
therapist(s) to modify those (unacceptable or maladaptive) behaviors specifically targeted for therapeutic-remedial intervention
To persuade (exhort and inspire) the patient to "endorse" and "own up to" the
ways of acting that were prescribed for him or her by the therapist
To induce the patient to become emotionally committed to assimilating and
personalizing (i.e., to feel strongly about) the ways of behaving and of reasoning initially instigated by the therapist
To induce the patient to make a public commitment (i.e., to assert in front of
others his or her intention) to practice the training regimen outlined for him
or her by the therapist(s), until the behavior becomes second nature to him
or her (i.e., habituated)
The principal therapeutic leverage in this therapeutic model for brain-injury
survivors is the frank application by the therapist(s) of various methods of
active persuasion, exhortation, and modeling of the desired behavior. Both the
call for specific behavior changes and the formulation of the priorities and
order of the behavior change emanates from the therapist. This is in stark
contrast to the therapy model for non-brain-injured persons, in which the
changes are initiated by the patient and ordered according to a self-generated
set of priorities.
Therefore, when viewed in contrast with one another, the two models have

282

III. Rehabilitation Techniques

quite different injunctions. The model for insight therapy implies: "First
examine your generic attitudes, resolve to change them, and then voluntarily
commit yourself to change your behaviors. When a sufficient number of these
behaviors have changed, your life-style will also be changed." The model for
therapeutic intervention with brain-injured persons has the following injunction: "First do as you are asked and shown, which is in your best interest and is
absolutely necessary for a better future adjustment. Now that you know how
and why to act this way, act with conviction. Do it out of your own will-not
just to please others. Once you grasp this way of behaving and thinking,
commit yourself to practice and make these behaviors part of your habit system. When you incorporate these 'templates,' your outlook on life (i. e., your
attitudes) and your life-style will improve. You will be able to attain a more
satisfactory adjustment."
The differences between the two models of therapy are fundamental and
have important ethical, psychological, and didactic implications. However, it
is beyond the scope of this chapter to discuss such issues. Rather, we will confine ourselves to outlining the principal features of a paradigmatic smallgroup remedial-therapeutic procedure developed especially for brain-injury
survivors. For reasons that will become apparent, this procedure is termed
group exercise rather than the more conventional term group therapy.
4 THE PARADIGMATIC GROUP EXERCISE

The following sections present an abbreviated and excerpted version of an


entire videotaped group exercise that is typical of the group exercises that take
place daily as part of the New York University (N.Y.U.) Head Trauma Program for neuropsychological rehabilitation. This clinical sample illustrates the
specially designed small-group procedure for brain-injury survivors. Various
aspects of the procedure will also be discussed.
4.1 The Setting

It is time for the daily one-hour group exercise. All eight patients (P-l through
P-8) currently in the program are gathered in the group room. Also present are
six members of the staff (S-l through S-6), four significant others (SO-l
through SO-4-both parents of one patient, the husband of another, and the
sister of a third), and two visiting professionals (V-l and V-2). All participants
arc closely seated in a semicircle. A video camera is set in place to tape the
group exercise. Opposite the video camera, at the other end of the open circle,
a large easel is set up with a two-foot by three-foot multicolored poster. To the
right of the easel, one chair is occupied by the staff person (S-l) designated to
be the leader of the exercise. To the left of the poster are two reserved seats,
one for the patient (P-l) called to occupy this "hot seat" (i. e., to perform in
front of the audience and the video camera) and the other for the staff person
(S-2) acting as "coach" to P-l on the hot seat.

11. Structured Group Treatment

283

4.2 Preparations

The leader invites P-l to occupy the hot seat and S-2 to occupy the seat to the
left of P-l and act as coach during the group exercise. As the leader, the patient, and the coach take their respective positions, staff members distribute
feedback sheets to the other seven patients (P-2 through P-8), checking that
each patient has a pen or pencil for entering notes and observations about P-l 's
performance in the hot seat. When all the participants are ready, the leader
begins the proceedings.
4.3 Induction and Instructions

The leader turns to P-l and begins the process of induction or preparation:
LEADER:

It is your turn now to be in the hot seat. You have seen how others have done
this exercise. The purpose of this exercise is to tell the group about two accomplishments in your life before the injury of which you are proud, and to explain
what these accomplishments or achievements meant to you. One of these
accomplishments should be an objective accomplishment [points to the poster].
As you can see, an objective accomplishment is something which had been
formally acknowledged by others [points to examples cited on the poster] such
as a citation, a grade at school, a certificate, a trophy, mention in the newspapers, etc. The second achievement should be a subjective one. A subjective
achievement, as you can see [again points to poster] is something you did or
achieved that meant a great deal to you personally. Others may not have
known about it. To repeat, your task is to tell the group about two accomplishments of yours, one objective and one subjective, which meant a great deal to
you. Take all the time you need to decide which of your past accomplishments
you wish to tell us about, to plan carefully how you will express your thoughts,
and to make sure that you say things just the way you planned them. You
know that in addition to making sure that you plan in advance what to say,
you will also have to pay attention to how you tell your story. This is an exercise of both thinking and planning and of social communication. When you
"deliver the goods," you have to do it in a way that captures the attention of
your listeners, in a way that makes them interested in what you have to say
and at the same time makes you appear as a likeable individual. So, plan what
to say, then deliver it in a socially interesting and pleasing manner. Watch the
tone of your voice and your manner of delivery. Look into the eyes of your
listeners and use the proper body language ....

The leader turns to the other patients and instructs them to carefully observe
P-l's performance on the hot seat and to make notes on their feedback sheets
under each of three headings: 1) How well did he or she comply with the intellectual requirements of the task? 2) Did he or she project the right tone of voice
and appear as interesting and likeable? 3) Did he or she accept coaching gracefully and put the coaching to good use? The leader tells the patients: "Your
notes will serve as guidelines for your feedback when P-l is finished with his
performance. "

284

III. Rehabilitation Techniques

Turning to P-l, the leader states: "Now, I will turn you over to your coach
(S-2). He will give you some special advice before you begin the exerCIse.
Your coach will also repeat the instructions one more time."
COACH:

Let me briefly remind you that my job is to help you perform this assignment
on the hot seat as well as possible. But before we begin, you must give me permission to coach you in front of the group. Because if you accept coaching and
permit me to interrupt you whenever I feel that a consultation between the two
of us will be helpful, then you will not resent being corrected in front of others.
Also, if you give me permission to coach you, chances are that you will not feel
bossed around by me. And you will not have to worry about people losing
their respect for you, because when you give permission to the coach, you are
in charge.

P-l:

All right, you've got my permission.

COACH:

Thank you. I will try to do it as gently and respectfully as possible. I will touch
you on your arm when I feel like interrupting you for some coaching.

P-I:

Fine. I am ready.

COACH:

All right. Here is a quick reminder of what your assignment is. [Coach then
reiterates the instructions given by the leader.]

4.4 Performance and Coaching

Having completed the instructions, the coach tells P-l to start the exercise.
P-l:

[turning to the audience] The objective accomplishment concerns my brief


career as promoter of concerts. Steve-my partner-and I did that. We promoted some of the most famous concerts [mentions some names of wellknown rock musicians] ... what quickly became apparent is that we needed a
lot of money to promote concerts. . .. [At this point, P-l tells in great and
meandering detail some ways in which he and his partner went about raising
the necessary funds. He becomes hopelessly entangled in peripheral issues.
Having apparently forgotten his main point, he stops and turns to the coach
with a quizzical look on his face.]

COACH:

All right, you are reporting to the group about an objective accomplishment of
which you are proud. This involved promoting concerts. Go ahead and tell the
group what exactly was the achievement and why you picked this example of
your objective accomplishments in the past.

P-l:

[in a petulant tone of voice] I was going to do that, but. ...

COACH:

[in a gentle, soothing voice] I believe you. You really don't have to give us
affidavits of honesty. I was simply trying to get you to deliver the "punch line"
of your story ....

P-l:

[still annoyed] But you were suggesting that I wasn't about to do that.

COACH:

[patiently, in an unruffled tone, with a smile aimed at defusing P-l 's obviously
hurt feelings] You know, your honor is not really at stake here. Your coach
simply did his job. If you tell the group your achievement in promoting those

11. Structured Group Treatment

285

concerts and why you feel so good about having accomplished that, you are
home free with your objective accomplishment.
P-l:

The accomplishment was that my partner and I promoted some of the best
concerts .... We were successful in pulling it off. .. good reviews in the papers
... I felt good. Wow! I did it all by myself. [turns to coach with a "seeking
approval" look]

COACH:

Since you looked at me, let me tell you that you have done the job. I think that
you have delivered your objective accomplishment. May I suggest that you go
on now to the next one?

P-l:

My subjective accomplishment is my career with the _ _ Corporation and


also how good I feel about my relationships with my family .... I feel very
strongly about it. ... [P-l begins to tell about his job as supervisor of sales at
__ Corporation. After meandering along for some time, he says the following:] With that came a lot of responsibility and slowly but surely I was not
grasping everything that I should have grasped. [P-l turns to the coach with a
puzzled look on his face]

COACH:

[realizing that P-l has forgotten his topic and has begun to reminisce about his
difficulties on the job after the head injury, he decides to reorient P-l] So your
subjective accomplishment had something to do with the way you combined
your work for your company and your family life. What was that exactly, and
why do you feel so good about it?

P-l:

I was able to balance my work with my relationships with my family. [turns to


the coach] More?

COACH:

Yes, you could amplify a bit on your statement. You reported that you were
pleased at being able to balance your dedication to a demanding career with
your devotion to your family life. What was so special about that? Try to explain
why this was such an important achievement in your life. Think things though
first, then explain to the group.

P-l:

It shows the inner strength that I had as an individual. .. to be able to balance


my career with my home life. [turning to the coach, he then adds:] I was
planning to complete my objective accomplishment ... [rambles on, apparently
forgetting that he is reporting on his subjective accomplishment]

COACH:

[reassuringly] No, no, you have done a good job with the objective accomplishment. You are finished with that. It has been sealed and delivered in good
order. What you are talking about now is how proud you were about being
able to devote yourself equally well to a demanding career and to your family
life. This obviously meant so much to you. This is the story of your subjective
accomplishment. So, if you feel like adding something to it, go ahead. But get
ready to wrap it up.

P-l:

[ignoring or misinterpreting the coach's latest comment, begins to review his


initial thoughts. With a tone of controlled anger and petulance in his voice, he
adds: 1Then you commented that I explain ... which took the meat and potatoes
away from what I was about to say.

286

III. Rehabilitation Techniques

COACH:

[gently touching P-l's arm and looking at him briefly in a calming tone of
voice] Let me see if! understood you correctly. Did something I asked or said
throw a monkey wrench in the system? Are you trying to tell me that I stole
your thunder by jumping in too soon with coaching?

P-l:

[looks embarrassed] No, no! It's my interpretation ... I did not say it right. .. I
did notciceliver.

COACH:

[puts his hand on P-l's shoulder] I don't agree with you. You have actually delivered on both the objective and the subjective accomplishments. [rapidly restates the essence of both ofP-l's accomplishments] But let me ask you if you
would like to accept one more challenge before we wrap things up?

P-l:

[with some anxiety] Yes, go ahead!

COACH:

When you told us about your subjective accomplishment, you told us that you
are proud of being able to do justice to both your career and to your family life.
In other words, you performed as a high-level executive for a big company and
at the same time led a quality family life. This came through loud and clear. Now
I would like you to reflect a bit and tell us why this would be a great achievement in everybody else's life as well, not just for you personally. [coach restates
his question in a simplified way and asks:] Do you get my point?

P-l:

[in a sarcastic tone] Yes indeed, and you are repeating yourself.

COACH:

All right, what is your answer to my question?

P-l:

[looks puzzled] Subjectively, I have the inner strength ... [continues to restate
what he has already said before, missing the point of the coach's last question.
The coach decides to model (i.e., demonstrate) the correct way of answering
the question.]

COACH:

Let me put myself in your shoes and demonstrate how I would answer the
question if I were you. [briefly reiterates the question as to why the ability to
balance the demands of a busy career with a quality family life is a great achievement] If I were you, I would answer the question like this: The reason I think
the ability to balance one's career with a close family life is a true achievement
is because in our society many people fail to accomplish that. The result is that
many people give up a quality family life, sacrificing everything for their
career. [turning to P-l] What do you think about your coach's demonstration?
Does it make sense to you?

P-l:

That thought crossed my mind .... [circumlocutes about what he had wanted
to convey in his earlier answers, implying that the coach unnecessarily intervened] I was ready to come around to it.

COACH:

I see. What you are telling me is that the coach reacted too soon with coaching.
Had he waited, you would have eventually come around to answering the
question on your own. You did not really need his help. You did not need his
demonstration.

P-l:

That's right!

COACH:

I see. Well, as your coach, my job is to help you make your performance as
clear and crisp as possible, to make things so explicit that everyone of your

11. Structured Group Treatment

287

listeners understands what you are trying to get across. This means to help
you fully articulate your thoughts. That is why I intervened. That was our
agreement when you gave me permission to coach you, remember?
P-I:

[appearing mollified] Y (,S.


if you had that answer in mind, I wish you had come out with it immediately, because that was the shortest and most effective way to reply to the question. This exercise was designed to help people communicate their thoughts
effectively. [turns to leader] We are finished.

COACH: SO,

4.5 Feedback
LEADER:

[turning to P-l J Before you receive feedback from your peers and the staff, it is
your job, as you know, to be the first to evaluate your own performance.

The leader then engages P-l in the self-evaluation process. In response to the
question of how well he planned and organized his thoughts and delivered the
contents of his statements, P-l replies that "intellectually I feel that I gave all
the information." When asked whether he demonstrated the ability to "think
on your feet," P-l felt that he "ad-libbed" satisfactorily on the hot seat. Interpresonally, P-l felt that he "came across rather nicely" but that "maybe I tooted
my hom a bit too loudly." In respect to the question of his acceptance of
coaching, P-l felt that he "worked with him [i.e., the coach] pretty nicely."
At this point, the leader turns to the other patients and invites each one to
use the feedback sheet to appraise P-l's performance "cleady, concisely, and
with the appropriate feelings." For the sake of brevity, only the key statements
from each patient are cited below:
P-2:

I thought that you organized your words well. You were not the least
nervous-your eye contact was very good ... you were not too prideful and
you accepted coaching very well.

P-3:

I felt that you stayed on track and gave sufficient information, what was
needed. You showed that you were proud of your accomplishments, I could
fed it. You accepted the coach's suggestions.

P-4:

[speaks in third person] He jumped around and lost me a couple of times ...
too much interference by the coach [turns to the coach and admonishes him]
Leave him alone next time!

LEADER:

[intervenes and reminds P-4 that feedback is to be given to P-l only and that
the feedback should be delivered personally, face to face, in a friendly manner
and not as a finger-wagging admonition]

P-4:

[apologizes to P-l and completes her feedback in a "softer" and more personal
tone]

P-S:

[moderately aphasic, he reads his feedback from his notes, which is permitted]
Your hot seat performance was straight and clear. You were smooth and loud
and coaching you communicated very well ... accepting ... very good job.

288

III. Rehabilitation Techniques

P-6:

The intellectual part, you were very interesting. You supplied enough information, maybe at times too elaborate. Interpersonally your posture was good.
Maybe you could speak a little louder, though. You came across very proud of
your accomplishments. Coaching, you impressed me. Compared with your
first time on the hot seat [when P-l openly fought against coaching] you
turned to the coach a couple of times. I am really impressed.

P-7:

Being the first one [on the hot seat in this exercise] you were outstanding,
think. You had a bit of difficulty with organization but for being the first onc,
I could see that you were shaky at times, but you did OK.

P-S:

I enjoyed listening to you. Your voice was conversational and you were
organized. Particularly the way you summarized. You were proud and it was
clear why you were proud. You did what was asked of you by the coach.

At this point, the leader asks for feedback from the staff and other participants (i.e., the significant others and the visitors). An abridged verSIOn,
illustrating the nature and tone of this feedback, follows:
LEADER:

Tighten your seat belt because I have a lot of information to give you. In some
ways, your performance was a success; in other respects, there is a need to
improve things further. There was no inappropriate tooting of your horn.
You came across as a warm and personable communicator. You turned to the
coach when you experienced difficulties. But there was some open arguing
with the coach, some testiness. At times, the help offered by the coach was not
graciously accepted; you quibbled ... [explains need for trusting coach and
accepting help without being defensive about it. Points out how, in fact, after
each intervention by the coach, P-l was able to come up with a "good finish"]
I told you that I will speak to you straight because I respect you as a man. Are
you offended?

P-l:

Oh, no. Not at all!

S-3:

You came across as an attractive, warm person in the way you communicated.
[briefly mentions P-l's tendency to lose focus when irritated]

S-4:

I was impressed with your improvements in the intellectual sphere since your
last exercise on the hot seat. [mentions P-l's articulateness and conciseness
after coaching, and P-l's turning to the coach, but agrees with the leader that
"at times, you argued with the coach instead of accepting his suggestions
straight"]

S-5:

This was a much improved performance compared with your first one in the
hot seat. Good, but a bit rough at the edges .... [proceeds to point out good
points and then, discussing the argumentative style exhibited by P-l when
coached, says] I can empathize with you. For someone who has been a takecharge man, a leader, it is very difficult to accept help or admit that you didn't
understand something. But you really must try; there is no other way ....
[continues with some compassionate exhortations to inspire P-l to "bite the
bullet"]

11. Structured Group Treatment

289

SO-I:

I started out being very impressed with your performance at the beginning but
then suddenly you just ran out of steam. I agree that you turned to coach for
help but you also fought him. The trust was just not there.

SO-2:

I happen to think that you were pretty well-organized and pretty good at
accepting coaching. It is very hard to be in the hot seat. I think you did your
best.

S-6:

You started out well-organized ... considering the cognitive deficits with
which you must wrestle, this was an impressive performance. [cites several
highlights] I too agree that there was a tug-of-war with the coach. [explains
with empathy "how hard you have been struggling with accepting help since
the beginning"] ... Even though there are some things that must be worked
on, do not become discouraged; there were some beautiful things in your
performance today. Rome was not built in one day.

COACH:

[prefaces his statement with an invitation to P-l to discuss further what he was
about to say next] I have told you already that the greatest compliment we can
pay you is to tell you the truth, without beating around the bush. By talking to
you straight, man to man, I am expressing my confidence in your integrity
and gutsiness as a person. I am not talking down, or around you, like to a
"brittle" patient, but am showing you that I respect you as a man. Your performance today, I would characterize it by the saying: "The spirit was willing but
the flesh was weak."

The coach reviews three of P-l's demonstrated characteristic problems: 1) a


difficulty in being able to think on his feet and reason through problems;
2) a tendency to become irritable and emotionally flooded under stress, which
exacerbates his analytic-synthetic difficulties; and 3) the tendency to argue
about constructive criticism because his sense of self-esteem has been offended.
The coach then openly exhorts P-1:
COACH:

LEADER:

P-l:

Your rehabilitation depends on your ability to swallow these truths without


choking on them. I believe you have the character and the wisdom to do it. We
are here to help. If you will not be able to tackle these problems, then I fear for
your future. On the other hand, if you trust us and accept coaching-knowing
full well that in this group you need not fear losing our respect-there is much
hope for getting you back into shape. [expresses his empathy with P-l, asserting] Naturally, you find it hard to accept help in public and to admit that you
can't do certain things which came to you very easily prior to your injury. My
heart goes out to you. Ambitious, hard-driving people like to be in charge;
they hate being told by others what to do. But the road toward your rehabilitation demands that you face the truth with courage and dignity. Trust uswe won't let you down!
[turning to P-l] How do you feel now?
I don't have much to say right now. Intellectually, I feel that I have been browbeaten [coach puts his arm around P-l's shoulder; other staff members say
"yes!"] ... But I will work on it. [appears subdued]

290

III. Rehabilitation Techniques

COACH:

Did you feel humiliated by what I said to you?

P-l:

Yes!

COACH:

Oh, I am sorry; this should not be. We truly respect you and do not wish to
hurt your feelings. We\yil! have to do something about this.

P-l:

I know, in my head, intellectually, I understand, but it is hard to take ....

S-6:

We care for you, you know it!

P-l:

Yes, I know it, and I truly appreciate what you are trying to do.

LEADER:

This is an excellent beginning toward the solution. Because of that, this


exerCIse was a success ....

4.6 Follow-up

A few minutes after the group exercise was over, the leader and the coach had
a brief personal session with P-1. During this session, the main features of
P-1's performance were repeated and the various messages were reiterated in a
supportive and compassionate atmosphere. P-1 was visibly moved, and he
promised to work hard on overcoming his tendency to resent and fight coaching. He was encouraged to take the videotape home and view it together with
his wife. Several days later, P-1 and his wife attended a follow-up session on the
same exercise. This exercise was a more personalized version (i.e., only P-1,
his wife, and several members of the staff were present). Again, a multicolored
poster was prepared and the video camera was used to record the proceedings.
The poster outlined, on one side, P-1's resistance to being confronted or
corrected about his deficits. The language was direct, evocative, dramatic, and
emphatic (e.g., "This was a take-charge, proud, executive type." "He liked to
tell others what to do, not to be told by others what and how to do." "This
kind of man finds it hard to swallow being taught by others." "He finds it very
difficult to admit that he can't do something as well as he used to. "). On the
other side of the poster, some of the key reasons why P-1 "must learn to accept
coaching openly, without resentment, and without feeling put down or put
out" were outlined.
P-1, with his wife's active participation in the session (with both following
explicit cues of the group leader), performed another exercise in which the
staff-alternating between gentle confrontation, frank persuasion, and emphatic support-helped P-1 absorb and acknowledge the messages delivered by the
poster. The videotape of the exercise was taken home to be viewed and further
assimilated by P-1. In this way, P-1 gradually became "transformed." By the
end of the 20-week treatment cycle, he was able to attain the minimum degree
of acceptance of his disability and adaptability necessary to ensure his stable
future adjustment.
5 THE GROUP EXERCISE AS AN OPTIMAL STRUCTURE

The group exercise possesses certain structural features that make it possible
1) to optimize patients' attention and concentration and ensure stability over

11. Structured Group Treatment

291

time; 2) to assist patients in formulating their thoughts (via the availability of


modeling and instant cues, or "coaching"), monitoring and correcting their
performance; and 3) with the benefit of the systematically guided self-critique
and feedback from others, to help patients attain the proper "closure" after the
exerCIse.
The principal structural aspects of the group exercise have been described
elsewhere [2, 3]. At this point, we will merely identify them. These consist
of 1) the preselected theme; 2) the "poster" and careful and explicit induction
and instruction; 3) the total control and guidance of the leadership, designed to
ensure that only one person speaks at a time and that comments are well-targeted and concise; 4) the ritualized and orchestrated aspects of the exercise, designed to facilitate the participation of the timid and inhibit the unmodulated
reactions of the disinhibited; and 5) the systematic and guided coaching and
feedback, designed to facilitate the patient's performance and allow for instant
verification of the performance by both the person on the hot seat and the
listeners.
6 THE GROUP EXERCISE AS A VEHICLE FOR
PUBLIC DISCLOSURE AND COMMITMENT

The paradigmatic group exercise incorporates both the various psychodrama


and encounter types of techniques and certain classic elements of persuasion
and healing so aptly described by Frank [29]. It also embraces certain insights
from social psychology, such as the use of exhortation to increase the "achievement drive" described by McClelland [30].
At the core of this technique is the inducement of patients 1) to assert, in the
presence of others, their deficits and the need to assume a realistic view of what
can and should be done about ameliorating the effects of their deficits; and 2) to
commit themselves publicly to work toward modifying their maladaptive
behaviors in accordance with the prescriptions and guidance of rehabilitation
specialists.
Clinical experience in the neuropsychological rehabilitation of brain-injury
survivors in the last decade [8, 26, 31] has shown the superiority of variants of
this group exercise method over the more conventional one-on-one or group
approaches to psychotherapy with this population.
7 THE GROUP EXERCISE AS A REHABILITATION ALGORITHM

Although this group exercise technique was designed to complement a series


of other activities in the holistic neuropsychological rehabilitation of the braininjury survivor, the technique contains all the necessary elements for achieving
the objective. In a recent publication [8], we presented a model for the holistic
rehabilitation of brain-injured people. The model posed the need for the braininjured person to advance through a hierarchy of six distinct intervention processes in order to be able to reconstitute (to the degree possible) his or her "ego
identity, " which had been shattered by the head injury and its sequelae. We use
the concept of ego identity as defined by Erikson [32].

292

III. Rehabilitation Techniques

The model is illustrated in Figure 11-3 as a mountain-climbing metaphor.


Applying the model to the group exercise, the brain-injury survivors 1) must
be helped to become properly engaged (i.e., focus their attention and sustain
their concentration); 2) must be helped in becoming aware of their deficits
(while retaining the will to overcome their handicaps and to practice the prescribed remedial exercises; 3) must be assisted-through systematic trainingin the mastery of various "templates" of behavior; 4) must compensate for their
deficits; 5) must attain increased functional competence; and (its corollary) 6)
must accept their existential situation. Thus, the successful passage of a patient
through each of these six stages culminates in the reconstitution of his or
her ego identity. Among the characteristics of ego identity are an enhanced
self-acceptance, calm resignation to one's fate, and peace of mind.
The group exercise, by its very structure, and as a vehicle for public "owning up" to one's deficiencies, fostering a public commitment to modify dysfunctional behaviors, and conditioning the patient to voluntarily and publicly
accept soliciting and receiving help (i.e., coaching without feeling "demeaned"
by it) is, in effect, a vehicle for the internalization of a rehabilitation algorithm.

DESPAIR
Figure 11-3. A model for rehabilitation following a head injury.

11. Structured Group Treatment

293

8 THE GROUP EXERCISE AS A COGNITIVE HIERARCHY

The group exercise was designed as a procedure for systematically addressing


several needs of the brain-injured person. At one end of the continuum is the
desire to facilitate the patient's ability to process information and to formulate
and effectively communicate ideas and feelings. At the opposite end is the concern that the patient achieve the minimum degrees of awareness, compliance,
and malleability so as to be capable of attaining acceptance and a reconstituted
ego identity. It is therefore important that in selecting the themes (i.e., contents) of the various exercises a proper balance be maintained between the two
opposmg concerns.
To address this issue, the New York University Head Trauma Program has
devised a "curriculum" of themes [2, 3]. This hierarchy of increasingly more
intellectually challenging themes for the group exercises seeks to accommodate the cognitive as well as the clinical needs of the patient who is undergoing
holistic neuropsychological rehabilitation. This hierarchy consists of the following themes: 1) introduce yourself; 2) tell about two of your achievements
(in your life prior to the injury); 3) tell about two personal qualities you admire
in yourself; 4) tell one member of the group about some qualities you admire
in him or her; 5) pretend you are a member of the staff and interview a prospective candidate for the program [the candidate is the patient himself or herself as
role-played by a member of the staff]; 6) evaluate your progress (as well as the
remaining unsolved problems) in collaboration with a member of the staff in
front of the group.
9 SUMMARY

This chapter has described a structured group treatment procedure for braininjury survivors. This therapeutic methodology has been successfully applied
in various clinic settings.
There can be no single, omnibus neuropsychological rehabilitation technique for people with diffuse brain injury. To attain the best outcomes, a
number of complementary remedial interventions need to be orchestrated and
applied within a holistic rehabilitation framework. Still, the paradigmatic
group exercise method proposed here is a powerful and versatile clinical tool.
It contains a number of elements designed to foster awareness and acceptance
of the consequences of one's brain injury, increase malleability to treatments,
and help in adopting a realistic stance towards what is possible to achieve in
rehabilitation.
REFERENCES
1. Ben-Yishay, Y, Ben-Nachum, Z., Cohen, A., Gross, Y., Hoofien, D., Rattok, J. and
Diller, L. (1978). Digest of a two year comprehensive clinical rehabilitation research program
for out patient head injured Israeli veterans. N.Y.U. Rehabil. Monogr. 59, 1-61.
2. Ben-Yishay, Y. (1979). Structured group techniques for heterogeneous groups of head trauma
patients. N.Y.U. Rehabil. Monogr. 60, 39-88.
3. Ben-Yishay, Y., Lakin, P., Ross, B., Rattok, J., Cohen, J. and Diller, L. (1980). Developing a

294

4.

5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.

III. Rehabilitation Techniques

core curriculum for group exercises designed for head trauma patients who are undergoing
rehabilitation. N.Y.U. Rehabil. Monogr. 61,175-235.
Lakin, P., Ben-Yishay, Y., Rattok, J., Ross, B., Silver, S., Thomas, j.L., Fawzi, E.M.,
Hamza, M.H. and Diller, L. (1981). Special procedures for assessing aspects of interpersonal
skills of head trauma patients undergoing rehabilitation. N. Y. U. Rehabil. Monogr. 62,
68-102.
Ross, B., Ben-Yishay, Y., Lakin, P., Rattok,j., Silver, S., Thomas, L. and Diller, L. (1982).
Using a "therapeutic community" to modify the behavior of head trauma patients in
rehabilitation. N.Y.U. Rehabil. Monogr. 64, 58-91.
Rattok, j., Ross, B., Silver, S., Thomas, L., Ben-Yishay, Y. and Diller, L. (1983). Understanding the world of work: A small group exercise for head trauma patients in rehabilitation.
N. Y. U. Rehabil. Monogr. 66, 92-113.
Ben-Yishay, Y., Lakin, P., Ross, B., Rattok, j., Piasetsky, E. and Diller, L. (1983). Psychotherapy following severe brain injury: Issues and answers. N.Y.U. Rehabil. Monogr. 66,
127-148.
Ben-Yishay, Y., Rattok, ]., Lakin, P., Piasetsky, E., Ross, B., Silver, S., Zide, E., and
Ezrachi, I. (1985). Neuropsychological rehabilitation: Quest for a holistic approach. Semin.
Neurol. 5, 252-277.
Benton, A. (1979). Behavioral consequences of closed head injury. In Central NeYl'ous System
Trauma Research Status Report, NINCDS, G.L. Odom, ed., National Institutes of Health,
Washington, DC, pp. 220-231.
Lishman, W.A. (1968). Brain damage in relation to psychiatric disability after head injury. Br.
]. Psychiatry 114, 373-470.
Brosin, H.W. (1965). Psychiatric conditions following head injury. In American Handbook qf
Psychiatry, Vol. 2, Arieti S., ed., Basic Books, New York, pp. 1175-1202.
Boll, T.]. (1978). Diagnosing brain impairment. In Diagllosis of Me/Ita I Disorders: A Hmldbook,
Wolman B.B., ed., Plenum, New York.
Rimel, R.W., Giordani, M.A., Barth,j.T., Toll, TJ. andJane,].A. (1981). Disability caused
by minor head injury. Neurosurgery 9,221-228.
Jane,j.A., Rimel, R.W., Pobereskin, L.H., Tyson, G.W. and Gennarelli, T.A. (1980). Outcome and pathology of head injury. In Proceedings of the Fourth Chicago Conference on
Neural Trauma, Chicago, September.
Oppenheimer, D.R. (1968). Microscopic lesions in the brain following head injury. j. Neurol.
Neurosurg. Psychiatry 31, 299-306.
Jane, j.A. and Rimel, W.R. (1982). Prognosis in head injury. Clin. Neurosurg. 29, 346-252.
Goldstein, K. (1959). Functional disturbances in brain damage. In Americall Handbook of
Psychiatry, Vol. 4, Arieti S., ed., Basic Books, New York, pp. 182-207.
Moore, j. (1980). Neuroanatomical considerations to recovery of function following brain
injury. In Recovery of FIII/ctio/I: Theoretical Considerations for Brain II/jury Rehabilitation, BachY-Rita, P., ed., University Park Press, Baltimore, pp. 9-90.
Ben-Yishay, Y., Rattok, J., Ross, B., Lakin, P., Ezrachi, 0., Silver, S. and Diller, L. (1982).
Rehabilitation of cognitive and perceptual defects in people with traumatic brain damage.
N.Y.U. Rehabil. Monogr. 64,127-176.
Reik, T (1964). Listening with the Third Ear. Pyramid Books, New York.
Tarachow, S. (1963). All bztroductiofl to Psychotherapy. International Universities Press, New
York.
Goldstein, K. (1952). The effects of brain damage on the personality. Paper presented at the
annual meeting of the American Psychoanalytic Association, Atlantic City, NJ, May.
Goldstein, K. (1951). Humall Nature ill the Light ofPsycllOpathology, Harvard University Press,
Cambridge, MA.
Bellak, L. (1977). Psychiatric states in adults with minimal brain dysfunction. Psychiatr. Ann.
7, 58-76.
Grimm, B.H. and Bleiberg,j. (1986) Psychological rehabilifation in traumatic brain injury. In
Hmldbook of Clinical Neuropsychology, Vol. 2, Filskov, S.B. and Boll, Tj., eds., John W'iley
and Sons, New York, pp. 495-527.
Prigatano, G.P. et aI., (1986). Neuropsychological Rehabilitation After Brain Injury. Johns
Hopkins University Press, Baltimore.

11. Structured Group Treatment

295

27. Block, S.H. (1987). Psychotherapy of the individual with brain injury. Brain Injury 1,
203-206.
28. Allport, G. W. (1961). Pattern alld Growth in Personality, Holt, Rinehart and Winston, New
York.
29. Frank, J.D. (1963). Persuasion and HealilJR. Shocken Books, New York.
30. McClelland, D. (1968). The Achieving Society. Van Nostrand, New York.
31. Ben-Yishay, Y. and Prigatano, G.P. (in press). Cognitive remediation. In Rehabilitation of the
Head Iniured Adult, 2nd ed., Griffith, E.R., Rosenthall M. and Bond M., eds., F.A. Davis,
Philadelphia.
32. Erikson, E.H. (1968). Identity: Youth and Crisis. W.W. Norton and Co., New York.

12. LONG-TERM FAMILY INTERVENTION

HARVEY E. JACOBS

1 INTRODUCTION

The effects of persisting sequelae of traumatic brain injury (TBI) on patients


and family members have been well documented. A number of studies have
identified a broad range of effects on family relationships relative to the time
since injury, age and sex of the survivor, premorbid life style, family composition, level of recovery, behavior problems, family resources, and a host
of other factors [1-8]. These findings indicate that there is no one course of
adaptation and reaction on the part of family members, but rather a multitude
of individualized circumstances and effects that may shape family functioning.
It is not surprising that families experience such effects. A rich body of research and literature on family systems in the face of chronic or catastrophic
illness already exists-for example, studies about dementia [9, 10], marital
therapy [11], childhood disorders [12, 13], alcoholism [14], drug dependency
[15], and psychiatric disabilities [16]. The general -reaction-adaptation process
of families has also been the subject of much discussion, most notably by
Kubler-Ross [17], with others offering complementary paradigms [18-20].
The specific components of reaction and adaptation may vary across these
models, but there is the general consensus that an adaptation process begins
with shock and denial and proceeds through steps of anger, accommodation,
bargaining, grief, eventual acceptance, and continuation of living.
Families act as well as react to such situations, however, and they have played
an integral role in treatment and recovery for other populations of people with
297

298

III. Rehabilitation Techniques

disabilities. For example, the comprehensive network of services and systems


to meet the needs of people with developmental disabilities was principally
developed through the efforts of concerned parents and citizens [21, 22]. In the
field of mental health, Falloon et al. [16] demonstrated that family members
who were given proper training and support were highly effective in preventing relapse among persons with schizophrenia. This study was also significant because it reversed a long-held view that families were more likely to
exacerbate rather than reduce the symptoms of this disorder [23]. Programs
addressing alcoholism [14], drug dependency [15], childhood adjustment [12,
13], and other disorders frequently incorporate family interaction as a prime
requisite in treatment and outcome maintenance.
2 FAMILY REACTIONS TO TRAUMATIC BRAIN INJURY

Parallel family-process and outcome findings are also being documented


following TBI. Muir and Haffey [24] have suggested that families of TBI
survivors experience many of the same stages of adjustment to illness and
death that have been noted by Kubler-Ross [17] but progress through these
stages dynamically rather than successively-a process they call "mobile
mourning." The Kubler-Ross model portrays a sequence of stages concluding
with the death of the victim and the continuation of other family members'
lives. For the family of the traumatic head-injury survivor, however, there is
frequently no "final resolution." Instead, there is a series of cyclic events,
beginning with the medical emergency and continuing through physical
rehabilitation, return to the community, resocialization, vocational rehabilitation, and the long-term psychosocial adjustments that require ongoing adaptation. Family members of survivors of TBI circulate through the stages of
adjustment according to the fortunes and prognosis of the patient.
The initial shock of the accident is resolved during the patient's medical
recovery, only to be supplanted by problems faced during rehabilitation. For
those patients who are unable to return home, the continuing struggle to
maintain long-term treatment and placement causes continuing fluctuation in
adaptation [25]. The psychosocial problems of the survivor who does return
home frequently overshadow the medical sequelae [26]. Recurring behavior
problems, legal or financial obligations, social isolation, vocational misfortunes, and community pressures frequently follow the survivor and affect
family adaptation in the process.
3 A MODEL OF LEARNED HELPLESSNESS

The ultimate frustration and sense of powerlessness that may occur for both
patients and families as each new cycle of problems is presented can result in
a condition known as "learned helplessness" [27]. Under circumstances of
repeated problems and no resolution, people develop the perception (in effect,
"learn") that they have no control over major events that affect their lives.
They may become dependent, depressed, and passive to the specific issues that
foster this perception.

12. Long-Term Family Intervention

299

In a review of this concept, Seligman [27] noted that learned helplessness


usually begins in one specific area of a person's life and generalizes to other
areas over time. During its early stages, learned helplessness is highly specific,
and people can clearly discriminate between situations of learned helplessness
and other life events. If learned helplessness expands to other life areas, this
distinction may begin to blur across daily events. In a classic vicious cycle, the
harder that people try to manage the situation (even though they remain unsuccessful), the more withdrawn and self-defeated they become. In effect, they
begin to lose perception of the controlling variables in their lives, becoming
less effective in meeting their daily needs, more despondent, and more depressed. Each new episode of perceived uncontrollable and unpredictable events
fosters more helplessness and lack of control.
Fortunately, the process of learned helplessness can also be "unlearned" if
people can be taught how to control part of their environment and independently implement this knowledge. In the initial steps in this unlearning process,
the factors controlling a person's life must be explained and demonstrated. He
or she must then be assisted to successfully implement effective strategies in an
actual problem area. Further success depends on effective generalization of
control to other areas of daily life. Similarly, the earlier that intervention
begins-and the less severe the level of helplessness-the sooner noted problems can be resolved.
4 FACTORS THAT FOSTER LEARNED HELPLESSNESS

Unfortunately, the nature of treatment following a brain injury may naturally


facilitate the development oflearned helplessness. News of the injury is a major
shock to family members, who will be stunned by its sudden onset. Most
people either will be unable to comprehend the information or will deny the
severity of the patient's admitting medical condition. During early phases
of treatment, there is little that family members can do to directly help the
patient, and they are frequently overwhelmed by the volume of information
that they are trying to absorb. The treatment team may try to involve them in
some of the decision-making processes, as well as in the consent procedures
that are legally required, but the family members are likely to be ignorant of
the complex medical and restorative issues that the survivor is facing. In this
way, the role of passivity and dependency in a situation outside of their control
gains a foothold during early phases of treatment.
Attempts to involve family members during the survivor's inpatient rehabilitation may also be too few and too far between to reverse a patient's (and
family's) already-developing dependency. Even after they receive supportive
therapy and education about recovery from brain injury, many family members still expect the survivor to come home "cured." In other cases, family
members may remain somewhat removed from the treatment-either by their
choice or by the choice of the treatment team. The family's decision not to get
personally involved in daily treatment often revolves around lack of time.
Once the survivor is medically stable, many family members must attend to

300

III. Rehabilitation Techniques

the other personal responsibilities and obligations that were disrupted by the
initial catastrophic event. The treatment team may also feel that the family's
involvement is counterproductive to their own goals for the patient. For
example, they may think that the family "babies" the patient too much, takes
up too much staff time asking questions, and so on. Among those family
members who do not become involved, whether by family or team decision,
dependency and lack of control can continue to develop.
In most situations, case management of the patient is transferred to the
family following discharge from inpatient treatment, when many other treatment services are also curtailed. Ironically, this is also the time when the more
pervasive psychosocial, behavioral, and financial problems begin to take preeminence over medical issues [8]. Although professional help is crucial-both
to facilitate transition of case management and to teach families how to meet
these new issues-professional help is more difficult to secure. At this point,
some family members may find themselves frozen and helpless because they
learned to rely on professional help during earlier stages of recovery. They may
not know what to do when formal assistance is no longer available. Others
may demonstrate the drive and ambition, but not the skills, to address problems as they arise; this results in "spinning one's wheels" and ultimate frustration. A few will effectively manage the situation. However, most families will
be caught in a web of relative confusion, loss, and lack of direction.
5 MODELS OF SUPPORT

Formal rehabilitation programs for survivors of traumatic brain injury are


in their infancy, compared to programs for people with other disabilities. For
this reason, few resources are available to meet the long-term needs of
survivors ofTBI. At the national level, the establishment of the National Head
Injury Foundation (NHIF) in 1981 was the premier step in the process of
awareness and paralleled the establishment of family-initiated groups for other
populations with disabilities, such as the Association for Retarded Citizens
(ARC) and the National Alliance for the Mentally III (NAMI). Over the
past eight years, the NHIF advocacy and support groups have made much
progress in gaining public awareness and personal support. However, the
effort required to identify the broad spectrum of needs for this population also
points to the extensive work that remains in translating awareness into active
programs.
Private-sector interests have also been active in addressing some of these
needs. The field has grown from a less than a handful of programs ten years
ago to more than 600 today [28]. However, fewer than 10% of these are specifically designed for post-rehabilitation services, and most are out of the financial reach of those in need.
Nonprofit, service-based organizations offer another option to some longterm care needs. In recent years, new groups have formed, and old ones have
adapted, to meet the needs of this growing population. For example, Project

12. Long-Term Family Intervention

301

Headway and Accessible Space Incorporated-New Beginnings for Brain Injury


are representative of organizations that have recently been developed specifically to meet the long-term needs of TEl survivors. In addition, more established organizations (such as Easter Seals and United Cerebral Palsy) have also
begun to address this population by incorporating TBI survivors into existing programs as well as by establishing specific programs for the survivors
themselves.
Despite all of these efforts, the primary responsibility for care and services to
survivors of TBI still falls to the families. Thomsen [6] noted that 73% of all
survivors of severe traumatic brain injury could not live independently and
returned to their families for assistance. Jacobs [3] reported that in situations
in which assistance was required, family members provided the assistance 82%
of the time, whereas professionals provided assistance 18% of the time. In the
same report, it was also noted that 42% of the survivors surveyed required
full-time supervision. In 40% of these cases, a family member had given up a
full-time job or education to provide this supervision.
There is no one model or form oflong-term service delivery for this population. Each survivor has a unique set ofliving arrangements, skill levels, degree
of support, independence, financial resources, and a host of other factors. No
matter what services each individual needs, however, programs must focus
upon increasing self-sufficiency and independence-in effect, upon helping
the person return home. Clearly, family members will playa vital role in
long term TBI rehabilitation services, regardless of the models that evolve.
6 WHAT IS "THE FAMILY"?

Before we can advocate that families be involved in rehabilitation programming, we need to understand exactly what a family is. The family may be
viewed as a system in which a composite of individual members work together,
at least in part, to meet both individual and mutual goals. Over the course of
time, each family member synergistically consumes and contributes to the
resources of the group. These resources include money, material goods, love,
affection, attention, teaching, support, and other needs of each individual.
Simplistically speaking, young children may consume more resources than
they provide to their parents, but this "imbalance" is reciprocated by the love
they return, and in later years through assistance to their aged parents. The
breadwinner of the family, who provides vital financial resources, may also
require significant attention and affection from others in exchange for his or
her labor.
When the give and take of resources are in balance, the family is also likely
to be in balance, because each person's needs are being met. When the balance
is disrupted, family problems are likely to arise. Decompensation may include
the dissolution of the entire family, marital break-up, one or more members'
moving out, increased fighting, loss of communication, decreased positive
affect, and lack of direction. Such decompensation may be caused by lack of

302

III. Rehabilitation Techniques

resources (e.g., by the loss of a job) or a need to redistribute resources and


goals (e. g., in the case of a traumatic or prolonged illness of one family member). Although most family systems develop sufficient reserves to adjust to
short-term imbalances in resources and priorities, few have sufficient emotional, financial, or social resources for protracted periods of imbalance. In
such cases, they may begin to decompensate. Accordingly, formal intervention
must focus on identification of the components, resources, and priorities of the
family system, as well as on how to balance these issues in light of changed
demands and possible redefinition of the family unit.
Although the family-systems approach provides important perspectives for
formal assessment and intervention, it can also cause one to lose perspective of
the individual members who make up the system. Each member of the family
system is also a unique individual who has his or her own specific goals, needs,
methods of coping, and adaptation that will influence his or her reaction to the
injury and its subsequent sequelae. Hence, although it is important to consider
overall family-system dynamics in long-term rehabilitation, it is just as important to recognize the individual family members' concerns. Ultimately,
family training and treatment involves work with individual members, including the survivor, according to each person's needs, interest, motivation,
and ability. There is no such thing as "the family," but rather a composite of
individuals.
7 THE ROLE OF FAMILIES

By now, it should be evident that the role of the family in the rehabilitation
process is implicit, although there is no one static course or role that all families
should follow. Each case must be judged individually on a number of interrelated factors, including family dynamics before the trauma, role changes
after the trauma, the patient's status and progress, individual reactions, treatment variables, economics, resources, and other outside influences.
The family can make many positive contributions to the overall rehabilitation process:
1. First, supportive family members can often provide more contact hours
to the survivor than outpatient treatment programs can, or they can be the sole
source of treatment when programs are not available. Even moderately functioning families are available almost constantly, compared to the limited hours
of costly professionals [9].
2. Family members may be more motivated to continue long-term and
intensive treatment in search of small but important gains, when others have
given up. The significant effect that a concerned and caring family member
can have on the quality of a patient's outcome has been repeatedly acknowledged in the rehabilitation literature [24, 29, 30]. On the other hand, negative
family roles can retard progress if they are not treated [31, 34].
3. Experience in other clinical areas such as child management [12, 35, 36]

12. Long-Term Family Intervention

303

has demonstrated that many daily training and rehabilitation programs can be
conducted by the family, with proper guidance, at a lower cost than that of
professional service delivery.
4. Clinical evidence indicates that family members may personally benefit,
through a sense of understanding and accomplishment, when they are actively
involved in the rehabilitation process [37]. Families are often able to work out
problems of guilt, helplessness, and anger by becoming a productive force in
the patient's recovery [27, 38].
5. There is the financial reality of the cost of rehabilitation to society. Unless there are major breakthroughs in treatment or financial aid, there are few
alternatives available to families, except to use their own resources for the
patient's extended treatment [39].
6. As was previously noted, informed and concerned family members are
typically the strongest advocates for the person with the disability. Through
the years, in the absence of needed services, groups of affected families of other
disabled populations-such as ARC and NAMI-have organized to provide,
or lobby for, appropriate treatment and community involvement [21, 22].
8 THE ROLE OF PROFESSIONALS

Professionals (e. g., physicians, therapists, case managers) also constitute a


system of individual members who provide specialized services according
to changing patient needs over a course of treatment. Professional roles also
change, making it important for each team member to retain a perspective of
his or her role in the system.
Out of necessity, members of the professional team are in absolute control
during the initial phases of medical treatment, when they are charged with
saving the victim's life. Although team members may treat families compassionately, they may also see family members as "accessories" to the patient. The
professional team's duties and levels of accountability at this point leave little
room for compromise, and few family members have the knowledge or emotional strength to effectively address issues about the patient's condition or care.
As the patient's medical risk subsides and as restorative treatment begins,
the distinction between family members and professionals begins to blur:
Their roles meld and their importance ultimately becomes reversed. Early
rehabilitation efforts still require substantial professional treatment, with few
opportunities for the family's involvement. Eventually, new opportunities
arise and some family members may begin to take an active role, through their
own initiative or through the staff's. This may begin with participation in
everyday decision-making, by their taking an active role in treatment, or by
taking the patient home for brief visits. This is the point at which the treatment team must begin to fit its services into the family structure and begin to
relinquish control.
Transfer of control affects professionals and family members alike. As the
family members strive to assume and comprehend the size of the task that they

304

III. Rehabilitation Techniques

must undertake, the professional team must be able to give up what was once
an absolute level of control. Different team members will interact with the
family at different points of the treatment continuum. The neurosurgeon's
involvement may be early and brief, whereas psychosocial-services personnel
pick up the case in the neurology ward and follow it past the patient's discharge from formal rehabilitation.
Ultimately, it is important to remember that professionals enter into the
family by virtue of a catastrophe and will leave it at some point during its resolution. However, the patient's family will remain, although its form and
ability to work as a unit will have been redefined as a result of the pressures of
the entire ordeal. As professionals, we can help families in this process of redevelopment as a unit, but we must remember that we were "hired" to help
them with an overwhelming but specific problem [25]. We are visitors, rather
than members of the family.
9 A MODEL FOR FAMILY TRAINING

It is not realistic to expect that families and survivors can manage to fulfill all
of the survivors' continuing rehabilitation needs by themselves, although repeated research has noted that families assume this role by default once formal
rehabilitation has ended [40]. However, the training, resources, and perspective required for successful long-term service delivery are beyond the average
family's abilities. For example, consider the years of training required to develop proficiency in just one treatment area (e.g., neuropsychology, physical
therapy, or communication), let alone the many disciplines involved in most
treatment settings. Few families have this training prior to the injury, and
most cannot be expected to develop these skills after the injury. The level of
financial and logistical support required to meet continuing needs is also beyond
the budget of the average family.
However, despite scarce resources, multiple priorities, and few alternatives, family members of dependent TBI survivors become case managers and
service-delivery agents. It is incumbent upon professionals to help them become more effective in their roles and to help them break patterns of learned
helplessness. We can accomplish this by helping them to regain at least partial
control of the situation and by teaching them how to systematically address
small and manageable units of the overall problem [41].
At a minimum, families must begin to understand not only the complexity
of needs that the survivor faces but also the nature and course of long-term
recovery from TBI. As a next step, the concept of prioritization becomes important. In some cases, the family system may be able to provide identified
services; in other cases, families will need to get assistance from outside sources.
In still other cases, resolution may not be possible or necessary. Learning
how to "let go" of certain problems in order to prevent their exacerbation is
just as important as effective intervention. In the future, coordinated advocacy
among families at local levels will be crucial for the development of low-cost,
long-term service delivery networks.

12. Long-Term Family Intervention

305

A family "training" model (as compared to a family "therapy" model) may


address these issues. The family training model is oriented towards outcome
and problem-solving, in contrast to the process-oriented approach of more
traditional therapeutic models. It might ultimately be argued that both models
incorporate support as well as skill development and that the issue is more of
words than actions. However, the emphasis may provide an important distinction. By focusing on problem-solving techniques and outcomes, rather
than on problem identification and process, participants take a stronger role in
development and can more concretely measure their accomplishments by
the progress they make on specific issues. Obviously, support and guidance
throughout the training process are crucial, as they are in a therapeutic approach, but in a training model such support is likely to be more focused and
directed at helping participants reach predetermined points of success.
9.1 Education

Determining how to set priorities requires knowledge about the course of longterm recovery from head injury. In a series of training programs that we have
conducted for families oflong-term TBI survivors [40-42], we have observed
(and become concerned about) the patchwork of knowledge that many families
have developed. Most family members can identify the cause of their survivor's injuries, but few can localize the damage or know how it affects physical, cognitive, or behavioral issues. Knowledge about basic physiology (e.g.,
what nerve cells look like or how they work) and physical recovery is beyond
the scope of most people. Many families report having no formal training in
these areas or attribute what little knowledge they do have to meetings with
staff while the survivor was in an inpatient rehabilitation unit, during the first
weeks after the accident. Now, several years post-injury, after the survivor's
brain physiology has changed and the recovery process has progressed, many
families rely on outdated and incorrect information in making their decisions.
For example, we noted that a number of families still expected brain cells to
regrow five to ten years after the onset of the injury. They also thought that
brain-cell regrowth, as opposed to new learning, would continue to be the
basic process of skill restoration.
Families' knowledge about cognitive and behavioral processes is also limited.
Most families can state that the brain-injury survivor has problems with
memory, concentration, emotional lability, or other related issues, but they do
not understand the causes of these problems, how they interact, or how they
can be addressed. For example, a family may identify that a survivor has
memory problems, but they cannot distinguish short-term problems from
long-term problems, nor do they understand the meaning of this distinction
(e.g., "How come he can remember all of his old football games, but he can't
remember to come home on time?").
Similarly, relationships between different types of problems are also difficult
to address-for example, the interaction between distractibility and poor frustration tolerance. Although families may understand that brain-injury sur-

306

III. Rehabilitation Techniques

vivors are distracted and become easily frustrated, they may need assistance
in identifying the order of relationships between the two problems. The families may thus assume that the survivors would not forget what they were doing
if they could only quit losing their temper, when in fact their inability to concentrate drives their frustration! The answer in such cases does not lie in trying
to change the survivors' "will." It comes from helping families understand
more about cognitive deficits and how to address such issues.
These types of problems not only demonstrate a need for basic knowledge
but also indicate the importance of the amount and timeliness of the information presented. Family members are very busy and do not have large amounts
of time to spend searching for information. They need materials that are condensed and to-the-point, yet easy to understand. Most professional literature is
too dense and abstract for nonprofessional consumption and may not be appropriate to family needs. Basic knowledge about the process ofTBI recovery
and available treatment systems is essential to effective decision-making.
However, family time and comprehension are limited by both skill level and
other family life demands. Furthermore, each survivor of a traumatic brain
injury-along with his or her family-presents unique skills and deficits. For
all of these reasons, the most important discussions in any educational program
(after the initial training in the basic aspects oflong-term effects and recovery)
may be about how the families can find additional information when it is
needed, and how to digest it.
9.2 Problem Identification and Problem Selection

A second major problem is the number and diversity of problems that the
dependent survivor might face. A traumatic head injury is an instantaneous
and spontaneous event that can have drastic repercussions across almost every
aspect of life. Because so many of these problems are directly related to the
injury, there is often a natural assumption that all of these problems can be
solved with one solution. Unfortunately, interrelated problems rarely present
integrated solutions. A good analogy might be a piece of string that has been
dropped on the floor and that has become tangled. Although it took only one
act to create the chaos, it takes a good deal of time and patience to untangle the
string, one knot at a time.
Grouping and operationalizing problems (i.e., stating issues in terms of
goals and solutions rather than as problems) is often the most important task
that families have to address. Faced with a multitude of problems and with
limited resources, families must be able to select and prioritize those issues that
present the greatest need and for which solutions and resources are available.
Behavioral taxonomy (the reduction of complex skills into their discrete
behaviors) provides a good basis for teaching problem-grouping and operationalizing. By breaking down major problem areas into more discrete issues,
families can begin to focus on the actual components of the larger problemswhich they may also be capable of addressing.

12. Long-Term Family Intervention

307

Consider, for example, the problem of returning to work. Work consists of


a complex set of behaviors, encompassing technical skills, reliable attendance,
the ability to get along with others on the job, the ability to handle criticism,
the ability to have the rest of one's life well enough organized so as not to interfere with work, the obtaining of reliable transportation to and from the job,
and a multitude of other abilities.
Each of these specific abilities can be further broken down into smaller units
whenever it is relevant to the presenting problem. For example, if attendance
and reliable transportation pose no problems, they do not have to be further
dissected. On the other hand, if accepting criticism is a problem for the individual, this area might be broken down to identify the precise issues of concern.
For example, it might be helpful to know if the issue involves the person's
accepting criticism from coworkers or employers, accepting corrective feedback (as compared to negative feedback), accepting criticism when learning a
new task (as compared to one that the person is supposed to know), and so on.
Smaller and smaller analyses of the problem situation can be developed until an
addressable unit of behavior is isolated for intervention.
Two problems will generally interfere with such a focused task analysis.
First, it may be difficult for family members to accept working towards relatively small gains, given the multitude and complexity of the problems being
faced. Second, the problems and priorities that the family faces are dynamic
rather than static. Frequently, just after a specific problem has been selected
and articulated, a new challenge may spring up, capturing everyone's attention
and distracting the family from the work that has just been accomplished.
These two issues can best be managed by selecting topical challenges that are
central to long-standing problems of the survivor and the family; by making
sure that there are adequate resources within the family to address spurious
issues as they arise, without jeopardizing their other systematic efforts; and by
providing firm but supportive direction to keep family members on track.
9.3 Resource Assessment

Resources involve much more than money to pay for daily living expenses and
rehabilitation services. Many families possess significant resources that may be
hidden or not actively considered; we often call such resources "in-kind."
These may include time that each family member can or will spend on survivors' needs; the access that members have to existing community services;
the influence that they may have to encourage the development of new services
in the community; individual skills that family members may have, according
to their profession or experience; their ability to make changes in daily routines
to meet treatment or care needs; and their ability to share family resources with
one another.
One of the biggest challenges in long-term recovery is the fact that at some
point after returning home the survivor of the traumatic brain injury no longer
receives the same high level of family priorities and resources that may have

308

III. Rehabilitation Techniques

been available while he or she was in the hospital. As time goes on, other family
members' priorities must be considered again. To balance the equation of
resources and needs within the family, the survivor's needs must meld with
issues that arise. For example, the breadwinners of the family, who might have
been able to take off some time immediately after the injury, need to return to
work. Children in the family need support and guidance during their development. Strained marital relations need to be addressed. Life must go on.
In ongoing family training groups, we find that family members are willing
to learn and implement new techniques. However, like most of us, they are
pressed for time. For example, in one family, the parents had let their small
business fall into financial disarray during the initial months of the medical
catastrophe. After a year had elapsed, they needed to devote an extensive
amount of time to the business so that the family could continue to have an
income. In another family, a sister who took care of a survivor needed to return
to college. A survivor who was a high-powered lawyer prior to her injury
found that the amount of time she had to spend on legal issues and on reassigning cases to other colleagues took substantial time away from her own rehabilitation programming needs. Like other survivors, she found that the return
to the community and family brings new needs and new responsibilities.
To comprehend the significance of in-kind resources and manage all their
resources, many families need to understand how their family unit functions as
a system. This concept may seem obvious to many family members. However,
charting each family member's needs and contributions to the family unit helps
to clearly delineate resources-as well as the priorities of the survivor, relative
to other family members. This process also helps provide a more objective
perspective for targeting problems and formulating interventions.
9.4 Intervention

To move from a dependent role to an active role in long-term management of


the survivor's care, families have to make decisions based on their knowledge,
assessment of their resources, and identification of the presenting problems.
Perhaps the most important factor in the process of resource assessment is
ranking the survivor's presenting issues according to type, severity, and
resource needs. As has been previously noted, some problems require immediate professional attention and are beyond the scope of family ability (e.g.,
medical issues, specialized training). Others are long-term issues that can wait
or that may not even be appropriate to address at the present time (e.g., trying to get the person to go back to work before major aggressive behavior
problems are under control).
Discussions about problems that require outside resources and treatment
programs usually end at the decision-making process, after the problem has
been targeted, resources have been identified, and services have been obtained.
This process involves finding appropriate therapy programs, schools, and

12. Long-Term Family Intervention

309

professional services. On the other hand, many of the issues faced by survivors
and their families reflect daily life and can therefore be addressed within the
family structure. Problems that revolve around day-to-day issues in the home
can frequently be improved or resolved through the application of behavior
analysis techniques.
As noted earlier, behavior analysis and behavioral family training techniques
have made strong contributions to other populations, including the developmentally disabled and psychiatrically impaired. They have also been effective
with children and in marital/family situations. Although these techniques are
obviously not a panacea for all issues that the family of a brain-injury survivor
faces, they are effective in addressing those issues-including skill building,
communication, and behavioral control-in which the daily environment has
an influence on individual actions.
A popular notion is that behavioral techniques emphasize motivational control. However, behavior analysis more correctly focuses on environmental
relationships and behavior (including situational cues and antecedents that occur
before the behavior), the topography (form) of the desired behavior, and
control of consequences.
The following figure (Figure 12-1) presents a flowchart for behavior change
that is frequently used with family members and professionals alike in teaching
basic behavior-analysis strategies. It is immediately obvious that the majority
of the flowchart is devoted to problem specification and assessment, rather than
intervention. Once the precise relationships between environmental events
and behavior are understood, intervention is often "simply" a matter of
manipulating these functional relations.
Most of the work of stating and operationalizing the problems and goals will
have been accomplished during problem-identification training. However, it
is important that all parties involved in the intervention, including the survivor,
agree on these issues before continuing into the flowchart. The greatest cause
of failure for any intervention is either imprecision or lack of agreement about
the goals and objectives.
Because behavior analysis interventions are based on the manipulation of
functional relationships between controlling variables and behavior, the identification of the controlling variables in the equation is crucial to successful outcomes. A significant problem is that many controlling variables may be responsible for any specific behavior, and some of these variables may not be in
the realm of either family or professional intervention. For example, a person
with significant frontal lobe involvement may be highly impulsive, but repair
of frontal lobe damage is currently beyond our abilities. However, other
variables such as the amount of environmental stimulation, the size of demands, the cues and support for staying "on task," and the reinforcement for
each step accomplished are within reach and can be highly effective in helping
to manage the situation.

310

III. Rehabilitation Techniques

STATE THE PROBLEM AND


GOAL IN OPERATIONAL
BEHAVIORAL TERMS

Figure 12-1. Flow chart for behavior change.

One successful strategy is to ask families to identify two sets of controlling


variables-those variables that are outside the realm of control (such as neurological damage and medical limitations) and those environmental variables
that are more amenable to intervention. Generally, this second set can be further
delineated into antecedent events, which occur prior to the behavior of interest,
and consequences, which follow the behavior. (Again, as is indicated by the
flowchart, all parties need to agree about these variables before any data are
collected; this will ensure consistency in the information to be collected.)
Data collection has become synonymous with behavior analysis and provides
important information even before an intervention is established. Incorrect
estimation of the frequency and magnitude of the target behavior also frequently
occurs due to the situational aspects of the behavior. For example, people will

12. Long-Term Family Intervention

311

often overestimate the frequency of a specific target behavior because of its


magnitude when it occurs-as in the case of a singular but dramatic outburst
of frustration. Similarly, less dramatic behaviors may be underestimated, even
though data collection may demonstrate a much higher frequency of occurrence. Following data collection, it is not unusual for families and therapists to
have to reassess the significance of the survivor's target behavior, after they see
the discrepancies between what they assume to be the behavior and the actual
presentation of the behavior. In some cases, people will determine that the problem is not of sufficient magnitude to address, and they will therefore select
another problem. In other situations, the overall problem for which a specific
behavior was targeted for intervention will still be perceived to exist, and it
will be necessary to start back at the beginning of the flowchart by redefining
the problem and the goal.
The hypothetical identification of variables that control behavior and the
documentation of the actual relationships of those variables to behavior constitute two separate issues. By collecting data about selected controlling variables
in conjunction with the occurrence of the behavior, one can gain an empirical
perspective about functional relations between these two issues. Discrepancies
between assumed controlling variables and those that are actually recorded can
then be alleviated by redefining the noted controlling variables and assessing
their amenability to intervention. Formal intervention cannot be pursued until
after the behaviors and goals have been operationally defined, the controlling
variables have been documented, and the functional relationships between the
two have been empirically determined.
After the careful assessment of behavioral and environmental relationships
has been completed, most interventions will consist of rearranging the environmental and interpersonal interactions to foster behavioral change. In many cases,
it may be another family member, rather than the survivor, who will have to
make the adjustments. For example, in one case, a mother complained that her
son yelled at her or would stomp off whenever she tried to talk to him. She
soon realized that this mainly occurred when she tried to correct him. When
she took a different approach in her corrective feedback, she helped to reduce
family tensions. Another survivor, who had previously been reported to be
irresponsible, was noted to have significant memory problems and to lose track
of her responsibilities when she was asked to do more than one thing at a time.
These problems were reduced after interventions helped her rearrange her tasks
and cues for her responsibilities.
In other situations, survivors may be asked to make significant changes in
behavior and must participate in planning and carrying out the intervention.
Families learn quickly that they cannot make anyone "do" something, but that,
through careful communication, they can work together to attain mutually
agreed-upon goals.
Training in basic principles of reinforcement, stimulus control, and shaping
techniques also proves helpful in situations of motivational control, direction,

312

III. Rehabilitation Techniques

and new learning. Most families can begin with simple applications of these
procedures, especially when they are applied to well-specified issues. Initial
training should focus on the acquisition, development, and use of behavioral
principles; at this point, proper technique is considered to be more important
than the magnitude of change acquired. Later, larger and more complex issues
can be addressed as the family members' proficiency improves.
Regardless of the type of intervention attempted, it is essential that collection of data about controlling variables and behaviors continue throughout the
intervention, so that outcome can be monitored. This is frequently the most
difficult step for family members to adhere to. Once they notice the beginnings
of a behavior change, they are likely to stop recording data and rely on their intuition instead. However, behavior change is a dynamic process, and its course
may alter or even reverse during the process of intervention.
Without an empirical base from which to monitor progress, one cannot be
certain if initial changes are being maintained, or even if the functional relationship between the noted change in behavior and intervention is continuing. For
example, in one case, a mother developed a behavioral program to help increase
her daughter's self-care and appearance skills. The program began successfully,
and the mother was astounded by the rapid changes. Several weeks later, the
daughter returned to her former disheveled appearance, even though the behavioral program was still in operation. The mother was beside herself. An
investigation of the situation showed that, at the same time that the program
was being implemented, the daughter had found a boyfriend and had developed a relationship that had lasted several weeks. Hence, the improvement in
the patient was primarily a result of her having the boyfriend; the mother's
intervention had little power.
When interventions prove unsuccessful, it is time to reassess the situation,
behavior, and controlling variables, and then plan a new strategy. The length
of time one should pursue a specific program depends upon intervention goals,
the person's baseline behavior at the start of the intervention, and the severity
or importance of the problem.
In most cases, interventions that prove successful must be maintained to
facilitate continued behavior change, and this proves to be another stumbling
block for continued success. It is crucial to remember that behavior is a function of its environment and is responsive to the antecedent stimuli and consequences of its occurrence. A family's reversion to old patterns of interaction
following a change in the survivor's behavior can bring about a return to previous forms of behavior. For this reason, programming for maintenance and
generalization is important.
The level or intensity of programming that is required to maintain behavior
change will generally not be as great as the effort required to bring about the
initial change. However, for continued success, long-term planning is just as
crucial as the initial intervention. Over time, cues, reinforcement schedules,
interaction patterns, and the general specificity of the intervention program

12. Long-Term Family Intervention

313

can fade as the problem subsides and as newly developed patterns of responding become more "natural. " The amount of time required for such a transition
will depend on the rate of behavior change and the durability of such change,
both of which will be shown by the continuing collection of data.
As their interventions succeed, families come to understand that they are not
helpless to address the multitude of issues facing them. As a result, learned helplessness begins to subside, and family members can take on larger and larger
issues with greater resilience and with a greater opportunity for positive
outcome.
9.5 Advocacy

Ultimately, families will find that many issues are outside of their realm of
control and that no alternative community or professional services are available. In fact, the development of a comprehensive and responsive system of
services for this population may come only after families learn to organize and
become advocates for such services. The efforts for achieving national advocacy
that are being undertaken by the National Head Injury Foundation (NHIF) and
professional task forces on head-trauma rehabilitation are meeting with increasing success. However, local efforts by families and professionals must be
intensified, both to take advantage of these national developments and to develop
unique services to meet local needs. For example, the recently signed cooperative
agreement between the NHIF and the Office of Special Education and Rehabilitation Services (OSERS) mandates vocational services for TBI survivors
through state vocational rehabilitation agencies. It is up to each state, however,
to formulate and implement these plans.
Needs for community activities, socialization centers, and support groups
are local issues that require community efforts. However, TBI rehabilitation is
not the only issue that communities face, and active lobbying, program development, and advocacy by those affected by traumatic brain injury are the only
means by which any such services will be developed.
At support group meetings, families are often challenged to justify why their
needs are more important than any other social issue that their community faces
and why anyone else should even get involved in their problems. At first, families are shocked when their needs are so belittled, but often they are then asked
what they have done to help others: "How many of you have attended an
American Cancer Society meeting in the past year or worked for the rights of
people with psychiatric disabilities?" Through such interactions, it becomes
clear that all of us in society have our individual needs, that we are our own
best advocates, and that we cannot expect others to prioritize our burdens over
their own. We can, however, educate others about our needs and ask them for
their assistance in helping us achieve our goals.
Family advocacy follows out of problem-solving and resource assessment.
Effective local problem-solving support groups form the foundation for a sense
of community and ultimately for social advocacy. The increasingly strong

314

III. Rehabilitation Techniques

national base provided by the NHIF and its state chapters is a source of direction and inspiration. However, as individul families identify priorities within
their own homes, they must work with other families who have similar problems in order to make their priorities a community issue-that is, to gain
access to community resources.
10 SUMMARY

The long-term issues that families and TBI survivors face are dramatic, dynamic, and pervasive. Many individual components of these issues have been
previously noted and faced by other groups of people who have had catastrophic illness or who are chronically disabled. However, it is the suddeness of
the injury, the survivor's age (generally young) when the injury occurs, and
the pervasive and dynamic long-term needs of the dependent survivor, as well
as the established nature of existing TEl rehabilitation systems, that make this
population unique and can place families in states of learned helplessness.
Although families may not be equipped to meet-or may not be capable of
meeting-the complex long-term needs of this population, they have assumed
the role by default, due to a lack of other available services. It is not possible to
expect families to meet all of the long-term needs of the survivors, but they
can meet some, and, with proper training and assistance, they can become
more effective in securing other services. Pragmatically, families can become
more effective case managers when they have had proper training in long-term
TBI sequelae, problem identification, problem solving, resource utilization,
and basic, environmentally oriented interventions. Initial intervention must
focus on developing effective management skills, with later efforts directed
toward progressively larger problems. If families are capable of assuming some
of these responsibilities, they may also break the barrier oflearned helplessness,
thereby becoming stronger advocates for their needs.
Ultimately, for a comprehensive service-delivery network to be established,
local family groups must become advocates for services within their own communities. Advocacy forms naturally out of groups of families who can focus
on specific mutual needs and can systematically gain support from others, but
families cannot expect anyone else to "carry the torch" for them.
Professionals, of course, can-and must-play an integral role in awareness,
cooperation, and service dc.livery with families, in a team effort. Professionals
assume most of the responsibility for treatment during the initial stages following the injury, but gradually fade from direct services over time. However,
their overall effectiveness will be measured in part by their ability to help
survlvors and their families "carry the torch" once formal treatment has
ended.
REFERENCES

J. Neurosurg. Nurs.
13, 165-169.
2. Jacobs, H.E. (1987). The Los Angeles head injury survey: Project rationale and design impli1. Mauss-Clum, N. and Ryan, M.R. (1981). Brain injury and the family.

12. Long-Term Family Intervention

315

cations. j. Head Trauma Rehabi!. 2, 37-50.


3. Jacobs, H.E. (1988). The Los Angeles head injury survey: Procedures and preliminary findings. Arch. Phys. Med. Rehabi!. 69, 425-431.
4. Najenson, T., Grosswasser, Z., Mendelson, L. and Hackett, P. 1980. Rehabilitation outcome
of brain damaged patients after severe head injury. Internat. Rehabi!. Med. 2, 17-22.
5. Oddy, M., Humphrey, M. and Uttley, D. (1984). Stresses upon the relatives of head-injured
patients. Br. j. Psychiatry 133, 507-513.
6. Thomsen, I. V. (1974). The patient with severe head injury and his family. Scand. j. Rehabi!.
Med. 6, 180-183.
7. Weddell, R., Oddy, M. and Jenkins, D. (1980). Social adjustment after rehabilitation: A two
year follow-up of patients with severe head injury. Psycho!. Med. 10, 257-263.
8. Lezak, M.D. (1978). Living with the ch3racterologically altered brain injured patient. J. Clin.
Psychiatry 39, 592-598.
9. Mace, N.L. and Rabins, P.V. (1982). The 36 Hour Day: A Family Guide to Carillgfora Persoll
with Disease, Related Demelltillg Illness and Memory Loss in Later Life. Johns Hopkins Press,
Baltimore.
10. Zarit, S.H. and Zarit, J. M. (1982). Families under stress: Interventions for caregivers of senile
dementia patients. Psychotherapy: Theory, Res. & Pract. 19, 461-471.
11. Liberman, R.P., Wheeler, E.G., de Visser, L.A. J.M., Kuehnel, J. and Kuehnel, T. (1980).
Handbook 4 Marital Therapy: A Positive Approach to Helping Troubled Relationships. Plenum
Press, New York.
12. Becker, W.C (1971). Pam/ts Are Teachers, Research Press, Champaign, IL.
13. Masten, A.S. (1979). Family therapy as a treatment for children: A critical review of outCome
research. Fam. Process 18, 323-336.
14. Steinglass, P. (1976). Experimenting with family treatment approaches to alcoholism, 19501975: A review. Fam. Process 15, 97-124.
15. Stanton, M.D. (1979). Family treatment approaches to drug abuse problems: A review. Fam.
Process 18, 251-280.
16. Falloon, I.R.H., Boyd, J.L., McGill, CW., Razani, j., Moss, H.B. and Gilderman, A.M.
(1982). Family management in the prevention of exacerbations of schizophrenia: A controlled
study. N. Eng!. J. Med. 306,1437-1440.
17. Ktibler-Ross, E. (1969). Orl Death and Dyillg. Macmillan and Company, New York.
18. Fink, S. 1967. Crisis and motivation: a theoretical model. Arch. Phys. Med. Rehabi!. 48,
592-597.
19. Pearson, J. (1973). Behavioral aspects of Huntington's chorea. Adv. Neuro!. 1,701-712.
20. Weisman, A. (1972). On Death alld Dying. Behavioral Publications, New York.
21. Matson, J.L. and McCartney, J.R., eds. (1981). Handbook of Behavior Modificatioll with the
Mentally Retarded. Plenum Press, New York.
22. Ruegamer, L.C, Kroth, R., and Wagonseller, B.R. (1982). Public Law 94-142: Putting Good
bltentiolls to Work. Research Press, Champaign, IL.
23. Jacobs, H.E., Donahoe, CP. and Falloon, I.R.H. (1985). Rehabilitation of the chronic
schizophrenic: Areas of intervention. In All/lIIal Review of Rehabilitatioll, Vol. IV, Pan, E.,
Backer, T., Vash, C and Newman, S., eds., Springer Publishing Compa.ny, New York,
pp.83-113.
24. Muir, CR. and Haffey, W.j. (1984). Psychological and neuropsychological interventions in
the mobile mourning process. In Behavioral Assessment mid Rehabilitation of the Traumatically
Brain Damaged. Edelstein, B.A. and Couture, E.T. eds., Plenum Press, New York, pp.
247-272.
25. Jacobs, H.E., Muir, C.A. and Cline, J. (1986). Family reactions to persistent vegetative state.
J. Head Trauma Rehabi!. 1, 55-62.
26. Lezak, M.D. (1967). Psychological implications of traumatic brain damage for the patient's
family. Rehabi!. Psycho!. 31, 241-250.
27. Seligman, M.E.P. (1975). Helplessness: 011 Depression, Developmellt and Death. W.H. Freeman
and Company, San Francisco.
28. National Head Injury Foundation. (1988). National Directory of Head Irljury Rehabilitation
Services, NHIF, Southborough, MA.
29. Power, P.W. and Dell Orto, A.E., eds. (1980). Role of the Family in the Rehabilitation of the
Physically Disabled. University Park Press, Baltimore.
30. Wright, G.N. (1980). Total Rehabilitatioll. Little, Brown, Boston.

316

III. Rehabilitation Techniques

31. Galloway, J.P. and Goldstein, H.K. (1971). A Follow-lip Study of the Illfluellces of Group
Therapy with Relatives 011 the Rehabilitatioll Potelltial of Rehabilitatioll Clicl/ts. (Unpublished
manuscript, Delgado Community College, New Orleans, LA.)
32. Lindenberg, RE. (1977). Work with families in rehabilitation. Rehabil. Counseling Bull. 20,
67-76.
33. Neff, W.R. (1959). Succcss of a Rehabilitatioll Program: A Follow-up Study of the Vocatiollal
Adiustmellt Cel/ter. MOI/Ograph 3, Jewish Vocational Service, Chicago.
34. Olshansky, S. and Beach, D. (1975). Special report. Rehabil. Lit. 36, 251-253.
35. Falloon, I.R.H. and Liberman, R.P. (1983). Behavioral therapy for families with child
management problems. In Helpillg Families with Special Problems, Textor, M. Red., Jason
Aronson, New York, pp. 121-147.
36. Patterson, G.R. (1975). Families: Applicatiolls of Social Leamillg to Family Life. Research Press,
Champaign, II..
37. Sbordone, R. (1983). The emotional reaction of family members of head injured patients.
Paper presented at the International Traumatic Head Injury Conference, London.
38. Grief, E. and Matarazzo, R.G. (1982). Behavioral Approaches to Rehabilitatio//: Copillg with
ChaT/ge. Springer Publishing Company, New York.
39. Gagnon, R (1984). The AARC: A Model.for Effective LOII.~ Term Carefor the Traumatically Head
Illiured. (Unpublished manuscript, American Head Trauma Alliance, Los Angeles.)
40. Jacobs, H.E. (1986). Patient and family outcomes from traumatic head injury: Treatment
roles for long-term sequelae. Workshop presented at the Seventh Annual Braintree Traumatic
Head Injury Conference, Braintree, MA.
41. Jacobs, H.E., Muir, C and Wixom, C (1986). Assessment and intervention in family
training. Workshop presented at the Fifth Annual National Symposium, National Head
Injury Foundation, Chicago.
42. Muir, C, Jacobs, H.E. and Martel, M. (1987). Family training. Workshop presented at the
Sixth Annual National Symposium, National Head Injury Foundation, San Diego.

13. MANAGEMENT OF AGGRESSIVE BEHAVIOR


FOLLOWING TRAUMATIC BRAIN INJURY

WILLIAM J. HAFFEY AND JOHN W. SCIBAK

1 INTRODUCTION

Survivors of traumatic brain injury (TEl) often exhibit aggressive behaviors


that can frighten, alienate, intimidate, or place stress upon their caregivers,
treatment personnel, and family members. Such aggression presents a formidable barrier to the TBl survivor's resumption of preinjury activities, roles, and
status. The challenge to rehabilitation professionals and families is to discover
ways of interacting with these TBl survivors that will promote learning of the
prosocial behaviors necessary to support community resettlement.
1.1 Case Illustration: Mark

Mark was 19 years old when he was struck by a passing truck as he was walking along a freeway breakdown lane. This 6'2", 220-pound college football
player, who was comatose for two weeks following his accident, was admitted
to an inpatient rehabilitation program one month after his TBl. During the
first week in this program, he was discovered in an agitated state. He had
managed to climb onto his bed while still strapped in his wheelchair. He was
attempting to jump from his bed through the window. He was so agitated that
physical restraint measures were necessary to prevent him from injuring
himself. During these interventions, he punched one nurse and kicked
another.
Preparation of this manuscript was funded by the New Medico Head Injury System. The authors wish to thank
Eileen Haffey and jean Langevin for their editorial assistance and clerical support. Special thanks arc due to Peter
Eames, MD, for his critical review of an earlier version of this chapter.

317

318

III. Rehabilitation Techniques

1.2 Case Illustration: Jack

Jack was a 5'8", 170-pound high school wrestler and motocross racer. He sustained a severe closed-head injury when he lost control of his motorcycle when
he was riding off-road in the desert. He enrolled in a residential community reentry program for brain-injury survivors 18 months after his injury.
Jack's behavior was typically characterized by psychomotor and emotional
agitation. During the initial weeks in the program, he would become extremely agitated whenever he was frustrated. Sometimes this frustration would
result in diffusely directed tirades. At other times, his obscene invectives would
be directed at a staff member or another resident. In these latter situations, he
would be extremely belligerent and would often approach the person in a
physically threatening manner. At times when his dense left homonymous
hemianopsia and unilateral neglect [1] impaired his performance, he would
hurl objects across the room or attempt to destroy these materials. He would
then leave the scene in such an agitated state that he sometimes crashed into
people or objects. This would then elicit another round of verbal abuse and
psychomotor agitation. Once Jack calmed down, however, he would act as if
nothing out of the ordinary had occurred.

1.3 Case Illustration: Al

Al was a 29-year-old mechanic who was injured when a bus tire exploded next
to him. The impact of the explosion threw him across the garage. He struck
his head against a hydraulic lift when he landed. He had a severe closed-head
injury complicated by the concussive effects of the explosion. One result of his
severe brain injury was a total loss of hearing.
After his discharge from an inpatient brain-injury rehabilitation program,
Al returned home to his wife and two young daughters, who lived in a
rural area. During the next year, Al and his family relied on educational and
counseling services that were readily available in their local area, because the
rehabilitation center was about 60 miles away.
Approximately one year after Al returned home, his wife called the rehabilitation center. She was very distressed. She complained that she was at her wit's
end due to her husband's extreme jealousy, his violent temper outbursts, his
threats of assault, and his manhandling of his daughters when he was attempting to reestablish himself as the family disciplinarian. The event that prompted
her to call the rehabilitation psychologist for help was AI's physical threat to
attack her with a butcher knife. When she ordered him out of the family home,
he turned the knife on himself, screaming that he might as well kill himself,
since there was no place for him even in his own home. His wife convinced
him to give her the butcher knife. Al then spent the rest of the evening in
the cab of his pickup truck. The next day, Al and his wife began outpatient
counseling with the rehabilitation psychologist.

13. Management of Aggressive Behavior

319

1.4 Case Illustration: Mary

Mary was a 33-year-old woman who sustained a severe head injury in a singlecar accident. Her injuries also resulted in serious facial scarring, and a residual
left hemiparesis required her to use a leg brace and a quadcane. Despite her
hemiparesis, she was an imposing physical presence, due to her large bone
structure, her height (6'), her facial demeanor, and the way she acted towards
others.
During the initial month in a residential community reentry program for
brain-injury survivors, Mary either isolated herself or became embroiled in
conflicts with the staff and other residents. She was alternately melancholic or
agitated. Contributing factors included her depression over the breakup of a
long-term lesbian relationship just prior to her accident; her fears that she
would not be able to attract another lover because of her facial scarring and
residual hemiparesis; her fears that she would not be able to resume her occupation as a courier; her anger at her dependency on her elderly parents (who
she felt blamed all of her life's misfortunes on her sexual preference); and her
verbal and physical aggression towards anyone whom she perceived as invading her personal space. She had been hospitalized for depression prior
to her head injury. In each instance, the precipitating circumstance was the
dissolution of a sexual relationship.
At one point during the initial weeks of her stay in the rehabilitation program,
Mary threw a brick at a staff person during a disagreement. On another occasion, she had left the living room to go to her bedroom during an argument
with a staff member. The staff member pursued her down the hallway, at
which point Mary turned and struck the staff member with her quadcane.
1.5 Scope of the Chapter

This chapter will examine two fundamental questions:

1. What conditions give rise to and maintain the TBI survivor's aggressIve
responses?
2. What conditions will most likely gIve nse to and maintain alternative
prosocial responses?
Within each of these questions, we will examine factors within the TBl
survivor and factors associated with his or her interaction with the environment.
This examination will include an analysis of the contributions of preinjury and
postinjury conditions. The framework for this chapter follows from our conviction that the design of treatment interventions requires a comprehensive
assessment of the TBI survivor's aggression. This assessment should yield data
that enable treatment personnel to describe the nature of the individual's
aggression, identify probable underlying reasons for the aggression, and select
interventions that are specifically targeted at the determinants of the aggression.

320

III. Rehabilitation Techniques

The design of such a data-based intervention plan requires a conceptual


understanding of the nature and determinants of human aggression. Consequently, we will begin this chapter with an examination of the implications of
adopting a social-science perspective about aggression. This is in contrast to an
ethological, biological, or psychoanalytic perspective. We also believe that a
comprehensive management strategy should be guided by a theoretical construct of human learning. Therefore, we will discuss the implications of social
learning theory for the management of the TBI survivor's aggression.
2 HUMAN AGGRESSION

2.1 Definition of Aggression

Bandura [2] defines aggression as "injurious and destructive behavior that is


socially defined as aggressive" (p. 8). Tedeschi, Smith, and Brown [3] have
stated that aggression is the exercise of coercive power that "refers to the use
of threats and punishments to gain compliance" (p. 548). They also emphasize
that "no action can be identified as aggressive or violent without taking into
account the value system of the perceiver" (p. 557).
This social science perspective about aggression implies that:
1. The aggressor is seeking to control a situation so that his or her perceived
needs are met.
2. Acting aggressively is problematic not because it fails to meet the aggressor's needs, but rather because people in the environment judge that the
behavior violates accepted norms for that situation.
3. Behavior must meet the individual's perceived needs and respect situational
norms if it is to be judged adaptive (prosocial) by both the aggressor and
others in the environment.
These implications provide the first insight into what should be involved in
assessing the TBI survivor's aggression. Such an evaluation should attempt to
identify 1) the TBI survivor's perceived needs, 2) the situational norms that
were violated, 3) the survivor's awareness of these norms, and 4) the potential
alternative behaviors that could satisfy the needs of both the brain-injured
person and others in the environment.
2.2 Goal-Directed Nature of Aggression

A social-science perspective holds that aggression is purposeful [4]. The


aggressor intends to accomplish a goal by acting aggressively. The maladaptive-or adaptive-quality of the behavior is often determined by how others
in the environment judge the aggressor's intentions. Some researchers make a
distinction between two types of aggression, which are differentiated by their
behavioral characteristics [5-9]. Overt, hostile aggression is typically emotionally charged and most often involves behavior that is excessive with respect to
situational norms. This emotional aggression is qualitatively different from instru-

13. Management of Aggressive Behavior

321

mental aggression. This latter type of aggression characteristically involves the


use of power or force to control a situation. Emotional aggression is most
often a response to aversive stimulation, whereas instrumental aggression
is most often motivated by external incentives such as food, attention, and
material objects.
The importance of this distinction is that the TBI survivors' use of aggression can be examined in the context of the incentives they are trying to obtain
and in the situations they are trying to escape or avoid. Such an analysis can
yield valuable information relative to identifying prosocial (situationally
adaptive) alternative strategies that could be substituted for the currently
employed maladaptive strategies.
The insight that behavior is goal-directed is also valuable because it overcomes the pessimism that follows from the assumption that aggressive behavior
is instinctual. Bornstein and his colleagues [10] have stated that the early psychoanalytic theories of Freud [11] and the ethological theories of Lorenz [12]
inhibited "the development and empirical validation of effective treatment
strategies for the control of anger and aggression" (p. 603) [10]. Berkowitz
[13] concludes that "few research-oriented zoologists and psychologists now
subscribe" to the instinctual theory of aggression (p. 355).
Social learning theorists maintain that aggressive behavior is learned [2, 4,
13]. Although aggressive tendencies exist to support the survival of the individual, the expression of aggressive behavior is determined by learning. Initially, people acquire aggressive behaviors through instrumental learning (i. e.,
direct response) and vicarious learning (i.e., observing others). When someone
behaves aggressively, consequences (e.g., reinforcement or punishment) determine the probability of that behavior's recurrence in similar situations. "Reinforcing" consequences serve to maintain the acquired aggressive behavior;
"punishing" consequences suppress the behavior. The significance of this
social learning approach to aggressive behavior is that aggression can be
modified using social learning principles and techniques [2].
2.3 Stimulus Conditions Associated with Aggression

Valuable insights can be gained from an analysis of situations in which aggressive tendencies are typically evoked. When people who are engaged in a goaldirected behavior cannot achieve the desired goal, they usually become aroused
emotionally. Most people refer to this arousal as "frustration." The particular
behavioral response employed to deal with these feelings of frustration is a
function of experiential learning. For example, some individuals might expend
more effort or might select an alternative task approach. Either of these responses could lead to the desired goal. In this sense, frustration can lead to
adaptive responses. However, people can also respond to frustration maladaptively-for example, by ceasing to attempt to achieve the goal. If that
goal is critical to their welfare, such a response fails to meet essential needs. If
people react in a hostile manner (affective aggression), they are considered to
be behaving maladaptively, even if their perceived needs are met.

322

III. Rehabilitation Techniques

It is important to emphasize that the thwarting of goal-directed behavior


evokes aggressive tendencies. A person's behavior in response to those tendencies is determined by previous learning and can be considered either adaptive or maladaptive, depending on its appropriateness. Bandura [2] noted that
training can influence whether a person behaves aggressively or appropriately
when he or she experiences frustration. The TBI survivor who behaves aggressively can be trained to behave adaptively if the conditions that affect learning
are managed effectively.
Because TBI survivors often cannot perform tasks as effectively or efficiently as they could prior to the injury, they frequently experience frustration
throughout the course of recovery. Specific training in dealing with frustration
should be a central aspect of rehabilitative training. This will increase the likelihood that TBI survivors will respond adaptively rather than maladaptively.
This is especially true when a TBI survivor has succeeded in using aggression
to either obtain social attention or to escape or avoid frustrating situations.
Fear is another experience that evokes aggressive tendencies. People experience fear when they perceive a threat to their safety or well-being. This fear
gives rise to any number of responses, including aggression. For example,
Mark (see Section 1.1) believed that he was in prison. He thought that his
brother had parked his pick-up truck next to his bedroom window and was
signalling him to jump into the truck bed to effect a prison break. When the
nurses attempted to keep him from escaping, he felt trapped. His aggression
was related to both the thwarting of his goal-directed behavior and to his fears
about remaining in prison. We have seen other TBI survivors who incorrectly
interpret an environmental situation asa threat and respond in a combative
way to protect themselves from this perceived threat.
Human beings, like other animals, are very protective of their territory.
Threats to that territory can evoke aggressive tendencies. For example, Mary's
aggression (see Section 1.4) was sometimes linked to perceived threats to her
personal territorial boundaries. Interaction with people in general, and heterosexuals in particular, was aversive to her. She responded by isolating herself
whenever she could. Whenever circumstances required that she interact with
others, she would communicate through her behavior that she did not want
her personal space "invaded. " She had to be in control of how someone entered
her private territory. If a person didn't understand her signals or chose to ignore
them, Mary would become verbally aggressive in the hope that she could drive
them away. On those occasiops when the person would persist, she threatened
violence; and on at least two occasions she had become involved in physical
altercations with people who "wouldn't stop bugging" her. Subsequent to her
injury, the incidence of this type of territorial aggression had markedly increased, but in all cases it was restricted to her making verbal threats to harm
others.
Sex-related aggression is another common human experience; it often involves territorial aggression, as well. For example, one role that was pivotal to

13. Management of Aggressive Behavior

323

Al's (Section 1.3) psychological identity and well-being was that of husband
and father. He was very protective of the territory of his family and home.
Prior to his injury, he was suspicious of interactions between his wife and
other men. Occasionally, he would express his aggressive tendencies in verbal
attacks when he believed that his wife was having sexual relations with other
men (sex-related aggression). When he was drinking, stresses in their relationship that would otherwise be ignored could trigger a verbal assault about her
(supposed) unfaithfulness (irritable aggression). After the injury, which rendered Al functionally deaf, his frustration at not being able to hear what his
wife \-vas saying on the phone (an aversive stimulus) triggered verbal assaults
about her supposed infidelity that eventually escalated to threats of physical
harm to her and to her imagined lover(s).
We just pointed out that the probability of AI's becoming verbally abusive
increased whenever he was drinking. In addition, many other conditions or
"states" within an individual can reduce tolerance in annoying or irritating
situations. These internal states can increase the probability that a person will
react in an overtly aggressive manner. The physiological and psychological
effects of alcohol and drug intake, fatigue, pain, sleep deprivation, anxiety,
depression, or stress each influence the extent to which a stimulus will be
considered aversive and the probability that aggression will occur [10].
We have discussed a variety of stimulus conditions that tend to evoke aggressive tendencies. Since TBI survivors experience many of these conditions
during the course of their recovery, it is important to review each individual's
postinjury daily activity pattern. Such a review can reveal stimulus conditions
in which aggressive tendencies were evoked and can identify those situations
that were usually associated with aggressive behavior. Such an analysis is valuable on two levels-diagnosis (it may provide information about the structure
of aggressive incidents that have already occurred) as well as prophylaxis (it
may help identify, for the future, situations in which aggression might occur).
3 SOCIAL LEARNING PERSPECTIVE

Experiments employing classic and operant conditioning paradigms have


demonstrated that the behavior of an organism that is capable of learning is
determined by the organism's interaction with environmental events [14-15].
3.1 Antecedents

One variable in the person-environment interaction that must be analyzed or


explored is the environmental events (antecedents) that preceded a person's
aggressive behavior. We have already discussed the stimulus conditions that
can evoke aggressive tendencies. For example, "sensory overload" from the
environment can be aversive to a TBI survivor. The person may become
aggressive to provoke the staff to remove him or her to a less stimulating
environment. A review of the antecedent conditions could reveal this precipitating condition. If they can recognize the stimuli that lead to the aggressive

324

III. Rehabilitation Techniques

behavior and can then alter those stimuli, staff members may be able to reduce
the probability of the person's acting aggressively to provoke staff into removing him or her from the specific setting to a less stimulating environment.
3.2 Consequences

The other component in the person-environment interaction is the events that


immediately follow the aggressive behavior-the consequences. There are
two main classes of consequences: 1) reward (reinforcers), and 2) punishment (the withholding or withdrawal of reinforcers and the presentation of
an aversive event).
Reinforcement is a process in which a particular consequence that follows a
behavior increases the likelihood of that behavior's recurrence. For example, if
a TBI survivor receives social praise immediately after telling a therapist he was
feeling angry and frustrated at not being able to perform a task and if this is
shown to increase the frequency of his expressing negative feelings verbally-as
opposed to his previous habit of throwing task materials across the roomthen social praise is a positive reinforcer. In this example, positive reinforcement has resulted in an increase in the target behavior (verbal expression of
frustration). If this presentation of social praise does not eventually lead to an
increase in the frequency of the target behavior, then social praise is not a positive reinforcer for that person, even though such praise seems rewarding.
The frequency and duration of maladaptive behavior can also be affected by
withholding or withdrawing positive reinforcers immediately after the occurrence of such behavior. It can also be influenced by the staff's applying a consequence that the TBI survivor would rather not experience (i. e., it is aversive)
immediately following an aggressive behavior. Later in this chapter, we will
discuss the use of social learning principles and techniques to treat TBI survivors
who are behaving aggressively (see Sections 6 and 7).
3.3 Contingency Management

In a social learning approach, the overriding goal is to establish associations between the behavior exhibited and the consequences that follow this behavior.
It is the systematic and consistent occurrence of such associations that is often
labeled "contingency management." Learning occurs through this process. It
is very important to understand that reducing the frequency of aggressive behaviors can be accomplished using reinforcement and punishment techniques.
However, increasing the frequency of adaptive (situationally appropriate) behavior can only be achieved with reinforcement techniques. A management
strategy that fails to incorporate this reality is one that is likely to produce only
temporary effects. Training and reinforcement of pro social responses are essential if the effects of treatment are to generalize to the discharge environment.
3.4 Summary

When examining a TBI survivor's aggressive behavior, staff members should


think in terms of the behavioral event. A behavioral event consists of the ante-

13. Management of Aggressive Behavior

325

cedents, the behavior, and the consequences. These are the so-called "ABC's" of
behavior analysis. Behavior can be altered in a variety of ways. Staff members
can 1) alter the stimuli that evoke aggression, 2) alter the consequences that are
reinforcing the individual for behaving aggressively, and 3) reward nonaggressive behaviors that can achieve the same goal that was previously sought
through aggressive behavior.
4 ASSESSMENT

4.1 Social-Behavioral Diagnosis

Before they can initiate a treatment program based on social learning principles
and techniques, staff must conduct a diagnostic evaluation of the TBI survivor's
behavior. The end product of the social-behavioral assessment should be 1) an
operational definition of the aggressive behaviors, 2) data regarding the frequency and severity of these aggressive behaviors, 3) hypotheses regarding the
stimulus situations (antecedents) that tend to elicit these maladaptive responses,
and (4) environmental conditions (consequences) that appear to maintain them.
4.1.1 Target Behaviors

The common method of collecting behavioral assessment data is direct observation and measurement of the TBI survivor's behavior in the treatment
setting. The first step is to specify the behaviors that will be observed and
measured. These target behaviors must be described in clear, unambiguous
Table 13-1. Sample menu of aggressive behaviors
Code number

3
4
5
6
7

8
9
10
11
12
13

14
15
16
17
18
19

20

Behavior
Verbally abuses others.
Verbally threatens to harm property.
Verbally threatens to harm people.
Physically threatens to harm property.
Physically threatens to harm people.
Violates others' personal life space when agitated but does not specifically
threaten, harm, or make contact with others.
Physically strikes, hits, kicks, or bites others.
Destroyslrenders harm to another's property.
Destroyslrenders harm to his/her own property.
Takes another's personal property without authorization.
Makes suicidal/self-destructive gestures.
Attempts suicide.
Engages in hyperagitated behavior, including rapid pacing and excessive
movement or other high levels of psychomotor activity.
Engages in explosive verbal or physical outbursts.
Engages in argumentative/oppositional behavior when asked to perform a
behavior.
Engages in screaming, shouting behaviors unrelated to explosive outbursts
(#14) or oppositional behavior (#15).
Makes sexually offensive/vulgar remarks.
Makes sexually offensive/vulgar gestures.
Touches others in sexually offensive/aggressive ways.
Engages in other forms of disruptive, attention-seeking behaviors.

326

III. Rehabilitation Techniques

terms so that there is agreement among the raters regarding the behaviors of
interest. Bornstein and his colleagues [10] used a 12-item list of target aggressive behaviors. We have found it useful to create a master list (menu) of maladaptive behaviors that, if expressed by the TBI survivor, would typically be a
barrier to successful community reintegration. A subset of these maladaptive
behaviors is presented in Table 13-1. For these data to be useful for planning
an appropriate intervention strategy, the observation and recording of these
target behaviors must include all three elements of the behavioral analysisthe maladaptive behavior, the antecedents, and the consequences.
4.1.2 Baseline Measurement

Baseline refers to the frequency of target behaviors prior to the initiation of


treatment. Recording of baseline data is most useful for planning interventions
when staff members document information pertaining to all three clements of
the behavioral analysis (the maladaptive behavior, the antecedents, and the
consequences). An example of a form used to record such information is
displayed in Table 13-2.
A major issue that must be addressed in this process is observer error and
bias [16]. A principal source of error is inadequate training of the people who
are collecting baseline data. It is important to ensure that the observers understand the target behaviors. The list of target behaviors should be designed so
that the operational definitions are mutually exclusive. In addition, we have
found it useful for observers to practice categorizing behavioral responses
during their orientation to the baseline recording process. Finally, we have
employed senior clinicians to serve as role models by demonstrating how to
complete the rating forms. This process of on-line orientation and continued
training via supervision helps to reduce errors resulting from raters' poor
understanding of the target behaviors or the recording requirements.
Table 13-2. Data Collection Form: Aggressive Behaviors
Patient name: _ _ _ _ _ _ _ _ __
Behavior
code
number

Severity
rating

Date/
Time of
occurrence

Antecedents

Consequences

Comments

Staff
initials

13. Management of Aggressive Behavior

327

Kent and Foster [17] noted that "consensual observer drift" can be a major
source of error. This occurs when some observers alter the criteria used to
define the target behavior without informing other observers. In short, they
drift from predetermined operational definitions of the target behavior. This
can lead to some observers recording particular responses as target behaviors,
even though they do not belong in that category. In other cases, some observers may arbitrarily decide not to include certain actions of the TBI survivor in
their recording, when in fact they should have documented these acts. The
net effect is that there is either an inflation or underestimation of the actual
frequency of the target behaviors.
A related problem is one of "expectation bias." When conducting a baseline
measurement, an observer may expect that a maladaptive behavior will have a
high frequency. Similarly, in the intervention stage, an observer may expect
that the maladaptive behavior will occur less frequently than in the baseline
period.
The problem arises when the particular behavior emitted by the TBI survivor is not clear-cut. For example, verbal abusiveness may be the target
behavior. The operational definition includes the notion that the abusive
language has to be directed at a person. During the baseline period, the TBI
survivor who begins cursing indiscriminately, after a staff member reminds
him or her of a required task that he or she has not yet performed, may be
rated as being verbally abusive to that staff person. In the treatment phase,
however, the same behavior may be interpreted as a relatively harmless verbal
expression of frustration.
Another variation of expectation bias occurs when a behavior (e.g., indiscriminate cursing in response to frustration) is rated as something else (e. g.,
verbal abusiveness) due to the TBI survivor's history of being verbally abusive.
Verbal abusiveness is expected, and so the behavior is rated as such. As with
consensual observer drift, the net effect is an overestimation or an underestimation of the actual frequency of the maladaptive behavior. Both of these
sources of error can be reduced through training and ongoing supervision of
multiple observers. Quality control mechanisms for data collection in both the
assessment and treatment phases are another means of reducing errors.
There are excellent texts that address the issues involved in the mechanics of
baseline measurement and ongoing data collection in behavioral programming
[18, 19]. The clinical management team needs to decide the length of the baseline measurement period, the times during which a patient's behavior will be
observed and measured, the measurement parameters, the methods that will
be employed to ensure reliability, the methods of collapsing the raw data into
baseline profiles, and the methods used to analyze and interpret the baseline
data. Although a review of these issues is beyond the scope of this chapter, we
should mention that these decisions will be influenced by 1) the frequency and
intensity of the target behavior; 2) the type of setting in which these observations are made; 3) the number of baseline studies being conducted concurrently;

328

III. Rehabilitation Techniques

4) the prior experience, trammg, and nature of professional and paraprofessional observers; 5) the relative significance of the target behavior for achieving
clinical outcomes; and 6) the time frames for producing an initial report to
third-party payers.
We have tended to record every instance of the aggressive behaviors listed in
Table 13-1 during the baseline measurement period. The behavioral technicians describe the behavior (by code number) and assign it a severity rating.
One example of a severity rating code is presented in Table 13-3. The raters
also document the circumstances that preceded the. maladaptive behavior
(antecedents) and the events that followed (consequences). We have left space
in the data collection form to allow the rater to add brief comments that he or
she believes are pertinent to the situation.
The baseline data recorded on these forms provide the basis for establishing
precisely the nature and severity of the individual's problems. Herbert and her
colleagues [20] demonstrated that parents' perceptions relative to the frequency, intensity, and duration of behavior are not very valid when compared
to the objective data. Staff members' perceptions are also susceptible to errors
of judgment when compared to objective measurements of behavior [21].
A second function of baseline data is to provide a standard for determining
the relative efficacy of treatment interventions. This can best be accomplished
by employing single-subject paradigms. As contrasted to group designs in
which comparisons are made between a treatment group and a control group,
single-subject paradigms focus on change within a single individual as a function of treatment. The frequency, intensity, or duration of a target behavior
during the baseline period is compared with the levels observed when the
intervention is in place. In this way, the TBI survivor serves as his or her own
Table 13-3. Problem Behavior Rating Scale
A behavior is a social problem when it is socially inappropriate or disruptive (i. e., its frequency,
intensity, or duration interferes with ongoing environmental routines or activities).
(1) Mild
Behavior is inappropriate or somewhat disruptive, but it does not interfere
with activities or routines in the environment. (If the person makes
physical contact with others in any way that is situationally inappropriate,
rate as 2,3, or 4 as indicated.)
(2) Moderate

Behavior interferes with or interrupts normal ongoing environmental


routines or activities. The person must be atteuded to, but resumption of
normal routine/activities can be accomplished without requiring the
person to leave the setting. If physical contact that is situation ally
inappropriate occurs, no risk of physical harm or danger is involved. (If
risk of harm or danger is involved, rate 3 or 4.)

(3) Severe

Behavior is so disruptive that restoration of normal activities in the setting is


not possible until the patient leaves the setting, either voluntarily or with
assistance. The behavior endangers the patient or other individuals, but no
one is actually harmed or injured.

(4) Extreme

The behavior results in at least one person being harmed during the episode.

13. Management of Aggressive Behavior

329

control. Most single-subject designs (e. g., multiple baseline, simultaneous


treatment, or changing criterion) can be used to evaluate the effectiveness of
treatments for aggressive behavior. However, given the nature of aggression
and the potential risk of harm to anyone in the environment, it would be
neither ethical nor clinically advisable to utilize a reversal design. For further
information about the single-subject designs, the reader should consult Barlow
and Hersen [22], Johnston and Pennypacker [23], or Kazdin [24].
4.1.3 Behavioral Interview

Direct observational data can be supplemented by diagnostic data collected in a


behavioral interview. Even though some TBI survivors cannot participate in .
such an interview, many of them provide valuable information about their
understanding of their aggressive behavior. Despite how divergent their
viewpoint might be from the viewpoint of more objective observers, the TBI
survivors are reporting reality as they see it.
The interviewer should focus on the environmental events that trigger
aggression in a particular patient. Attention should be paid to the person's
thoughts, feelings, and interpretations of circumstances that were associated
with the occurrence of aggression. The specific types of aggressive behaviors
that have occurred should be explored, and the consequences-as well as the
individual's reaction to those consequences-should be discussed. The TBI
survivor should be asked about what alternative responses he or she could have
employed and the reasons why those were not used. Any predictions that the
patient has about what will likely occur in the future, given similar circumstances, should be elicited. Behavioral interviews can also include a review
of the results of self-report inventories such as the Novaco Anger Inventory
[25] and of self-monitoring checklists completed by the TBI survivor. Data
regarding these same factors can also be collected from family members.
4.2 Neurophysiological Diagnosis

It is impossible to consider biologically predisposed aggressive tendencies


(whose expression is influenced by stimulus conditions and environmental
consequences) without addressing the central role of the brain. Stimuli, whether
from within the individual or from the outside, are interpretable because of a
complex chain of neurochemical and neurophysiological events. Learning and
memory, which govern how the person will respond to these stimuli, are likewise dependent on complex neurochemical and neurophysiological events.
In utero and postbirth anomalies in neurochemical and neurophysiological
functions can influence the individual's proclivity to, and engagement in,
aggressive behavior. The disruptions in neurochemical and neurophysiological function following a TBI have enormous potential for influencing
the manifestation and modifiability of aggressive behavior [26].

330

III. Rehabilitation Techniques

4.2.1 Preinjury Neurobiological Factors

One aspect of diagnosis that appears underemphasized in rehabilitation programs is the analysis of organic constitutional factors that may be contributing
to the TBI survivor's aggression. There is a lack of empirical data about the
incidence of preexisting conditions such as attention deficit disorder, organic
personality syndrome, antisocial personality disorder [27], episodic dyscontrol
syndrome [28-30], or other disorders that may involve impairments of brain
function. It is important to screen patients for these disorders and syndromes.
Such screening should include a specification of the types of maladaptive behaviors that the individual demonstrated and the situations in which these behaviors typically occurred. The types of intervention strategies employed to
treat these disorders-especially pharmacological interventions-and the relative efficacy of these interventions should also be explored. This type of data
may prove valuable in identifying the determinants of aggressive behavior and
in selecting the best intervention strategies.
4.2.2 Postirijury Neurophysiological Factor: Seizures

Jennett [31] reported that 5% of all hospitalized TEl survivors develop late epilepsy. The risk of a late seizure's occurring during the first four years after injury increases to 35% for those who had an intracerebral hematoma removed
within two weeks after injury. For individuals with a depressed skull fracture,
the risk ranges from 3% to 60%, depending on a variety of early clinical signs
associated with the head injury. For those TBI survivors who develop late
seizures, Jennett indicated that approximately one half of these seizures were of
the grand mal type, whereas approximately one fifth were restricted to temporallobe attacks ("psychomotor fits"). Bond [32] noted:
Unusual and unrecognized mental symptoms and behavioural patterns ... may go
unnoticed by the inexperienced clinician ... often with an associated period of great
difficulty for the family and inappropriate management for the patient. [po 149}

Annegers and his colleagues [33] reported lower risk rates in a populationbased study that addressed some of the methodological shortcomings of
Jennett's [31] approach (i.e., selection bias of neurosurgical admissions and
referrals and selective follow-up). These authors reported the incidence oflate
seizures at one year and at five years for children (less than 15 years old) and
adults with severe, moderate, and mild head trauma. For children, the rates at
five years were 7.4%, 1.6%, and 0.2%, respectively, whereas for adults the
rates were 13.3%, 1.6%, and 0.8%, respectively.
The relationship between posttraumatic epilepsy (PTE) and aggression is
exceedingly complex [26]. Pond [34] noted that children with petit mal seizures who were referred to the Maudsley Hospital demonstrated neurotic rather
than conduct disorders, whereas young brain-injured epileptics who had grand
mal seizures and focal epilepsy showed unpredictable aggressive and explosive
behaviors. Patients with non traumatic temporal lobe epilepsy (TLE) in this

13. Management of Aggressive Behavior

331

group had the most severe behavioral disturbances. Few studies have examined
the possible relationship between PTE and aggression in TBI survivors. Data
regarding aggression in patients with TLE and in those manifesting the "episodic dyscontrol" syndrome [28-30] may provide valuable insights into this
vexing question.
In TLE, unprovoked, sustained aggression is rare [35]. For example, an
international panel of 18 experts in epilepsy reviewed closed-circuit television
tapes of 33 epileptic attacks in 19 patients who were selected from a group of
approximately 5,400 epileptics [36]. Only 7 of the 19 patients demonstrated
aggression toward inanimate objects or another person. Two of these patients
were head-trauma survivors. These aggressive acts were of sudden onset,
apparently unplanned, and of short duration (the average mean score was 29
seconds). "Aggressive acts were simple, unsustained, and never supported by
a consecutive series of purposeful movements" (p. 715) [36]. The behavioral
responses during the seizure (ictally) were consistent with an angry state in five
of the seven patients, fear in another, and confusion in the remaining case. All
seven reported having no memory of the aggressive act. Lishman's [26] review
of published data led him to conclude:
In so far as there is any increased risk of violent or antisocial conduct among epileptics,
it is unlikely to arise from the attacks themselves but rather from the psychiatric complications of the epilepsy. (p. 346)

There are a number of published findings on interictal aggression in TLE

[37-39]. The aggression seems to be a response designed to retaliate for perceived slights. These epileptic patients can recall the antecedent event that
was associated with the aggressive response, and they exhibit remorse for their
aggression. By contrast, people with character disorders typically claim no
recall for the event and deny any feelings of guilt or remorse.
Eames [40] noted that "episodic dyscontrol" may be a better model than
TLE for understanding severe conduct disorders in some TBI survivors.
Maletzky [30] described 22 patients who manifested severe violent behavior,
often with only minimal provocation. None of these patients had TLE, but
14 had abnormal electroencephalograms (EEGs)-six with temporal-lobe
spiking and eight with nonspecific temporal-lobe EEG anomalies. Their attacks were similar to epileptic attacks. Moreover, there was a high incidence of
family history of epilepsy. Maletzky noted that the potential for violent behavior was increased by alcohol intake and occasionally by drug treatments
with chlordiazepoxide. He also reported a positive treatment effect in an uncontrolled trial of phenytoin that resulted in approximately a 75% reduction in
the frequency and severity of violent attacks in 19 of the 22 patients.
Maletzky hypothesized that limbic-system abnormalities were probably
associated with these patients' exceedingly low threshold for uncontrollable
anger. Eames [40] observed that the more likely underlying pathology for the
severe conduct disorders of some TBI survivors may be temporolimbic epi-

332 In. Rehabilitation Techniques

leptic synd~omes rather than temporal lobe epileptic syndromes. Lishman


[26], in his discussion of explosive behavior, observed that pathophysiological
changes in the periamygdaloid region of the temporal lobes may be at the root
of such violent behavior.
Our main point is that certain forms of aggressive behavior in TBI survivors
may be a function of pathophysiological changes related to the brain trauma.
Evaluation of the nature of the aggressive episodes and of the circumstances
surrounding these episodes is a critical component of selecting appropriate
pharmacological and behavioral interventions.
4.2.3 Postinjury NeurophysWJ<lgica/ Factor: Effects of Medications

Anticonvulsant drug interventions-whether to manage seizures or to prevent


their occurrence in TBI survivors-have come under scrutiny with respect to
their efficacy and their potential negative impact on patients' cognitive functioning [42, 43]. The assessment of a TBI survivor should include an examination
of these issues. Thompson and Trimble [44] reported that the reduction in use
of multiple anticonvulsants (polypharmacy) and the use of carbamazepine
resulted in improvements in cognitive functioning with no increase in seizure
frequency. In one subset of epileptic patients, the shift from polypharmacy to
single drug interventions resulted in improved seizure control as well as enhanced cognitive functioning. These researchers also reported positive effects
(increased cognitive functioning; no increased frequency of seizure activity)
with low serum-level concentrations of anticonvulsants as compared with
high serum concentrations (in most cases, below toxic levels) [45]. Sodium
valproate and carbamazepine showed fewer adverse effects on cognitive functioning than did phenytoin, when doses were in therapeutic ranges. It should
be noted that in both these studies, the subjects had chronic epilepsy with causes
other than TBl.
The issues of efficacy and of adverse effects on cognitive and motor recovery
from TBI are also pertinent with respect to the pharmacological management
of the TBI survivor's aggression. The use and relative benefit of antipsychotic
agents such as haloperidol [46], tricyclic antidepressants such as amitriptyline
[47], and benzodiazepine derivatives such as chlordiazepoxide [48], as well as
drugs with both anticonvulsant and antipsychotic properties such as carbamazepine [49] and lithium [50-52], are reviewed in other chapters in this
book. The emphasis here, however, is on the importance of a thorough evaluation of the effects of seizure activity, as well as pharmacological interventions related to seizure control and control of aggression, on the TBI survivor's
neurophysiological system.
4.3 Neuropsychiatric Diagnosis
4.3.1 Preinjury Factors

A comprehensive review of the individual's neuropsychiatric status previous


to the head injury can reveal any functional or constitutional factors that are con-

13. Management of Aggressive Behavior

333

tributing to postinjury aggression. For example, Mary's (Section 1.4) postinjury


aggression was related to her preinjury psychiatric condition. During the clinical neuropsychologist's assessment, it was discovered that prior to her injury
she had been hospitalized for the treatment of depression. She had a history of
using aggression as a means of protecting herself from those social interactions
that reinforced her chronic low self-esteem. Subsequent to her injury, her
parents took her to their family physician because of her aggressive outbursts.
Mary withheld information about her psychiatric condition from this physician. He recommended a psychiatric consultation, but he also prescribed
haloperidol as a temporary measure. Neither Mary nor her parents followed
through with the psychiatric consultation for the next three weeks. When
Mary was admitted to the residential community reentry program, she was
still taking haloperidol.
With Mary's approval, we contacted her former psychiatrist. Her current
situation was reviewed with the psychiatrist, who agreed to reevaluate her.
Her medication was altered to lithium, which had previously been effective in
managing her aggression and depression. She responded well to this change
during her three-month stay, and her aggressive episodes decreased. She was
discharged to her home, and she continued receiving pharmacological and
supportive psychiatric treatment.
4.3.2 Postinjury Factors

Neuropsychiatric and clinical psychological assessment should include an analysis of the personality structure of the TBI survivor and of his or her family
members. Catastrophic injury such as TBI assaults a person's sense of wellbeing. Muir and Haffey [53] employed the concepts of "partial death" and
"mobile mourning" to describe the psychological aspect of coping with the
TBI event and its consequences-for the TBI survivor and for the members of
his or her "personal system" as well. Each person develops and relies upon a
set of individual strategies to cope with psychological loss. Assessment of each
person's preinjury behavioral responses when confronted with loss ("partial
deaths") can provide valuable information for predicting probable behavioral
responses to the TBI and its consequences.
Dealing with such a traumatic disruption of one's life is often an aversive
event. As a result, the probability of aggressive tendencies being evoked at
various times in the recovery process is high. Exploration of specific aspects of
this process of dealing with the consequences of the TBI can help identify the
stimuli that each person considers aversive. Rank-ordering these aversives and
assigning relative weights to the perceived stress value of each stimulus can
help pinpoint "red-flag" situations.
For example, analysis of the red-flag situation of AI's wife's talking on the
telephone (Section 1.3) and a review of AI's premorbid responses to stimulus
situations that elicited aggressive behavior led to the identification of those
therapeutic interventions that would probably be most effective for managing
his aggression. Training Al and his wife to employ these strategies took place

334

III. Rehabilitation Techniques

in the context of conjoint counseling sessions. Their use of these strategies in


their home environment was followed by a decrease in the frequency and
intensity of AI's overt aggression toward his wife.
We have already mentioned that aggressive behavior can be altered by providing or withholding or withdrawing reinforcers. The neuropsychiatric or
clinical psychological assessment can include a careful review of the person's
preinjury likes and dislikes, activity patterns, or leisure pursuits. Such information can help guide the selection of potential reinforcers. It can also reveal
potential sources of aggression when it becomes clear that access to a potent
reinforcer is restricted or precluded, given the nature of posttraumatic deficits.
For example, driving a car may be very reinforcing to a 19-year-old resident of
a community reentry program, yet providing access to driving may be unwise
or logistically unfeasible. The assessment data can alert the treatment team to
this likely source of frustration, which could lead to aggressive behavior.
In examining an individual's posttraumatic behavior, the neuropsychiatrist
or clinical psychologist may discover that the person has been described as uncooperative, noncompliant, nonresponsive to a behavior modification program, actively resistant to staff interventions, or unmotivated. There are many
reasons why a TBI survivor may have been so described. The neuropsychiatric
assessment might reveal either of two disorders described by Wood and Eames
[54] and Wood [41, 55] that are barriers to the successful application of social
learning principles and techniques.
The first condition is an organic one. It involves impairment of those brain
systems associated with reward and pleasure. The behavioral outcome of such
impairment is that the TEl survivor lacks a hedonic response (i.e., he or she
makes little or no effort to earn rewards or to avoid pain or other aversive
events). Thus, the modification of such a TBI survivor's behavior through the
usc of reinforcers or aversive events is virtually impossible. If a TBI survivor's
postinjury behavior is predominantly one in which little or no effort to earn
rewards or avoid aversive events (e.g., painful stimuli) has been noted, then
the assessment should include a review of the potential damage to deep diencephalic brain structures. If the findings are that these structures or their
connections with frontal brain structures were severely damaged, then the
neuropsychiatric assessment data will have particular significance for the selection of initial treatment goals and intervention procedures. The initial treatment goals may be restricted to eliciting effort from the TBI survivor to earn a
reward-or at least to avoid an aversive event. The method might be a classic
conditioning paradigm (see Section 4.5). In the event that such a goal cannot
be achieved, it is likely that no other rehabilitative goal could be accomplished,
and plans would have to be made for transferring such a person to a long-term
care facility where he or she could be protected from harming himself or
herself or others.
The second condition that can be a barrier to effective rehabilitation is one
described by Lishman as the "posttraumatic hysterical syndrome" [26]. Wood

13. Management of Aggressive Behavior

335

[41] reported an incidence of 20% (10 patients) among those TBI survivors
who had very severe behavior disorders and who were admitted to the Kemsley
Unit during its first six years of operation. During a six-year period, one of the
authors of this chapter (W.J. H.) can recall only one such case, but his sample of
TBI survivors had less severe behavioral disorders. These individuals manifest
conversion symptoms (i.e., physical symptoms such as gait disturbances or
contractures) that cannot be explained neurologically. In other words, despite
a range of impairments and disabilities that were understandable, given their
brain injuries, these TBI survivors exhibited behaviors whose presentation
could only be accounted for by assuming that the abnormality was of psychological, not neurological, origin. The two main diagnostic signs were that
1) the abnormality was inconsistent with any known neurological impairment
or disease; and 2) there was evidence that the function that was no longer able
to be performed (e. g., movement of a limb; vision) given the dissociative state,
could be performed voluntarily, usually when the person was apparently unobserved. Wood [41] gave an example of a woman who resisted staff efforts to
range a severely contracted left arm but who used the limb to reach out to grab
a sandwich to which she was not entitled. After putting the sandwich in her
mouth, she returned the arm to its contracted position. In such cases of dissociative behavior, the person exerts effort in ways that are directly contrary to
achieving target behaviors.
Wood [41] noted that only one of the ten patients whose rehabilitation was
blocked by such dissociative behavior had behaviors in adolescence that could
have been seen as precursors of such bizarre behavior. Thus, the main focus of
the neuropsychiatric assessment would be on postinjury behavior and would
be directed at ruling out all other potential explanations for the observed
abnormality.
4.4 Neuropsychological Diagnosis
Most rehabilitative specialists have begun to accept the use of neuropsychological assessment in the formulation of treatment goals and the specification
of the stimulus-response conditions that might optimize rehabilitative learning
[56]. TBI survivors, like other disabled people, receive feedback on a daily
basis that they are not able to perform tasks as effectively or efficiently as they
did before the onset of their brain injury. The situations in which they experience failure subjectively are aversive, and thus are potential stimuli for aggression. A person's frustration can be reduced by defining, on a task-by-task basis,
his or her learning potential and the conditions under which learning will be
maximized. Information about cognitive functioning that is employed as a
result of those definitions can assist in decreasing the aversive character of
rehabilitative training, thus reducing the potential for aggression.
Posttraumatic cognitive deficits are often a major determinant of the TBI
survivor's aggression [57]. These brain-based cognitive disorders can interfere
with the following domains of cognitive behavior: 1) the registration and

336

III. Rehabilitation Techniques

interpretation of environmental events; 2) the storage and retrieval of information; 3) the programming and execution of goal-directed behavior, including the prediction of probable outcomes; and 4) the evaluation of whether
the performance resulted in the intended outcome.
Social-behavioral competence depends upon each of these four domains of
cognitive behavior. Deficits in the aforementioned domains of cognitive behavior can be the principal determinants of social-behavioral incompetence.
Whichever neuropsychological method is employed to gather diagnostic data,
the evaluation must identify a person's performance strengths and liabilities in
each of these four areas.
These data need to be integrated with the behavioral assessment data to help
identify any potential contributions that these cognitive deficits make to the
person's aggressive behavior.
Of equal importance is the formulation of predictions of the person's learning ability and of the conditions that will enhance (or interfere with) such learning. This is particularly critical if the data suggest that the person's cognitive
deficits may interfere with establishing the associations between his or her
behavior and the contingent consequences. If the cognitive assessment data
reveal substantial attentional or learning deficits, then a classic conditioning
paradigm should be tried initially, rather than an operant one. An example of a
classic conditioning paradigm would be the repeated pairing of 1) the verbal
command "look at me," and 2) moving the person's head so that eye contact
was made with the therapist. Over time, if the conditioning were successful,
the verbal command would be sufficient to lead to the desired behavior (i. e.,
eye contact). In operant conditioning, behavior is governed by the consequences that immediately follow the behavior, as contrasted to those events
that precede it. For example, if the conditioning were successful, praising a
person immediately after each instance of eye contact with the therapist would
eventually lead to an increase in the frequency or duration of eye contact.
Wood [55] has noted that a massed practice procedure is often useful to determine to what extent associative learning can be demonstrated. In such a procedure, the person would be required to produce a single target behavior
repeatedly throughout the entire therapy session. The contingent reinforcer
would be delivered immediately following each instance of target behavior.
The degree to which the conditioned response is manifested empirically establishes the person's learning ability.
4.5 Integration of the Diagnostic Findings

We have emphasized the importance of a comprehensive assessment of the


possible determinants of the TBI survivor's aggression. The final step in the
assessment process is the integration of the findings from each diagnostic
source.
The best way to achieve this integration is by establishing a team of professionals with expertise in each of the domains of assessment that we have ad-

13. Management of Aggressive Behavior

337

dressed in this section. Upon the completion of their respective assessments,


members of such a team would. convene to discuss the hypotheses that have
emerged from these assessments. The final diagnostic report would represent
a synthesis of the findings.
The establishment of the intervention plan follows logically from such a
diagnostic synthesis. It requires a coordinated decision-making process regarding the selection of which treatment methods will be instituted, the timing of
their initiation (as well as the subsequent review of their efficacy), the plan for
additional intervention strategies in crisis situations, and the provisional plan
for subsequent interventions in the event that the initial interventions fail to
produce the desired results.
This coordinated approach is contrasted to one in which each professional
determines interventions that follow from his or her assessment findings and
communicates this intention to the other professionals in the context of a team
conference. The latter process can result in complementary or contradictory
treatment interventions.
5 MANAGEMENT OF AGGRESSIVE BEHAVIOR
DURING THE ASSESSMENT PHASE

We have emphasized the importance of conducting comprehensive assessments

prior to implementing treatment interventions. Aggression must be managed


effectively during this period to protect the patients' and staff's safety and
welfare [58]. A number of techniques can meet immediate management needs
without compromising the baseline measurement or without beginning actual
treatment interventions.
Management of aggression can begin upon a person's admission to the treatment setting. Orientation of the TBI survivor to the daily workings of the
setting can include a review of what is expected. This can be done repeatedly in
the initial few days in simple terms.
For example, a staff member might say: "If you feel angry about something
around here, tell a staff member that you are feeling angry. The staff person
will help you work out a solution. If the staff person is too busy, he or she will
find another staff person for you to talk to. Just remember, feeling angry is
OK. Acting in a way that could hurt you or someone else is not OK." Another
example might be: "No one can strike another person. If you strike another
person, you will be told to stop. If you do not stop immediately, we will physically stop you. You will be taken to our 'time-out' room. If necessary, we
will place you in restraints." Clear messages about what is and is not permitted
is an orientation technique that may reduce the need for crisis intervention.
If the person becomes combative or engages in potentially harmful behavior,
the staff must intervene immediately. A staff that is properly trained in crisis
management is one of the most effective means of managing aggression. The
existence of a standard protocol for how to respond to a variety of crisis situations in ways that protect everyone's safety and welfare-in ways that do not

338

III. Rehabilitation Techniques

increase the probability of maladaptive behavior's recurring-is central to


effective management of the TBI survivor's aggression. The staff must be
sufficiently trained in following these prescribed strategies so that they can
respond immediately (almost automatically) and in a coordinated fashion.
In a situation involving assault, staff members in the vicinity must immediately decide if they can handle the situation according to the prescribed procedure. If not, the first step is to get back-up help. However, if staff members
believe that they can manage the situation without help from others, they
should quickly determine the appropriate course of action. One technique may
be to physically restrain the assailant. Each staff member is assigned a part of
the patient's body to restrain. After removing any items that they are wearing
or carrying that could cause injury, staff members move in unison-at a predetermined signal-to immobilize the assailant. This is done with "minimal
human interaction" between the staff and the patient; by this we mean that eye
contact and verbalizations should be avoided. The physical force exerted by
the staff should be restricted to the amount required to protect all parties from
danger. The entire process is designed to be implemented rapidly and efficiently. Although this may seem dehumanizing or overly mechanical, it is effective
in controlling the dangerous situation in ways that reduce the probability of
anyone's being injured. It is also designed to reduce the probability ofinadvertently reinforcing the TBI survivor's aggression. If social attention is reinforcing to the individual, intervening in a manner that provides such attention
could increase the chances of a recurrence of this behavior. The more sterile the
exchange, the better.
Training in such physical restraint procedures is essential for staff members
who deal with TBI survivors who demonstrate severe aggressive behaviors.
Such techniques can be learned from printed materials and videotapes [59-61]
or in training workshops [62].
During the time-limited assessment process, as we have said, staff must
intervene to stop an aggressive act in order to protect everyone's safety and
welfare. It is also important to restate that the baseline measures need to be
conducted before implementing treatment protocols. Because of the anxiety
that aggression evokes in staff members and other residents, there is a strong
temptation to begin employing pharmacological or punishment techniques to
reduce the frequency ofaggressive episodes before identifying the determinants
of the TEl survivor's aggression. The liability of such a premature institution
of treatment protocols is that it may mask the true nature of the behavioral
disorder. The result might be that inaccurate judgments about the nature of the
disorder may be made that subsequently may lead to a misdirected treatment
regimen. Another liability is that the assessment phase is often discontinued
once such interventions are instituted, with the result that potentially critical
information is never collected.
The principal means of resisting the temptation to implement treatment
interventions prematurely is a well-trained staff that functions effectively in

13. Management of Aggressive Behavior 339

cnsIs situations. If staff members believe that they can handle any crisis in
ways that prevent injury to themselves and others, they have greater tolerance
for the time-limited period of observation and measurement prior to initiating
treatment. This is especially true when they are integrally involved in recording behavioral observations that provide the data for designing treatment
strategies.
We in no way underestimate the difficulty involved in creating and maintaining such a therapeutic environment. Nonetheless, for centers that treat a
high volume of TBI survivors with severe behavioral disorders, we believe
that the time and energy invested in such a process is one of the primary means
of achieving efficacious results with this very challenging clinical population.
6 DECREASING AGGRESSIVE BEHAVIOR,
USING SOCIAL LEARNING PRINCIPLES AND TECHNIQUES

The community resettlement of a TBI survivor who is acting aggressively


depends on reducing the frequency or intensity of maladaptive behaviors and
on promoting reliance on prosocial behaviors to meet his or her perceived
needs. The remainder of this chapter will focus on the use of social learning
principles and techniques to achieve such outcomes in cases in which assessment findings indicate that such interventions are warranted.
6.1 Reinforcement

Philosophically and clinically, we are committed to Bellack and Hersen's [63]


dictum:
Punishment should not be used when positive procedures can be as effective. If there is
any doubt, positive procedures should be employed first and punishment applied only
if positive procedures have been unsatisfactory. (p. 210)

We cannot overemphasize the therapeutic potential of positive reinforcement


for reducing aggressive behavior as well as for eliciting, strengthening, or maintaining prosocial behavior. Unfortunately, positive reinforcement paradigms
seem to be underutilized for reducing aggressive responses. We believe that
this is because people realize that positive reinforcement increases the frequency
of the behavior that occurs immediately before the delivery of the reinforcing
consequence. Because the goal is to decrease aggression, it appears contradictory
to employ positive reinforcement to accomplish this goal.
The key to using positive reinforcement to decrease aggression is the production of an increase in the frequency of nonaggressive target behaviors in those
situations in which TBI survivors rely on aggression to meet their perceived
needs. Differential reinforcement is a technique by which positive reinforcement is employed to decrease aggressive behavior [64, 65]. This technique involves delivering positive reinforcers following nonaggressive responses in
general (this is called "differential reinforcement of other behaviors" or DRO)

340

III. Rehabilitation Techniques

or following a specific prosocial alternate behavior (known as "differential


reinforcement of alternative behaviors" or DRA). A less frequently used variation is "differential reinforcement of low rates" (DRL), in which reinforcement is delivered for a reduction in the overall frequency of aggression within
a specified time frame or an increase in the time interval that elapses between
aggressive episodes.
The DRO technique can be easily employed in a therapeutic milieu in which
staff actively focus on rewarding nonaggressive behavior. This seems obvious.
However, examination of staff-patient interactions in many units often reveals
that disruptive patients receive more attention, more corrective feedback, and
more staff time and energy than patients who present relatively few management problems. Therefore, the first step toward decreasing aggressive behavior
is a focused effort to observe and reward nonaggressive behavior.
The potency of the DRO technique can be enhanced if staff members verbally "label" (by telling patients) why reinforcers are being presented. This is
because there are so many possible nonaggressive responses that could be followed by positive reinforcement. Consequently, the association between these
many possible responses and the delivery of a positive reinforcer is relatively
weak. Verbal labeling helps identify the contingency. For example, if a therapist said "That was very good" after a TBI survivor expressed her frustration
verbally rather than by throwing task materials across the room, it is possible
that the survivor would not make the connection between her verbal statement
and the social praise that followed her comment. However, if the therapist
labeled the contingency by saying, "Right after our session, I'll take you for a
short walk around the grounds because you said you were feeling frustrated instead of throwing things," then the chances are increased that the association
between the target behavior and the reinforcer will be made.
At the beginning of an intervention plan designed to decrease frequent
aggression, DRO is often used in combination with punishment techniques
such as time-out, overcorrection, and positive practice (see Section 6.3, Punishment) to bring severe aggression under control [65]. Unfortunately, too often
the tendency is to abandon the DRO technique in favor of punishment interventions. We strongly agree with Kazdin's [66] statement, "The best use
of punishment in applied settings is as an ancillary technique to accompany
positive reinforcement" (p. 193).
The DRA technique involves rewarding the TBI survivors for employing a
more acceptable alternative response to deal with the situation in which they
are currently relying on aggressive responses to meet perceived needs. This
technique is often employed in conjunction with specific skill training in areas
such as assertiveness, conflict resolution, problem solving, and release of
frustration.
For example, because of his attention aI, perceptual, and motor deficits, Jack
(Section 1. 2) experienced high levels of frustration when he attempted to prepare meals. When he encountered problems in performing cooking tasks, he

13. Management of Aggressive Behavior

341

would hurl cooking utensils, food-and whatever else was readily availableacross the kitchen. Initially, we trained him to use alternative responses that
were already in his behavioral repertoire (e.g., cursing, or slamming his hand
on the countertop) to express his frustration. Reinforcement (verbal praise,
"points" that could later be exchanged for prearranged privileges) was delivered when he used alternative responses instead of more aggressive and potentially dangerous responses. Eventually, the reinforcement was delivered
only after he expressed his frustration verbally. This DRA technique was accompanied by the delivery of positive reinforcement for each component step
that he performed during the cooking task (i. e., successive approximation).
Eventually, he was rewarded only upon completion of the "terminal
behavior"-in this case, cooking the meal.
The DRL technique involves the delivery of positive reinforcement following a specified time period during which the overall frequency of an aggressive
behavior decreases to a specified level or the time interval between aggressive
acts decreases. We have never employed this technique for reducing combative
or other potentially dangerous behavior. However, we have used it for reducing nondangerous aggressive behavior such as verbal abusiveness, hyperagitation, and so forth. In most cases, we have used this approach to expand
time intervals between the occurrences of aggressive behaviors such as verbal
abusiveness. In some cases, we have employed it following an overall decrease
in aggressive behaviors (e.g., verbal threats to harm a person or some property) over periods of time such as a month. We typically did this in conjunction with a specification of performance goals for discharge (which we
labeled "graduation criteria").
The "differential reinforcement of incompatible responding" (DRI) technique provides positive reinforcement following a behavior that is topographically incompatible with aggression. For example, sitting quietly in a chair is
topographically incompatible with hitting someone. However, because aggression has so many potential expressions, this differential reinforcement technique cannot always practically be employed to reduce aggressive behavior. For
instance, a person who is reinforced for keeping his hands in his pockets (topographically incompatible with punching someone) could still be aggressive by
kicking someone.
We have been advocating the use of positive reinforcement to achieve a reduction in the frequency and severity of the TBI survivor's aggression. We
recognize that the potency of these positive reinforcers is at times inadequate to
overcome the reinforcement value of behaving aggressively. This requires us
to employ extinction and punishment paradigms in our quest to reduce the
TBI survivor's aggression.
6.2 Extinction

Extinction is a technique that is designed to achieve a reduction in the frequency


of target behavior [18] by removing the positive reinforcement that maintains

342

III. Rehabilitation Techniqnes

the behavior. To use an extinction paradigm, staff must identify and control
the maintaining consequence. This technique is often difficult to implement,
especially when the maintaining reinforcers occur intermittently. The decreased
frequency will typically occur gradually and will often be preceded by a temporary increase in the frequency of the maladaptive behavior. This increase,
which is called an "extinction burst," results from the person's attempt to
restore the maintaining consequence.
Extinction is of limited value for reducing most aggressive behaviors.
Aggressive behavior can have many maintaining consequences, and identification and removal of all of these reinforcers is often impossible. In cases in
which the aggression is already dangerous or an increase in the severity of the
disordered behavior would result in a dangerous situation, it would be clinically inappropriate to employ an extinction paradigm. Extinction procedures
tend to elicit rage reactions or frustration-related aggression. This is unacceptable when the level of aggression already poses a risk to people's safety and
well-being.
When aggression is reinforced by social attention, an extinction procedure
can be very useful. As long as the aggressive behavior does not have a high
potential for harm (e.g., verbal abuse, screaming), consciously ignoring the
behavior can be effective. The removal of the maintaining consequence (social
attention) can gradually lead to the reduction in this type of aggression. The
problem that some staff members encounter is that they believe that ignoring
the behavior is tantamount to approving it. Rather than seeing it as a treatment
intervention, they perceive it as irresponsible permissiveness. In their effort to
discipline the aggressive person, those staff members provide the reinforcing
attention. This undermines the efforts of other staff members and results in the
technique's being ineffective for controlling these behaviors. This is but one
example of the need for consistency in the staff's interactions with the TBI
surVIVOr.
Ongoing supervision and training are the primary means of ensuring that
each staff member is contributing to the learning process rather than inhibiting
it. It is also important to emphasize to such staff members how the appropriate
use of social learning techniques is a more effective teaching method than their
personal efforts at verbally correcting or disciplining the TBI survivors. Kazdin
[66] points out that "for most behaviors brought to treatment, the weak or
inconsistent effects of verbal reprimands are not sufficient to achieve therapeutic change" (p. 165). Alternative punishment techniques such as time-out,
response-cost, overcorrection, and positive practice-when used in concert
with positive reinforcement techniques-are more effective teaching strategies.
(These techniques will be described in greater detail in Section 6.3.)
6.3 Punishment

Sometimes TBI survivors' aggressive behaviors persist, even after staff members have employed differential reinforcement and extinction paradigms. In
such cases, it is often necessary to use punishment paradigms to control po-

13. Management of Aggressive Behavior

343

tentially dangerous situations or to get the aggression under sufficient control


to enable staff to employ other treatment techniques. Punishment techniques
include 1) the removal of the person from situations in which positive reinforcement is available (time-out procedures); 2) the contingent application of
predetermined fines or penalties (response-cost procedures); 3) the requirement that the environmental consequences of an aggressive act be corrected
(restitution) and that prosocial alternatives be repeatedly rehearsed (positive
practice); and 4) the contingent application of an aversive stimulus (positive
punishment procedures) [is].
6.3.1 Time-Out Procedures

The purpose of time-out procedures is to remove sources of positive reinforcement immediately following the aggressive behavior. This removal can occur
within the setting in which the aggression occurred (time-out-on-the-spot,
situational time-out) or it can involve separating the TBI survivor from the
setting and directing him or her to an environment that is sterile with respect
to availability of positive reinforcers (time-out room). Such a procedure is
aversive because it restricts access to positive reinforcers. Wood and Eames
[54] and Wood [41,55] have reported that use of time-out is effective in reducing the frequency or intensity of maladaptive behaviors with TBI survivors
who have severe conduct disorders.
Staff must address a number of issues when they are employing time-out
procedures. Time-out is more accurately labeled "time-out from positive reinforcement" [64]. Consequently, these procedures are effective only when the
TBI survivor has access to positive reinforcers in the treatment setting. If the
environments in which the TBI survivor spends his or her time (e.g.,
treatment unit, therapist's office) are not associated with the delivery of
positive reinforcers, a time-out procedure is rendered ineffective. If the procedure includes removal to a time-out room, such a process is nothing more
than social isolation [67].
Assuming that positive reinforcers are available, the time-out procedure
must result in removal of all opportunities that are positively reinforcing [64].
Therefore, sending a person from the treatment setting to his or her roomwhere television, radios, tape players, or other potentially reinforcing items
are available-defeats the purpose. Moreover, this situation could also reinforce the behavior that preceded the time-out procedure. "Time-out rooms"
should be relatively barren to ensure that positive reinforcement is unavailable.
There is another way in which removal to a sterile environment can produce
an undesired effect. If the TBI survivor is behaving aggressively to escape
from or avoid an aversive situation, then removing the person from the demand situation-even to an environment that is devoid of positive reinforcerswill increase the possibility of a recurrence of the aggressive behavior.
Time-out procedures are designed to effect learning. This is central to the
issue of criteria for release from time-out. Some staff, family members, patient
advocates, administrators, and regulatory agents argue for preestablished,

344

III. Rehabilitation Techniques

fixed-time intervals for time-out. Release occurs at the conclusion of this time
period, irrespective of the TBI survivor's behavior while he or she is in timeout. There are a number of motives for adopting this position, including a
fundamental aversion to punishment paradigms and a genuine concern for
ensuring patients' rights. However, the problem with such an approach is that
it does not foster the intended learning. To accomplish this objective, release
from time-out should be response-contingent. That is to say, the TBI survivor
must remain in time-out for a predetermined time interval and release from
time-out should occur only after the person has been behaving nonaggressively
for a set period of time [68]. For example, the time-out protocol could include
a minimum five-minute duration, with release only after one continuous
minute of nonaggression.
Matson and DiLorenzo [68] provide data supporting the effectiveness of a
response-contingent release approach. White, Nielson, and Johnson [69] specifically address the issue of time-out duration. They found that short durations
may be as effective as longer durations for suppressing maladaptive behavior
in some individuals, whereas longer durations were required to produce the
desired effect in others. The implication is that staff must empirically determine the optimal time parameters for using time-out with each TBI survivor.
Although time-out procedures can be effective, removal of the TBI survivor
from normal treatment environments does temporarily eliminate the possibility
of positively reinforcing adaptive behaviors. Therefore, as with all other effective punishment procedures, the outcome is suppression of maladaptive
behavior rather than the learning of adaptive behavior. However, such suppression is often essential in order to make learning under other conditions
feasible.
Our final observation about the use of time-out procedures that involve the
removal of the TBI survivor from treatment environments deals with the staff's
attitude. Because the TBI survivor's aggression is typically an aversive stimulus
to staff, the removal of that aversive stimulus from the treatment environment
can negatively reinforce the staff's use of time-out procedures [64]. If staff members are not vigilant, the use of a time-out procedure to reduce problem
behaviors can be deemed effective not because it results in a decrease in the disordered behavior (the objective of the procedure) but rather because it enables
staff to return to their clinical duties in an environment that is unencumbered
by the TBI survivor who is behaving maladaptively.
6.3.2 Response-Cost Procedures

These procedures involve the contingent loss of a positive reinforcer or the


contingent imposition of a penalty [66]. This penalty may take the form of a
fine or may involve requiring the person to perform an act that he or she
would not otherwise be required to perform. The theoretical basis of such a
procedure is that the contingently applied fine or penalty is aversive to the TBI
survivor. Because the person wants to avoid the aversive event, there is a de-

13. Management of Aggressive Behavior

345

creased probability that the maladaptive behavior that immediately preceded


the imposition of the response-cost procedure will recur.
The first implication of this theoretical construct is that the behaviors that
are followed by a response-cost procedure must be clearly specified, so that
staff and TBI survivors alike know when such procedures will be invoked.
Predetermining a finite set of severe maladaptive behaviors that will always be
followed by a response-cost intervention should minimize staff confusion and
inconsistency in applying the procedure.
Without clear specification of target behaviors and the corresponding response-cost procedure, individual staff members may fail to invoke the predetermined procedure following the TBI survivor's emitting a target behavior.
In other instances, a staff member might decide to apply a response-cost procedure following a maladaptive behavior that has not been included on the
predetermined protocol. This is inappropriate because the TBI survivor cannot be expected to avoid a penalty that he or she was unaware was forthcoming.
An additional advantage of such a pre~etermined list is that the list can be posted
for all to see. A TBI survivor who has memory problems can be given this list
in conjunction with whatever external memory aids are being used.
There are two additional issues that must be addressed when response-cost
procedures are being used. First, even if these procedures are not being used in
a token economy system [70], there is still an economic base to a procedure
that imposes fines. A person cannot pay a fine ifhe or she has no assets-unless
deficit spending is permitted. The TBI survivor could be emitting maladaptive
behaviors at such a high frequency that the total amount of fines outstrips his
or her potential earnings. Such a situation often precipitates frustration and
aggression or a refusal to participate further in such a program. One solution
to this problem is to dispense points or tokens initially for behaviors that only
approximate the target adaptive behaviors. The TBI survivor must ultimately
perform the terminal behavior to receive these points or tokens. (This process
is called "shaping.") This serves a dual purpose: the person is earning rewards,
and the deficit-spending situation is minimized or avoided.
The second issue involves the inherent delay between the application of the
penalty and the actual experience of the penalty. The loss of a positive reinforcer often does not immediately follow the maladaptive behavior. The greater
the delay between the behavior and the punishing event, the less effective the
procedure for suppressing the behavior [71]. This problem can be minimized
by reducing the delay as much as possible. At the time when access to the
expected reinforcer is denied, staff can highlight the contingent relationship by
verbally labeling the reason for the loss of the reinforcer.
6.3.3 Overcorrection and Positive Practice

Overcorrection is a punishment technique that requires the person to rectify


the environmental effects of his or her aggression. This mandated restitution is
often accompanied by repeated practice of alternative prosocial responses that

346

III. Rehabilitation Techniques

could have been employed in the situation. This combination of mandated restitution and positive practice helps the aggressor, other TBI survivors, and staff
see that maladaptive behavior has an immediate, highly visible consequence.
This consequence is more readily apparent than the time-delayed consequences
involved in response-cost interventions or the less visible consequence involved
in time-out procedures.
The result is often twofold-enhanced learning by TBI survivors and increased morale among staff members who have difficulty understanding the
nature of consequences in social-learning interventions. Staff members can be
taught to focus their energy on supervising the TBI survivor in restitution and
positive practice activities, rather than verbally reprimanding him or her. Their
need for seeing tangible consequences following aggressive behavior can be
channeled into helping the aggressor repeatedly practice (e. g., for five minutes)
responses that he or she could have employed to get the perceived needs met
in a more socially acceptable manner. This often involves staff's modeling
prosocial behaviors that the rehabilitation team has determined are necessary
to support the TBI survivor's community resettlement. Such a channeling of
the focus of staff members reduces their frustration and its concomitant potential for inappropriate and abusive application of punishment techniques. It
also increases staff members' satisfaction, because they perceive themselves as
teachers of appropriate behavior.
If the TBI survivor was engaging in a behavior that was potentially dangerous or if the person's emotional control was inadequate to enable him or her to
participate in such restitution and positive practice activities without risk of
another behavioral outburst, then other punishment techniques (time-out or
response-cost) should be applied immediately, Once the person has regained
control, staff can institute the overcorrection and positive practice interventions.
One means of restitution is apologizing to the people who bore the brunt of
the TBI survivor's aggression. This can be accomplished either through verbal
or written apologies to each individual or in larger community meetings in
which the person has to apologize publicly.
Positive practice can be one part of a more comprehensive social skills training approach. Employing overcorrection and positive practice techniques requires more staff time, energy, and focus than is typically expended once the
aggression is brought under control. Nevertheless, such an approach can be
extremely important if staff members are to communicate effectively to the
TBI survivor that aggression is counterproductive and if they are really going
to teach the person how to develop alternative ways to meet his or her needsways that will support community resettlement.
6.3.4 Positive Punishment

When the TBI survivor persists in severe aggressive behavior despite all other
treatment interventions, staff must confront the issue of whether to use posi-

13. Management of Aggressive Behavior

347

tive punishment techniques. Positive punishment involves the contingent


application of an aversive event, such as placing aromatic ammonia under the
person's nose, squirting water in the person's face or lemon juice into his or
her mouth, using corporal punishment, or applying mild electric shock.
The general population and many health-care professionals have a negative
response to such controversial measures. Clinical liabilities are also associated
with such procedures. For example, the person might respond by developing
a phobic-like response to all treatment personnel who have been involved in
such punishment [66]. This may render the treatment environment ineffective
for assisting the TBI survivor to achieve community resettlement. However,
when all else fails, and in cases in which self-destructiveness is extreme, one
might argue that such procedures serve the person's ultimate well-beingwhen the alternative is to permit such extreme self-destructiveness to occur
[55]. Nevertheless, such methods seem unwarranted and unfeasible in treating
most TBI survivors' aggression. In addition, regulations, statutory provisions,
or judicial orders may restrict or preclude their use, even in these extreme cases
[72].
6.4 Summary

Aggressive behavior must be brought under control to reduce the danger to


everyone's safety and welfare and to provide an opportunity for the TBI survivor to learn prosocial alternatives to his or her current aggressive responses.
Prior to instituting punishment techniques, staff should employ positive reinforcement techniques in cOrUunction with efforts at controlling stimulus
situations that are related to the person's aggression. Matson and DiLorenzo
[68] provide three rules of thumb that staff should follow when the use of
punishment is contemplated:
First, it must be documented that a maladaptive response is occurring at a sufficient
rates to warrant intervention. Second, controlling variables (i. e., antecedent and consequent conditions) must be delineated. Third, a demonstration of the lack of effectiveness of less intrusive techniques must be observed before punishment procedures are
attempted. (p. 160)

When clinically indicated, punishment techniques have been demonstrated


to be effective in suppressing maladaptive behavior, and they provide an alternative to other interventions such as the use of neuroleptic medications, whose
clinical effectiveness is often offset by their accompanying undesirable side effects. Repp and Deitz [73] have argued that withholding clinically appropriate
and effective punishment interventions would be unethical. Punishment techniques-when successful-result in a suppression of maladaptive behavior
that allows staff members to focus on their primary mission, which is to teach
TBI survivors those prosocial behaviors that are necessary to support their
resettlement in the community.

348

III. Rehabilitation Techniques

7 PROMOTING ADAPTIVE BEHAVIORS

Reducing the frequency of the TBI survivor's aggression is no small accomplishment. Unfortunately, such an outcome is typically insufficient to support
successful community resettlement. The second necessary element is teaching
the TBI survivor to rely on more socially desirable ways of meeting his or her
perceived needs. Positive reinforcement techniques are the principal means of
accomplishing this. We will now focus on issues involved in employing positive reinforcement paradigms to promote the acquisition and continued use of
prosocial behaviors.
7.1 Selection of Target Behaviors

Identification of the target behaviors that will be followed by the provision of


positive reinforcers is one issue confronting the rehabilitation team. We recommend that the team attempt to determine the response demands that the
TBI survivor will confront in his or her discharge environment. In some ways,
this is relatively easy with respect to aggression. Combative behavior is typically universally unacceptable. On the other hand, there is often a range of behaviors that people in various discharge environments will tolerate with respect
to dealing with frustration. The criteria that distinguish acceptable from unacceptable behavior vary from setting to setting. Different standards will be
operative in the home, on the job, or in school. Within each type of setting,
specific individuals will use different criteria to judge the adequacy of the TBI
survivor's responses.
The responses is an urban setting are often different from those in a rural
environment, and there are often regional differences as well. Behavior that is
tolerated in rural Texas may be considered deviant in rural Massachusetts.
Ethnic and socioeconomic groups can have different parameters of acceptable
and unacceptable behavior relative to other groups.
There are several reasons why it is important to pay specific attention to
the criteria that will be employed in various discharge settings to judge the
adequacy of the TBI survivor's responses.
First, typically there is a finite number of adaptive behaviors that TBI survivors who have severe behavioral disorders will be able to incorporate during
the course of most treatment programs. Second, TBI survivors who have cognitive deficits will typically demonstrate problems with transferring learned
behaviors from one situation to another. Third, TBI survivors will often have
to rely on trained compensatory strategies to meet situational demands. These
strategies must be ecologically relevant [74]; that is, the strategy must be suited
for use in the discharge environments. A strategy that a TBI survivor believes
will highlight to others that he or she is deviant from the norm will typically
not be employed unless it renders him or her successful. However, the reinforcement value of this success must be more potent that the aversive effects
associated with using the strategy. Unless those who design such compensatory
strategies pay specific attention to this reality, the probability of generalization
is extremely low.

13. Management of Aggressive Behavior

349

Fourth, interventions may have to be directed at altering the criteria that


key individuals in the discharge environment have been applying (and will
be applying) to judge a TEl survivor's response adequacy to increase the
probability of a positive outcome.
An analysis of the response requirements that will be placed upon the TEl
survivor in the discharge environment is therefore a critical step in selecting
the target behaviors that will be elicited and maintained using reinforcement
paradigms.
7.2 Selection of Positive Reinforcers

Reinforcement paradigms, whether for the reduction of aggression or the promotion of prosocial behaviors, are dependent on the delivery of positive reinforcers. There are three major classes of reinforcers: social, material, and
activity reinforcers.
Social reinforcers such as attention, praise, physical contact, a smile, and so
forth are potentially excellent positive reinforcers. They are easily administered,
occur naturally in most environments, and do not generally interrupt on-line
performance. Social reinforcers can easily be paired with many reinforcing
events, such as earning points that can be used to gain access to certain privileges that are reinforcing, or success in performing a task. As a result, their
potency as a reinforcer is enhanced [66]. Material reinforcers such as food,
beverages, cigarettes, and tokens are often used as incentives in contingency
management programs. Activity reinforcers can include activities that a person
engages in, or desires to engage in, frequently. These may include recreational
activities, therapeutic tasks that the person finds personally satisfying, passes
for activities outside the treatment facility, and so forth.
The selection of potential reinforcers is a trial-and-error process. Potential
reinforcers must be identified and then tested to determine whether their application in fact results in an increase in the frequency of the target behavior that
they immediately followed. This is referred to as the empirical law of ~{fect [64].
Staff members might ask the TEl survivor or a family member about preinjury interests and stimulus events that the TEl survivor experienced as motivating. Observation of current postinjury activity patterns can also provide
data about potential incentives. Whitman, Scibak, and Reid [75] point out that
although Premack's principle [76] asserts that a high probability event will
serve as a reinforcer for a lower probability event, the response deprivation
hypothesis specifies that any response that occurs occasionally can be a reinforcer if the person's relative access to this activity is denied [77]. Structuring a
program so that access to some activities must be earned (by demonstrating
adaptive behavior or by refraining from maladaptive behavior) may increase
the number of available reinforcers. Stimulus events that have been associated
with primary reinforcers can be potential reinforcers. Wood and Eames [54]
employed such conditioned reinforcers in their brain-injury unit.
Ease of administration of a reinforcer is an important variable to consider
when selecting reinforcers. As we noted earlier, this is one of the potential

350

III. Rehabilitation Techniques

benefits of social reinforcers as compared to many activity reinforcers. A simply administered activity reinforcer (e.g., 10 minutes of unsupervised rest) will
more likely be used when the alternative is an activity reinforcer that is complicated and time-consuming to administer (e. g., a supervised trip to the local
bowling alley). Social reinforcers typically involve little cost (i.e., personnel
time), as contrasted to certain material reinforcers (e. g., special foods). If
reinforcers are difficult to administer, if they require significant staff time and
effort, or if they are costly, the likelihood is that they will be poor choices for
reinforcers.
7.3 Implementing Positive Reinforcement Paradigms

Rehabilitation professionals tend to incorporate social reinforcement into their


training interactions. We have two observations that have previously been
well-received by such personnel.
First, social praise, smiles, a warm physical touch, and encouraging words
are often delivered more randomly than one might expect. The power of these
reinforcers can be enhanced if they are delivered immediately after the
performance of the target behavior or a component of that terminal behavior
(successive approximation). This helps the TBI survivor know precisely what
behavior is desired, in contrast to a more general association between the
therapist and the delivery of a positive reinforcer.
Second, these social reinforcers are often delivered along with specific verbal
directives that are designed to cue the person about the proper technique that
he or she should be employing to perform the behavior. Given that many TEl
survivors have difficulty discriminating stimuli, it is potentially more effective
to separate the delivery of critical verbal directives during task performance
from the delivery of the social reinforcer, which should be administered only
upon the person's achievement of the target behavior (whether it is one
component of the desired response or the terminal behavior itself).
We strongly advocate embedding all rehabilitative skill training in a social
learning context [78]. For example, systematic delivery of positive reinforcers
following the use of trained compensatory strategies in the natural environment
of the treatment milieu-as contrasted to the specific training context (e.g.,
therapist's office)-is central to the overall success of the therapeutic intervention. All clinical and paraprofessional staff members need to be informed about
the specific compensatory strategies that therapists have determined are the
most probable means for achieving behavioral competence. These compensatory behaviors are the target behaviors. All personnel can be trained to deliver
specific positive reinforcers immediately after the TBI survivor performs such
target behaviors. This is especially critical in those stimulus situations in which
the TBI survivor was previously behaving aggressively.
One principal method of decreasing aggression is the replacement of aggressive responses with trained adaptive responses. This is why reinforcement
should be delivered immediately after every instance of this alternative response

13. Management of Aggressive Behavior

351

(continuous reinforcement). Once these responses have been incorporated into


the TBI survivor's behavioral repertoire, then the reinforcement schedule can
be altered to an intermittent one. Acquiring a behavior typically requires a
continuous reinforcement schedule; stabilizing it is often best served by an
intermittent schedule. This shift in schedules is the first step toward improving
the probability of generalization of the trained effect. Fading of external reinforcers in favor of self-reinforcement is the second step. This is difficult to
accomplish with the TBI survivor. Nonetheless, the effort must be made, even
when it is likely that those in the discharge environment will have to be trained
in the use of those external reinforcement techniques that have been effective.
Family training in the use of social learning principles and techniques should
be an integral part of such treatment programs. Analysis of various family
members' preadmission reinforcement and punishment patterns in response to
the TBI survivor's behavior is a critical component of such training. Identification of these patterns will reveal what specific educative and supportive
interventions will be required to alter the family members' responses. Such an
approach is often difficult, because of their prior experience with the TBI survivor's maladaptive behavior, but it is an important means of increasing the
probability that generalization will occur. There is ample evidence that families
of disabled persons can be trained to apply behavior management procedures
in the home environment [79-81]. It seems unreasonable to assume that such
training would be less effective with families of TBI survivors.
Another underutilized reinforcement technique involves making participation in an event that is presumed to be reinforcing (because of its high frequency or because of comments made by the TBI survivor) contingent upon
participation in an event that is presumably aversive (as evidenced by compliance problems). This is referred to as the "Premack principle" [76]. We have
found this to be particularly useful when compliance is a problem in one
therapy (e.g., cognitive retraining), and there is a therapy that the patient/
resident particularly enjoys (e. g., physical therapy). The therapy schedule
could be arranged so that the physical therapist would begin a session only
after receiving a report about the patient'slresident's acceptable participation in
the previous cognitive retraining schedule. It should be noted that this is a defensible procedure only when the target behaviors in the cognitive retraining
session are critical to achieving discharge outcomes and when other alternatives
to establishing and maintaining such target behaviors cannot be employed.
8 IMPEDIMENTS TO SUCCESSFUL APPLICATION
OF SOCIAL LEARNING PRINCIPLES AND TECHNIQUES

Despite theoretical soundness, conceptual clarity, and technical expertise, rehabilitation personnel may find their efforts thwarted by factors within the
TBI survivor and by structural and attitudinal factors in the treatment environment. This chapter concludes with some observations about these impediments
to achieving desired outcomes.

352

III. Rehabilitation Techniques

8.1 Impairment Issues

The learning model presented above requires the occurrence of repeated associations between a behavior and a consequence. Severe attentional disorders,
especially when they are characterized by a breakdown in a person's alertness
and selectivity [82], can present a formidable barrier to establishing such associations. This is most typically seen in the agitated, confusional state in the
acute recovery phase. Fortunately, this is generally a transitory state, during
which the principal interventions are reduction of environmental contributions
to this state and protection of the person and others in the environment from
harm [83]. However, we have seen a small subset ofTBI survivors in the latter
stages of recovery whose attentional deficits were insurmountable barriers to
establishing the necessary associations between target behaviors and contingent
consequences.
The more common concern is the negative impact that persistent cognitive
deficits have on associative learning. As we have previously noted, the most
direct way to establish learning potential is to initiate a classic conditioning
paradigm and a massed practice strategy. Such an approach would enable a
decision to be made with respect to the person's capacity to benefit from the
treatment program. This process is cost-effective because people whose learning potential is inadequate can be identified sooner than typically occurs in
many rehabilitation programs. If this is the outcome, less costly management
options can be explored. On the other hand, those people who have severe cognitive deficits and who do demonstrate conditioned responses can be continued
in a treatment program. If an operant conditioning paradigm had initially been
employed, such people might not have demonstrated adequate learning. In
such cases, this method can help reduce the probability of treatment's being
discontinued prematurely.
Classic and operant conditioning paradigms have been employed successfully
with clinical populations of people who have substantive cognitive impairments, including autism [84], schizophrenia [85], mental retardation [86], and
developmental handicaps [87, 88]. These paradigms have also been successfully
employed with TEl survivors [41, 54, 55, 89, 90]. We do not mean to suggest
that classic and operant conditioning paradigms can be effective with all levels
of cognitively impaired TBI survivors. We are simply noting that some people
with substantial cognitive constraints have demonstrated learning in these
paradigms.
Motivational problems can present a substantial barrier to successful implementation of a contingency management approach. Wood and Eames [54] and
Wood [41, 55] have noted that a subset of their TEl survivors had organic
impairments that resulted in a condition in which reward, or even the avoidance
of pain, was an ineffective consequence for altering maladaptive behavior. We
have experienced cases in which the potency of empirically validated social,
material, or activity reinforcers was too weak to overcome the reinforcing
effect of avoiding situations in which the TEl survivor had to expend sub-

13. Management of Aggressive Behavior

353

stantial effort to perform the target behavior. As a result, the TBI survivor
would forego access to the otherwise reinforcing event, rather than put forth
the effort required to perform the target behavior.
In those situations in which the performance of the target behavior was essential to achieving placement in a less restrictive environment or would result
in the need for support from others in the projected discharge setting, we typically determined that the most palatable alternative was to make access to reinforcers such as specific types of food, snacks, cigarettes, selected recreational
pursuits, social phone calls, and visits home contingent upon performance of
target behaviors. Although it is clinically indicated, there are those who maintain that such an approach is unethical because it restricts access to what they
consider inalienable rights or privileges.
There are no easy solutions to the dilemma posed by the employment of
material and activity reinforcers in the treatment of TBI survivors with behavioral disorders of sufficient severity to interfere with their resettlement in
the community. Of course, the dignity of the individual must be respected. In
addition, his or her civil and legal rights are protected by laws and regulations
that vary from jurisdiction to jurisdiction. It is imperative that program personnel be cognizant of, and be in full compliance with, these mandates. (See
Chapter 14.) But there is also the ethical responsibility to employ techniques
that have been proven effective for reducing those behavioral liabilities that
restrict that person's access to home, school, work, and social environments.
Individual freedoms may be temperarily restricted within the context of a
management approach whose ultimate goal is community resettlement. This
eventually enables the person to have relatively unrestricted access to life experiences and reinforcing events that they otherwise would be barred from due to
their persistent socially unacceptable behaviors.
The other major learning constraint involves generalization. Many TBI
survivors demonstrate problems with transferring learning from one situation
to another. This relates to transfer of learning within the treatment setting
(e. g., from the therapist's office to the naturally occurring events in the
facility) as well as from the treatment setting to the discharge setting. We have
already discussed ways to improve the probability of such generalization. Without such specific programming, the probability of generalization of the trained
effect is extremely low [91]. The magnitude of this problem of generalization
can only be determined by systematic, long-term programmatic followup
studies. Eames and Wood's [92] followup study of the first 24 admissions to
the Kemsley Unit is an example of such an effort.
8.2 Structural Issues

Training rehabilitative staff or families is central to the appropriate use of social


learning principles and techniques. Research has indicated that typical staff
training programs do not improve on-the-job performance [93-95]. Two
reasons have been hypothesized. First, training typically occurs in a different

354

III. Rehabilitation Techniques

environment (e. g., a conference room or classroom) than that in which the
skills need to be employed. Second, skills that are acquired are not maintained
because staff members are not reinforced for using them.
The implications of these hypotheses are that training needs to occur at least
partially in the treatment setting [96]. Senior clinical specialists who have demonstrated competency in using these principles and techniques to manage
aggressive patientslresidents can be potent role models and master teachers in
the very settings in which they and other staff members have to function.
Training needs to focus specifically on what each staff member actually needs
to do (or not do) in specific situations. The underlying rationale for each of
these actions should be explained within the context of how this specific action
will assist the TBI survivor to gain better control of his or her behavior.
Training should involve on-site review of each staff person's behavior when
handling an aggressive episode or when implementing an intervention strategy.
In the context of such periodic reviews, performance requirements can be
clarified and supportive corrective feedback that outlines areas for continued
skill development can be offered. Because staff members' behavior is subject to
the same conditions that affect patientlresident behavior, efforts to improve it
should focus on antecedent events and consequences. With reference to antecedents, staff members can be provided with specific instructions about what to
do, when to do it, where to do it, and with whom to do it. Sneed and Bible
[97] and Iwata and his colleagues [93] have reported success using this strategy.
Another antecedent procedure is modeling, in which staff members observe the
clinical specialist's demonstrations of how to behave in specific situations [98].
Consequences have included supervisor praise [94], rearrangement of work
schedules [93], publicly posted feedback [99], private written feedback [100],
and disciplinary action (e.g., loss of pay, termination) [101].
The personnel time involved in such a training program represents a substantial investment on the part of the organization. Administrative personnel
and senior management must be educated about the value of such efforts. The
probability of eliciting a supportive response is generally enhanced by emphasizing factors like improved staff performance, reduced staff turnover, expansion of types of patients who could be treated, better patient outcomes, and the
marketing potential of these last two factors. Without such administrative and
senior management support, an ongoing training and development program
will likely either fail or be abandoned after initial success.
In addition to the competence of the people employing behavioral technologies, the facility itself can be an important variable [102]. For example, in
many treatment settings, physical space is usually less than optimal in both size
and design. This can lead to an environment in which "sensory overload"
becomes a problem for some patientslresidents. We have already noted how
this can lead to aggression and unintentional reinforcement of this aggression.
For patients who are easily distracted or who are experiencing difficulties in
screening out competing stimuli, use of "quiet areas" can reduce the prob-

13. Management of Aggressive Behavior

355

ability of aggressive behavior. Designing therapy schedules that take into


account patient response capacity can be difficult due to problems with maximizing therapists' efficiency, but such a step can often reduce the probability of
escape-motivated aggression. In rehabilitation centers in which productivity
standards are linked to the number of "therapy units" delivered to patients,
this type of clinical flexibility can run counter to fiscal demands placed on the
individual therapists. Administrative support and interdisciplinary cooperation
are essential to make such a system work.
8.3 Attitudinal Issues

A major barrier to the successful implementation of social learning principles


and techniques is rooted in the general public's negative attitude toward such
programs. Films like "A Clockwork Orange" and newspaper accounts of
patient abuse in programs employing aversive techniques reinforce this negativity. In addition to orientation and training of staff and family members
in the appropriate use of social learning principles and techniques, specific
review of each protocol, with an explanation of the rationale and objectives of
the specific procedure, is one way to counteract these misconceptions.
TEl survivors, family members, and administrative, clinical, and paraprofessional staff members who have been so oriented and trained may also react
negatively because of their aversion to treating previously "normal" adults as
if they were disturbed children or were mentally retarded. One solution is to
acknowledge that such methods have been used with success with such populations. Nonetheless, it can be explained that the use of the technology in this
situation is due to its potential benefit-returning the TEl survivor to his or
her preinjury setting at the highest level of functioning that can be achieved.
This in no way implies that the TEl survivor is necessarily a disordered child
or is mentally retarded. This negative attitude is most effectively countered
when the overriding atmosphere in the therapeutic setting is one of respect for
the rights and dignity of each TEl survivor.
8.4 Staff Training

Establishing such a milieu is totally dependent upon the attitude of the staff.
Working with severely aggressive TEl survivors requires a special type of
person, just as working on a burn unit or an oncology ward requires a certain
temperament and attitude. The approach that we have advocated is one that
emphasizes the systematic delivery of positive reinforcement. It is emotionally
challenging to reach out to a person and communicate genuine caring for that
individual when he or she may have spent part of the last hour verbally abusing
you, may have hurled a tray of food at you, or may have struck you. This becomes more difficult as time progresses and your sincere efforts at communicating genuine regard for that person are met with behaviors that seem
to indicate that he or she has absolutely no regard for you or anyone else.

356

III. Rehabilitation Techniques

Moreover, if the time, energy, and effort that you expend as a change agent
is met by persistent aberrant behavior, it is difficult to sustain such effort.
The solution to this problem is to select staff who recognize the demands
placed on them by this population and who demonstrate personality traits that
are suited for such work. Second, the staff members must work to build a
genuine support system among ,themselves. Staff must be given permission
during any shift to request temporary relief from specific patient management
duties. A person making such a request must feel that he or she is behaving in
ways that are consistent with appropriate self-regulation, rather than feel a
subtle message that if he or she were a "better" rehabilitationist, such special
treatment would not be required.
Staff cohesiveness is critical. Frequent informal parties for celebrating birthdays, anniversaries, and so forth are useful in building such rapport. More
formal group support meetings and process-issues meetings also help break
down attitudinal barriers.
Aggression evokes powerful responses in all of us. When staff have had to
resort to physically restraining a TBI survivor whose aggression presents a
danger to self and others, many powerful emotions may be triggered in those
staff who have had to respond to the crisis. In addition to having confidence in
one's own and other staff members' ability to handle the crisis (due to the consistent training in crisis intervention and physical management), each staff
member also has to have an appropriate means to deal with such emotions. We
have found that, once the crisis has been managed, those staff members who
were involved should leave the unit for a debriefing both to review technical
performance and to release any emotional residues from the situation. When
this type of administrative support is available in the work environment, staff
have a greater probability of behaving appropriately, and their attitude will
serve, rather than impede, the accomplishment of therapeutic objectives.
Another valuable attitudinal intervention has been helping the staff view the
world from the vantage point of the TBI survivor and his or her family members. Efforts directed at training the staff to "walk in the shoes" of these individuals have resulted in the development of a greater level of empathy. This
empathy has enabled them to weather the emotional assaults that they experience in dealing with these individuals. Even more so, it has enabled them to
be more responsive to their needs.
9 SUMMARY

The aggressive behavior of TBI survivors is a major barrier to their resettlement in the community. We have examined factors that can give rise to and
maintain either maladaptive or adaptive behavioral response to situations that
TBI survivors confront in their attempts to resume preinjury roles, activities,
and statuses. We have emphasized the critical importance of selecting intervention strategies based upon the use of social learning principles and techniques
for managing TBI survivors' aggression and for establishing behavioral re-

13. Management of Aggressive Behavior

357

sponses that could support their efforts at community resettlement. We hope


that the ideas in this chapter will contribute to reducing the degree of human
tragedy associated with traumatic brain injury.
REFERENCES
1. Heilman, K., Watson, R.T. and Valenstein, E. (1985). Neglect and related disorders. In
Clillical Neuropsycholo:.;y, 2nd ed., Heilman K. and Valenstein, E., eds., Oxford University
Press, New York, pp. 243-294.
2. Bandura, A. (1969). Principles of Behavior Modijicatioll, Holt, Rinehart, & Winston, New
York.
3. Tedeschi, J. T., Smith, R. B., III, and Brown, R. c., Jr. (1974). A reinterpretation of research
on aggression. Psycho!. Bul!. 8, 540-562.

4. Bandura, A. (1973). Aggressiol1: A Social Leaming Analysis. Prentice-Hall, Englewood Cliffs,


NJ.
5. Zillman, D. (1979). Hostility alld A:.;gressioll, Erlbaum, Hillsdale, NJ.
6. Eichelman, B., Elliott, R., and Barchas, J.D. (1981). Biochemical, pharmacological, and
genetic aspects of aggression. In Biobchaviol'al Aspects of Aggressioll, Hamburg D.A. and
Trudeau M.B., cds., Liss, New York, pp. 52-84.
7. Berkowitz, L., Cochran, S. and Embree, M. (1981). Physical pain and the goal ofaversivcly
stimulated aggression. J. Pers. Soc. Psycho!. 40, 687-700.
8. Feshback, S. (1964). The function of aggression and the regulation of aggressive drive.
I'sycho!. Rev. 71, 257-272.
9. Rule, B.G. and Nesdale, A. (1974). Differing functions of aggression. J. Pers. 42, 467-481.
10. Bornstein, P., Weisser, C. and Balleweg, B. (1985). Anger and violent behavior. In Handbook of Clinical Behavioral Therapy with Adults, Hersen M. and Bellack, A. S. eds., Plenum
Press, New York, pp. 603-628.
11. Freud, S. (1922). Beyolld the Pleasure Prillciple. International Psychoanalytic Press, London.
12. Lorenz, K. (1971). 011 Aggression. Bantam Books, New York.
13. Berkowitz, L. (1984). Human aggression. In Personality a/ld the Behavioral Disorders, Endler,
N.S. and Hunt, J. MeV. cds., Wiley, New York, pp. 349-372.
14. Pavlov, I.P. (1928). Lectures 011 Conditioned Reflexes, Gant, W.H. trans, International Publishers, N ew York.
15. Skinner, B. F. (1953). Sciellcc mId HUlnall Behavior. Macmillan, New York.
16. Kazdin, A.E. (1977). Artifact, bias, and complexity of assessment: The ABCs of reliability.
J. App!. Behav. Ana!. 10, 141-150.
17. Kent, R. N. and Foster, S.L. (1977). Direct observational procedures: Methodological issues
in naturalistic settings. In Handbook of Behavioral Assessment, Ciminero, A.R., Calhoun K.S.
and Adams H.E., eds., Wiley, New York, pp. 279-328.
18. Sulzer-Azaroff, B. and Mayer, G.R. (1977). Applyillg Behavior-analysis Procedures with ChiidYCII
mId Youth. Holt, Rinehart, & Winston, New York.
19. Sackett, G.P., ed. (1978). Observil1:'; Behavior: Vol. II. Data Collection alld Analysis Methods.
University Park Press, Baltimore.
20. Herbert, E.W., Pinkston, E.N., Hayden, M.L., Sajwaj, T.E., Pinkston, S., Cordua, G. and
Jackson, C. (1973). Adverse effects of differential parental attention. J. App!. Behav. Ana!. 6,
15-30.
21. Loeber, R. (1971). Engineering the behavioral engineer. J. App!. Behav. Ana!. 4, 321-326.
22. Barlow, D.H. and Hersen, M. (1984). Sil1:.;le Case ExperilnClltal DesigllS: Strate:.;iesfor Studyillg
Behavioral Chal1,!;e, 2nd ed. Pergamon, New York.
23. Johnston, J.M. and Pennypacker, H.S. (1980). Strategies mId Tactics of Humml Behavioral
Research. Lawrence Erlbaum Associates, Hillsdale, NJ.
24. Kazdin, A.E. (1982). Sillgle Case Research Desi:.;ns: Methods for Clinical and Applied Settings.
Oxford University Press, Oxford, UK.
25. Novaco, R.W. (1975). Anger COlltrol: The Deve/opmellt and Evaluation of an Experimelltal
TreatmCllt. Lexington Books, Lexington, MA.
26. Lishman, W.A. (1978) Organic Psychiatry. Blackwell, Oxford, UK.
27. American Psychiatric Association. (1980). Diagnostic and Statistical Mallual of MCIltal Disorders,
3rd ed. APA, Washington, DC.

358

III. Rehabilitation Techniques

28. Mark, V.H. and Ervin, F.R. (1970). ViolfllCC and the Brain. Harper and Row, New York.
29. Bach-y-Rita, G., Lion, J.R., Climent, C.E. and Ervin, F.R 1971. Episodic dyscontroi:
A study of 130 violent patients. Am. J. Psychiatr. 127, 1473-1478.
30. Maletzky, B.M. (1973). The episodic dyscontrol syndrome. Dis. Nerv. Sys. 34, 178-185.
31. Jennett, B. (1983). Post-traumatic epilepsy. In Rehabilitation of the Head Illjured Adult,
Rosenthal, M., Griffith, E.R., Bond M.R. and Miller,J.D., eds., F.A. Davis, Philadelphia,
pp. 119-124.
32. Bond, M. (1984). Psychiatry of closed head injury. In Closed Head Injury, Brooks, N. ed.,
Oxford University Press, Oxford, UK, pp. 148-178.
33. Annegers,J.F., Grabow,J.D., Groover, R.V., Laws, E.R.,Jr., Elveback, L.R and Kurland,
L.T. (1980). Seizures after head trauma: A population study. Neurology 300,683-689.
34. Pond, D.A. (1952). Psychiatric aspects of epilepsy in children. J. Ment. Sci. 98, 404-410.
35. Leventhal, B.L. and Brodie, H.K. (1981). The pharmacology of violence. In Biobehavioral
Aspects of Aggression, Hamburg, D.A. and Trudeau, M.B. eds., Liss, New York., pp. 85-106.
36. Delgado, A.V., Mattson, RH., King, L., Godensohn, E.S., Spiegel, H., Madsen, J.,
Crandall, P., Dreifuss, F. and Potter, RJ. (1981). The nature of aggression during epileptic
seizures. N. Engl. J. Med. 305, 711-716.
37. Waxman, S.G. and Geschwind, N. (1975). The interictal behavior syndrome of temporal
lobe epilepsy. Arch. Gen. Psychol. 32, 1580-1586.
38. Bear, D. and Fedio, P. (1977). Quantitative analysis of interictal behavior in temporal lobe
epilepsy. Arch. Neurol. 34, 454-467.
39. Devinsky, O. and Bear, D. (1984). Varieties of aggressive behavior in temporal epilepsy.
Am. J. Psychiatr. 141, 651-656.
40. Eames, P. In press. Posttraumatic epilepsy. In Closed Head Injury: Medical Mallageltlent,
Berrol S., ed., Oxford University Press, Oxford, UK.
41. Wood, R. (1987). Braill bljury Rehabilitation: A Neurobehavioral Approach. Aspen, Rockville,
MD.
42. Trimble, M.R. (1985). Psychopharmacology of Epilepsy. Wiley, New York.
43. Cope, N. In press, Psychopharmacologic management of traumatic head injury. In Closed
Head Injury: Medical ManageltlC11t, Berrol S., ed., Oxford University Press, Oxford, UK.
44. Thompson, P.J. and Trimble, M.R. (1982). Anticonvulsant drugs and cognitive function.
Epilepsia 25, 531- 544.
45. Thompson, P.J. and Trimble, M.R. (1983). Anticonvulsant serum levels: Relationship to
cognitive functioning. J. Neurol. Neurosurg. Psychiatr. 46, 227-233.
46. Rao, N., Jellinek, H.M. and Woolston, D.C. (1985). Agitation in closed head injury: Haloperidol effects on rehabilitation outcome. Arch. Phys. Med. Rehabil. 66, 30-34.
47. Jackson, R.D., Corrigan, J.D. and Arnett, J.A. (1985). Amitriptyline for agitation in head
injury. Arch. Phys. Med. Rehabil. 66, 180-181.
48. Monroe, RR (1975). Anticonvulsants in the treatment of aggression. J. Nerv. Ment. Dis.
160,119-126.
49. Tunks, E.R. and Dermer, S. W. (1977). Carbamazepine in the dyscontrol syndrome associated with limbic system dysfunction. J. Nerv. Ment. Dis. 164, 56-63.
50. Sheard, M.H. (1975). Lithium in the treatment of aggression. J. Nerv. Ment. Dis. 160,
108-118.
51. Sheard, M.H., Marini, J.L., Bridges, c.L. and Wagner, E. (1976). The effect of lithium on
impulsive aggressive behavior in man. Am. J. Psychiatr. 133, 1409-1413.
52. Sheard, M.H. and Marini, J.L. (1978). Treatment of human aggressive behavior: 4 case
studies of the effect of lithium. Compo Psychiatry 19, 37-45.
53. Muir, C.A. and Haffey, W.j. (1983). Treatment of behavioral deficits. In Rehabilitation of the
Head Injured Adult, Rosenthal, M., Griffith, E.R, Bond, M.R. and Miller, J.D. eds., F.A.
Davis, Philadelphia; pp. 381-393.
54. Wood, R and Eames, P. (1981). Application of behaviour modification in the treatment of
traumatically brain-injured adults. In ApplicatiOl1S of Conditioning Theory, Davey, G. ed.,
Methuen, New York, pp. 81-101.
55. Wood, R. (1984). Behaviour disorders following severe brain injury: their presentation and
psychological management. In Closed Head Injury, Brooks N., ed., Oxford University Press,
Oxford, UK, pp. 195-219.

13. Management of Aggressive Behavior

359

56. Haffey, W.]. (1984). Cognitive assessment and rehabilitative treatment planning. Paper presented at the American Congress of Rehabilitation Medicine, Boston, MA.
57. Malec,]. (1977). Training the brain-injured client in behavioral self-management skills. In
Behavioral Assessment and Rehabilitatiotl of the Traumatically Brain-DamaRed, Edelstein, B. and
Couture G., eds., Plenum, New York, pp. 121-150.
58. Thackrey, M. (1986). Therapeutics for ARRressiol/: PsycholoRicaliPhysical Crisis Intervention.
Human Sciences Press, New York.
59. New York State Office of Mental Retardation and Developmental Disabilities. (1980).
Physical Intervel/tiotl Tech'liques (videotape recording and text). Available from Bureau of
Staff Development and Training, New York State Office of Mental Retardation and
Developmental Disabilities, Albany, NY 12229.
60. Branson, R., Peterson, G., Infantino,]. and Grow, G. (1983). Aggression Control Techniques:
Instructor and Student Manual. Florida State University, Center for Educational Technology, Tallahassee.
61. Belanger, N. and Mullen, ].K. (1984). Safe Physical MmwRemellt (videotape and text). Available from Safe Physical Management Associates, New Bloomfield, PA 32301.
62. National Crisis Prevention Institute. (1986). Non-Violent Crisis Intervention Training
Workshop, 4011 W. Capital Drive. Milwaukee, WI 53216; (800) 558-8976.
63. Bellack, A.S. and Hersen, M. (1977). Operant approaches with children. In Behavior Modification: An Imroductory Textbook, Hersen, M. and Bellack, A.S. eds., Oxford, University
Press, New York, pp. 172-214.
64. Sulzer-Azaroff, B. and Mayer, G. (1986). Achieving EducatiOllal Excellence Using Behavioral
Strategies. Holt, Rinehart, and Winston, New York.
65. Repp, A. and Brulle, A. (1981). Reducing aggression of mentally retarded persons. In Handbook of Behavior Modification with the Mentally Retarded, Matson,].L. and McCartney,]. eds.,
Plenum, New York, pp. 177-210.
66. Kazdin, A.E. (1980). Behavior Modification ill Applied Settings. Dorsey, Homewood, IL.
67. Drabman, R. and Spitalnik, R. (1973). Social isolation as a punishment procedure: A controlled study. ]. Exp. Child Psycho!. 16, 236-249.
68. Matson, J.L. and DiLorenzo, T.M. (1984). Punishment and Its Alternatives. Springer, New
York.
69. White, G.D., Nielson, G. and Johnson, S.M. (1972). Timeout duration and the suppression
of deviant behavior in children. ]. App!. Behav. Anal. 5, 111-120.
70. Ayllon, T. and Azrin, N.H. (1968). The Token Economy. Appleton, New York.
71. Azrin, N.H. and Holz, W.e. (1966). Punishment. In Operant Behavior: Areas of Research and
Application, Honig W.R., ed., Appleton, New York, pp. 380-447.
72. Fisher, K. (1987). Agreement reached on aversi,,:es. APA Monitor 18, 24.
73. Repp, A.e. and Deitz, D.E. (1978). On the selective use of punishment-suggested guidelines for administrators. Ment. Retard. 16, 250-254.
74. Gillis, ].S. (1971). Ecological relevance ancl the study of cognitive disorders. In Cognitive
Studies. Vol. 2: Deficits in Cognition, Hellmuth]. ed., Brunner/Mazel, New York, pp. 1-15.
75. Whitman, T.L., Scibak, ].W. and Reid, D.H. (1983). Behavior Modification with the Severely
and Profoundly Retarded. Academic Press, New York.
76. Premack, D. (1959). Toward empirical behavior laws: I. Positive reinforcement. Psycho!.
Rev. 66, 219-233.
77. Konarski, E.R., Jr., Johnson, M.R., Crowell, e.R. and Whitman, T.L. (1980). Response
deprivation and reinforcement in applied settings: A preliminary analysis.]. App!. Behav.
Ana!. 13, 595-609.
78. Haffey, W.]. (1984). Social learning principles and techniques in the rehabilitation of brain
injury survivors. Paper presented at the Annual Meeting of the American Congress of Rehabilitation Medicine, Boston, MA.
79. Becker, W. (1970). Parents Are Teachers: A Child Mmwgement Program. Research Press,
Champaign, IL.
80. Patterson, G. (1971). Families: Applicatiol/S of Social Leaming to Family Life. Research Press,
Champaign, IL.
81. Fallon, I., Boyd, J., McGill, e., Stang,]. and Moss, H. (1981). Family management training
in the community care of schizophrenia. In New Directions for Mental Health Services: New

360

82.
83.
84.
85.
86.
87.
88.
89.
90.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.

III. Rehabilitation Techniques

Developmolts in Interventions with Families of Schizophrenics, Goldstein, M. ed., Jossey-Bass,


San Francisco, pp. 61-77.
van Zomeren, A.H., Brouwer, W.H. and Deelman, B.G. (1984). Attentional deficits: The
riddles of selectivity, speed, and alertness. In Closed Head Injury: Psychological, Social and
Family Consequences, Brooks, N. ed., Oxford University Press, Oxford, U.K., pp. 74-107.
Eames, P., Haffey, W. and Cope, N. In press. Treatment of behavioral deficits. In Rehahilitation of the Adult and Child with Traumatic Brain Injury, Griffith, E.R., Rosenthal, M., Bond,
M. and Miller, J.D., cds., F.A. Davis, Philadelphia.
Schriebman, L., Koegel, R.M., Clladop, M.H. and Egel, A.L. (1982). Autism, In international Handbook of Behavior Modijication and Therapy, Bellack, A.S., Hersen, M. and
Kazdin, A.E., eds., Plenum, New York, pp. 891-916.
Curran, J.P., Monti, P.M. and Corriveau, D.P. (1982). Treatment of schizophrenia. In International Handbook of Behavior Modijicati'>n and Therapy, Bellack, A. S., Hersen, M. and Kazdin,
A.E., eds., Plenum, New York, pp. 433-466.
N eisworth, J. T. and Madle, R. A. (1982). Retardation. In Intemational Handbook of Behavior
ModijicatiOll and Therapy, Bellack, A.S., Hersen, M. and Kazdin, A.E. eds., Plenum, New
York, pp. 853-890.
Luiselli, J.K. and Slocumb, P.R. (1983). Management of multiple aggressive behaviors by
differential reinforcement. J. Behav. Ther. Exp. Psychiatry 14, 343-347.
Altman, K. and Krupsaw, R. (1983). Suppressing aggressive-destructive behavior by delayed overcorrection. J. Behav. Ther. Exp. Psychiatry 14, 359-362.
Hollon, T.H. (1973). Behavior modification in a community hospital rehabilitation unit.
Arch. Phys. Med. Rehabil. 54, 65-68.
Horton, A.M. and Howe, N.R. (1981). Behavioral treatment of the traumatically braininjured: A case study. Percept. Mot. Skills 53, 349-350.
Stokes, T.F. and Baer, D.M. (1977). An implicit technology of generalization. J. Appl.
Behav. Anal. 10, 349-367.
Eames, P. and Wood, R. (1985). Rehabilitation after severe brain injury: A follow-up study
of a behavior modification approach. J. Neurol. Neurosurg. Psychiatry 48, 613-619.
Iwata, B.A., Bailey, J.S., Brown, K.M., Foshee, T.J. and Alpern, M.A. (1976). A performance-based lottery to improve residential care and training by institutional staff. J. Appl.
Behav. Anal. 9, 417-431.
Montegar, C.A., Reid, D.H., Madsen,C.H. and Ewell, M.D. (1977). Increasing institutional staff-to-resident interactions through in-service training and supervisor approval.
Behav. Ther. 8, 533-540.
Greene, B.F., Willis, B.S., Levy, R. and Bailey, J.S. (1978). Measuring client gains from
staff implemented programs. J. Appl. Behav. Anal. 11, 395-412.
Page, T.J., Iwata, B.A. and Reid, D.H. (1982). Pyramidal training: A large-scale application
with institutional staff. J. Appl. Behav. Anal. 15, 335-351.
Sneed, T.J. and Bible, G. H. (1979). An administrative procedure for improving staff performance in an institutional setting for retarded persons. Ment. Retard. 2, 92-94.
Gladstone, B.W. and Spencer, c.J. (1977). The effect of modeling on the contingent praise
of mental retardation counselors. J. Appl. Behav. Anal. 10, 75-84.
Panyan, M., Boozer, H. and Morris, N. (1970). Feedback to attendants as a reinforcer for
applying operant techniques. J. Appl. Behav. Anal. 3, 1-4.
Shoemaker, J. and Reid, D.H. (1980). Decreasing chronic absenteeism among institutional
staff: Effects of a low cost attendance program. J. Organizational Behav. Management 2,
317-328.
Gardner, J.M. (1970). Effects of reinforcement conditions on lateness and absence among
institutional personnel. Ohio Res. Q. 3, 315-316.
Canter, D. and Canter, S., eds. (1979). Desigl1ingfor Therapeutic Envil'Onmel1ts. John Wiley
& Sons, New York.

IV. PROFESSIONAL AND LEGAL ISSUES

14. PROFESSIONAL ISSUES IN NEUROPSYCHOLOGICAL


REHABILITATION

LANCE E. TREXLER

1 INTRODUCTION

Recently, the brain-injury and neuropsychological-rehabilitatiori segments of


the health-care industry have grown extensively. For the purposes of this chapter, the term neuropsychological rehabilitation refers to a constellation of services
that are designed 1) to improve the nonphysical, cognitive, language, and perceptual functions dependent on cerebral integrity; 2) to influence the patient's
psychological adaptation; and 3) to promote the patient's environmental reintegration. Professional health care workers have responded to this increase in
demand with a proliferation of services. The main emphasis of this chapter
is 1) to briefly examine the social trends that are contributing to the growing
neuropsychological rehabilitation market; 2) to determine the impact those
trends have had on the professional's ability to provide services to that market;
and 3) to explore some of the ethical issues and questions that have naturally
arisen in the early evolution of this field.
This chapter will not attempt to address all of the current professional and
ethical issues in neuropsychological rehabilitation. Rather, it will focus on considerations that seem most important for the future of neuropsychological rehabilitation. The various issues will be discussed largely from a critical standpoint,
based on the assumption that it is better to be conservatively judicious in how
we evaluate our practices and methods of neuropsychological rehabilitation
The author would like to sincerely thank Leonard Diller, PhD, for his numerous suggestions and his time in
reviewing this nlalluscript.

363

364

IV. Professional and Legal Issues

than to be blinded by prejudiced assumptions. Of course, whenever possible,


our practices should be guided by experimental validation. This posture is also
based on my belief that the clinical value and importance of the profession of
neuropsychological rehabilitation is worth protecting.
2 SOCIAL DEVELOPMENTS IN BRAIN INJURY REHABILITATION

The health-care industry in general has witnessed major changes over the last
two decades. Major changes in the complexion of health-care delivery systems
and insurance companies have occurred as a result of the utilization of diagnosisrelated groupings (DRGs) and preferred-provider organizations (PPOs), significant increases in health care costs and the corresponding request of employers
for less expensive insurance policies, and an increase in catastrophic case management systems. Simultaneously, the need for specialized services for the braindamaged has grown. A number of factors, including increased survivability,
legislation regarding rehabilitation, and reactions of the insurance industry,
have all contributed to the increase in demands for neuropsychological
rehabilitation.
2.1 Consumer Need

Traumatic brain injury (TBI) survivors and their families obviously have an
incentive to obtain services that will benefit their quality of life. The residual
behavioral, personality, neuropsychological, and physical difficulties following
brain injury often devastate family stability, consume financial resources, and
limit the patient from engaging in many productive activities. There is a probability that these difficulties (and their corresponding functional, vocational,
and avocational implications) will be life-long. Because many of the survivors
of TBI are relatively young, the potential duration of disability becomes
emotionally and economically devastating. These factors often lead families
to vehemently expect and pursue services.
2.2 The Insurance Industry

The insurance industry often has a substantial financial incentive to see that
function is restored in the brain-injury survivor. In many cases, insurers will
be responsible for paying disability income, workman's compensation, or
future medical expenses. Many cases involving trauma can easily represent
a million dollars or more in medical costs and disability payments.
Through case management, many insurance carriers actively manage highrisk cases so as to reduce long-term liability through restoration of the patient's
functioning. The more foresighted and progressive insurance carriers take an
active role in the management of brain-injury cases (case management), whereby they attempt to ensure that the patient receives the best possible care. This
approach follows from the belief that superior care minimizes long-term disability and, hence, long-term financial expense. Unfortunately, this approach
is more characteristic of carriers who are liable for long-term benefits, as com-

14. Neuropsychological Rehabilitation

365

pared to health insurance carriers (who are solely obligated to cover standard
and customary "medical" services). Further, some health-care insurance companies have implemented "case management" programs to control the referral
process so that the least expensive care is provided, regardless of the quality.
Although the most devastating disabilities following brain injury are often behavioral and neuropsychological, they are not likely to be considered "medical"
in nature by traditional health insurance carriers. As a result, the carriers avoid
reimbursement for services.
In many states, this situation has been complicated by legislative action. Some
states have enacted mandatory rehabilitation bills, which require the insurer to
reimburse rehabilitation services. Although legislative control has minimized
the probability of depriving brain-injury survivors and their families of necessary services, it has also laid the foundation for potential abuse by the healthcare industry. The so-called no-fault insurance states provide a fertile ground
for all patients referred to a facility to be found "appropriate" for admission,
and length of stay is often unlimited, as long as the patient is showing some
evidence of "progress."

3 IMPACT ON HEALTH CARE MODELS AND PROVIDERS

Rehabilitation practices designed in the 1940s to address the needs of patients


with predominantly orthopedic injuries represented a substantial improvement
in the centralization and integration of such services relative to a preexisting
fragmentation and very limited supply. Because of a more recent increased survivability among brain-injured patients-particularly those with head trauma,
but also those with brain damage secondary to infection, cardiac-related hypoxia, and vascular and metabolic insults-inpatient rehabilitation services
had to be modified to take into account the patients' neurobehavioral and neuropsychological impairments. Unfortunately, rehabilitation professions recognized the needs of the patient long before adequate models for staff organization
and professional roles had been developed. Feelings of inadequacy and ignorance promoted professional competition, jealousies, and territoriality. New
ideas often met with resistance as they confronted inadequate but familiar
professional roles.
The increase in survivability, the absence of appropriate program models,
and the new demands on professional roles have led to considerable variability
in rehabilitative outcome and have occasionally taxed the mental stability of
rehabilitation staff. The preparedness of the various professionals and the healthcare industry is still inadequate in many respects. Without careful, self-imposed
quality assurance-as well as peer review and continuing research-the credibility of neuropsychological rehabilitation may in some ways be compromised,
much as the initial history of clinical biofeedback was seriously damaged by
professional and nonprofessional abuses (predominantly through marketing
and distribution of technology without professional training and review).

366

IV. Professional and Legal Issues

3.1 Development of Specialized Services

Brain-injured people have difficulties with long-term adaptation and social reintegration. These difficulties have presented significant challenges for the health
care industry, particularly the need for diversification, reconceptualization,
and clarification of professional roles.
The early 1980s saw the development of inpatient units specializing in brain
injury, with a concomitant recognition of the patient's need for post-inpatient
care. Outpatient services with various orientations (e.g., vocational, transitional living, neuropsychological, family, and psychotherapeutic) were also
developed, but they did not easily fall within the medical model-or within
the reimbursement structure of most health care policies. These trends made a
substantial impact on the demand for rehabilitation professionals, particularly
psychologists.
Before the organizational structure and models for professional roles for the
newly developed inpatient and outpatient facilities had been established, most
programs paved their own road. The most comparable models of health care
delivery were in mental health care, but the applicability of traditional psychiatric services to the behavioral, neuropsychological, and environmental reintegration needs of the brain-injured was quite limited. However, the goals of
outpatient brain-injury programs were less medical and were more concerned
with behavioral and environmental issues. Leadership in the field was established on the basis of experience and training-or perhaps in some cases by
mere availability-rather than on the basis of the appropriateness of the profession. Little or no formal research was available to the practitioner regarding
what types of patients would benefit from what kinds of service, what outcomes
might be expected if patients were treated or not, or the efficacy of specific
interventions (e. g., cognitive rehabilitation techniques, psychotherapy). This
state of affairs is still far from being resolved, despite a quantum leap in
available information.
3.2 Professional Training

Only recently have different models of training and definitions of professional


roles for clinical neuropsychology been described [1]. Although approximately
23% of neuropsychologists who responded to a survey ofthe Clinical Neuropsychology Division of the American Psychological Association and the National
Academy of Neuropsychologists indicated that they engaged in rehabilitative
activities [2], no training guidelines have yet been developed for "interventive"
(rehabilitative) neuropsychology. Nevertheless, many rehabilitation professionals have arbitrarily adopted the professional title of "cognitive therapist."
Although cognitive therapy may be the approach that the professional embraces, the title "cognitive therapist" represents neither a professional degree
or course of postgraduate studies, nor a certification or license awarded by
an independent professional organization. At this point, neuropsychological

14. Neuropsychological Rehabilitation

367

rehabilitation remains a self-proclaimed profession, the nature of which is


determined by the individual.
The long-term consequences of this state of affairs have not yet been assessed;
however, the unfortunate history of clinical biofeedback should serve as a warning. More specifically, the early experimental findings that humans had the
potential to volition ally alter certain physiological processes (such as heart rate)
that were previously thought to be outside of conscious controlled to a premature proliferation of clinical applications. Biofeedback equipment was soon
available through the classified advertisements in many magazines. The equipment was a quick answer to an old problem. The effectiveness of the procedure
was attributed to the equipment, not to the theoretical or clinical sophistication
of the user. Because of the ensuing negative reputation, many insurance companies will still not recognize biofeedback services, despite more recent research
that demonstrates actual benefits.
Neuropsychological rehabilitation includes many domains of knowledge,
some of which have not been fully defined. No one discipline can claim to encompass them all. For example, professionals in this field need a firm foundation
in clinical and experimental psychology, but they also need a clear understanding of functional neuroanatomy, neuropathology, clinical and experimental
determinants of recovery, methods of assessment of brain-behavior relationships, and psychopharmacology. In addition, they must be aware of a host of
functional, vocational, and family reintegration issues that are unique to survivors of brain damage. Because there are no formal academic programs in rehabilitative neuropsychology, professional training in these areas must generally
be self-directed. Still another complication is that the scope of neuropsychological rehabilitation is continuously being redefined as a result of the explosive
progress in the behavioral neurosciences. No wonder the obstacles to achieving
expertise, or even competence, can seem overwhelming.
3.3 Proliferation of Services

In response to the demand for services generated by consumers and the insurance industry, the number of facilities or programs specializing in treatment
and rehabilitation of brain-injury survivors has grown exponentially over the
last decade. These programs can be inpatient, outpatient, or residential and can
focus on cognitive goals, vocational goals, or independent living goals-or
some combination thereof. Even more recently, some facilities have targeted
as their mission the management of coma or of minor head injury. The National
Head Injury Foundation (NHIF) reports that in 1980 there were 12 facilities
specializing in head injury, whereas in 1987 there were 618 (NHIF, personal
communication) .
Certainly the development of new services for a relatively new problem is
justified. For instance, when polio became a prevalent disease, treatment and
rehabilitation facilities were established throughout the country in relatively
short order. Considerable funds were allocated for basic scientific research and

368

IV. Professional and Legal Issues

development of treatment facilities. The costs of brain injury in terms of quality


oflife, lost vocational and economic productivity, disability benefits, and health
care are staggering.
The complaints of patients and their families who are seeking treatment are
vast and typically concern not only patients' psychological change, but also their
loss of neuropsychological function, their inability to live independently, and
their inability to obtain or maintain gainful employment. Significant others
often report that the patients' physical changes and neuropsychological losses
are more devastating than their losses of physical function. Many families have
indiclted that the person whom they had known "died" as a consequence of the
brain damage, but the body lived. This separation between mental and physiological functioning presents a schism unique to the disorder of brain damage.
Phenomenologically, brain damage can be an indefinite loss of self, concomitant with being left to survive. This tragedy certainly warrants all of the efforts
that our society can mobilize, at both the basic science and clinical levels.
The proliferation of brain-injury services has not been universally advantageous to the achievement of research and clinical goals. In some cases, the
development of brain-injury programs has been solely based on a niche in the
economic market. The onset of DRGs resulted in many empty beds, and hospital administrators have needed to find new services through which occupancy
could be recovered. Consequently, because brain-injury rehabilitation was a
growing market, and because inpatient rehabilitation was exempt from DRGs,
hospital units were hastily converted into ones that could serve the braininjured. A variety of outpatient and residential programs have been developed
specifically for the insurance industry, particularly the workman's compensation and liability carriers. The distinction between a clinically driven facility and
an economic-market-driven facility can be significant in terms of allocation of
resources, admission and discharge criteria, and level of professional preparation. From a business standpoint, there are certainly enough "customers";
families with a real problem and desperate for some help, and the professionally unsophisticated insurance companies with an economic problem (i.e., the
patient's long-term disability) on their hands. The viability of such an approach
(i. e., driven by the economic market) is apparent from the continuing proliferation of corporate facilities and programs. It is not clear, however, how these
rapidly growing programs can adequately train staff members at a corresponding rate. Adequate training in how to work with this population takes several
years, as well as considerable human resources. How, then, can a program add
300 staff members a year? Furthermore, few of these programs are based on
research concerning clinical determinants of recovery, predictors of benefit
from services, and treatment efficacy.
3.4 The Role of Differential Diagnosis

Professionals have a tendency to label pathology according to the perspective


they were taught. This practice is fraught with hazard in a field like neuropsy-

14. Neuropsychological Rehabilitation

369

chological rehabilitation, in which multiple and integrated types of expertise


are required, especially because professional training and competence criteria
have not been established. Furthermore, many of the behaviors exhibited by
brain-injured individuals defy traditional classification schemes or may arise
from multiple causes. For example, many neuropsychological examinations
have attributed a patient's memory "deficit" (i.e., the score didn't fall within
the average range) to his or her history of brain injury, even though the actual
reason for the abnormal score may have been that the patient was clinically
depressed, didn't like the technician, or had an attention deficit.
Mental health professionals sometimes attempt to diagnose and pharmacologically treat the brain-injury survivor within the traditional psychiatric model.
This is accomplished without benefit of knowledge about the neurobehavioral
effects of certain types of brain lesions.
The following case history further illustrates the difficulties involved in
differentiating the cause of any given symptom:
John was a 27-year-old man who had been transferred to our outpatient program from an out-of-state hospital. For five months following his severe TBI,
he had been treated in an inpatient rehabilitation unit; for the next three months,
he had been treated in the hospital's outpatient rehabilitation facility. John received a comprehensive evaluation at our center, and one of the findings was
that he loudly gasped and wheezed after he inhaled. His other symptoms included dysarthria, diplopia, and signs of damage to the brainstem. Medical records from a psychiatric consultation suggested thatJohn had suffered terrifying
emotional trauma in response to the accident even while he was unconscious.
(He had been in a deep coma for 21/2 weeks and suffered from retrograde and
posttraumatic amnesias.) As a result, he appeared to be "gasping for life." His
family was quite distressed that he had developed such a significant "behavioral"
problem, and family members were always blaming him for this behavior.
Our outpatient evaluation suggested a different hypothesis as to the cause of
the "gasping" behavior following inhalation. Because John exhibited a constellation of other oral and pharyngeal symptoms, we suspected that he suffered from a bilateral abductor-fold paralysis. We requested consultation with
a specialist in otorhinolaryngology. However, two days after the diagnostic
evaluation, John's gasping worsened to the point that he developed significant
closure of the airway, and the status of his life-supporting airway became questionable. He was taken to an emergency room where he received a tracheostomy, which probably would be permanent. Endoscopic examination suggested the presence of substantial amounts of scar tissue under the vocal folds,
which held them in a permanent, midline position. This had occurred as a result
of the prolonged placement of the breathing tubes while John was comatose.
I do not mean to imply, by my use of this example, that psychiatry is any
more vulnerable to misdiagnosis than is any other profession. Many patients
"fall between the cracks" of neurology, psychiatry, physiatry, and neuropsychology. The brain-injury survivor's symptom complex has interrelated com-

370

IV. Professional and Legal Issues

ponents that sometimes cross the boundaries of many different disciplines,


despite the fact that the varying symptoms can arise from the same underlying
pathology. The resulting clinical diagnosis and subsequent treatment can be
fragmented, and clinical care can be compromised.
3.5 Psychological Struggles and Conflicts of Staff Members

Because of the complexities of diagnosis, the expectations of patients and their


families, and the range of information that professionals must know and apply,
health care providers are placed in a very demanding and humbling position.
True, a professional's awareness that he or she lacks complete information
about patients (or about how to assist their adaptation and enhance their quality
of life) is a prerequisite to asking the right questions. However, the unremitting number and the increasing scope of questions can leave a professional
with feelings of inadequacy, defensiveness, depression, and anxiety. Over
time, the excitement of embarking on a new endeavor can be re'placed by
pessimism and discouragement.
Many staff members project these feelings onto the facility in which they
work or onto the patients and their families. The overt manifestations vary as
a function of the individual dynamics, but include professional territoriality,
setting of unrealistic expectations for themselves and their patients, or countertransference-identification and enmeshment with the patient. When families
are viewed as unsupportive or as inhibiting the rehabilitation process, it is
often a projection of the professional's feelings oflack of support or dislike of
the patients or their behavior. The management of neuropsychological rehabilitation staff often needs to be viewed as a process of group psychotherapy,
facilitation of the growth of the milieu, and continuing education; these
processes are as important as the more traditional personnel issues such as
performance and salary reviews and hiring and firing practices.
3.6 The Importance of a Program PhiIosopp.y

The most effective methods for managing and preventing staff problems are to
control the definition, revision, and implementation of the program philosophies. How the philosophies and priorities of the program are implemented
defines the culture in which the staff must work. Management and clinical
leadership staff are responsible for anticipating and managing the culture of the
rehabilitation program. The quality of the culture in a rehabilitation program
has certain significance for how the psychological struggles of the staff are
handled, which in turn influences patient care.
The first step for management and clinicallcadership staff is to jointly define
a program philosophy. Defining a program philosophy includes such issues
as quality and quantity of services, type of care (e. g., inpatient/outpatient,
behaviorally oriented, residential), determination of clinical priorities, commitment to research and training, and financial allocations. The program philosophy defines what the organization is and what it does. The assistance of a

14. Neuropsychological Rehabilitation

371

program consultant who is a professional in the field of neuropsychological


rehabilitation can be invaluable in this process.
Without an agreed-upon program philosophy, the stage can be set for staff
conflict and fragmentation of services. A program philosophy, and the process
of defining and revising it, can serve to unify the staff, build morale, and drive
both management and clinical decision-making. Program philosophy also
drives organizational structures and hierarchies, as well as job descriptions.
However, to derive a working and viable program philosophy and reap the
benefits of an integrated culture, all staff-from the receptionist to the
president-must share in at least some way the process of formulating the
program philosophy.
The second step for management concerns the implementation of program
philosophy. The primary method of implementation is to get staff "singing off
the same song sheet." Therefore, staff recruitment and training are a central
priority for program managers. When recruiting staff, it is at least as important
to consider the personality, professional priorities, and beliefs of the potential
staff member (especially as these factors mayor may not fit with the program
philosophy) as it is to consider their professional skills. These individual factors will obviously have a great deal to do with how each potential worker
might interact with co-workers and patients alike.
Staff selection should then be followed by including staff in the implementation and maintenance of the program philosophy. Implementation of program
philosophy can be monitored through regularly scheduled staff meetings, in
which minutes of the meeting, plans for action, and a corresponding timetable
for accomplishing the plans for action are maintained in a program development manual. This manual should also be maintained by all staff members,
with a section in which ideas and suggestions can be entered and used for discussion at staff meetings. Most important is the monitoring of priorities and
allocation of resources, both time and financial, to ensure that they coincide
with the stated program philosophies. Staff/patient ratios, clinical productivity
standards, allocated staffing, team meeting time, and educational resources are
also driven by the implementation of program philosophy.
A third aspect of staff management worthy of mention concerns the development of the rehabilitation program culture. As mentioned, the quality of the
culture determines the extent to which staff can effectively resolve conflict,
make decisions, and develop an effective team. The optimal culture provides
the staff with an environment in which they can learn and grow professionally;
this can enhance the quality of service provided by the program. Professional
development requires an environment in which it is acceptable for the individual staff member to not know everything and in which he or she can trust
other staff to give professional, productive feedback. Without trust and support
as integral aspects of a program culture, professional development is muted or
stagnated, and the possibility for the team to utilize the strengths of various
professionals is minimized.

372

IV. Professional and Legal Issues

A healthy program philosophy and culture permit internal and external


resources to be utilized for staff development and training. Individual staff
members can openly share their areas of expertise (and areas of weakness) as
well as summaries of professional training obtained outside the facility. Formal
and informal research, case reviews, and clinical staffings can be placed in context of the program philosophy. In the absence of a systematic vehicle for
discussing and modifying the program culture, an unclear or unhealthy program culture breeds discipline territoriality, a hesitancy to review cases, and a
tendency to withhold information.
4 ETHICAL ISSUES

The development of neuropsychological rehabilitation as a viable segment of


the health-care industry necessitates a corresponding development of ethical
approaches to professional conduct. At the moment, each professional is guided
by the ethical standards adopted by his or her professional association. Not all
situations encountered in the field of neuropsychological rehabilitation will be
addressed by the existing ethical standards. However, the long-term future of
neuropsychological rehabilitation will be influenced by the actions of today's
professionals.
For example, as previously discussed, the concepts and methodologies
employed in clinical biofeedback were quite viable. However, its credible
utilization was interrupted due to an overemphasis on technology (without
sufficient clinical training), inadequate research on efficacy, and indiscriminate
marketing and sales. Neuropsychological rehabilitation as a field is at risk for
comparable reasons. On the other hand, many facilities and programs have
applied professional ethics to their practice and have carefully documented
their interventions, selectively admitted patients based on admission criteria,
conducted formal program evaluation, obtained peer review and consultation,
and judiciously represented their services to the public. An exclusively marketdriven program is less likely to consider admission criteria or to fund programevaluation research.
It would seem likely that the development of the neuropsychological rehabilitation segment of the health-care market will parallel the evolution of
other new markets, in which there is an initial proliferation of the service
or product, the emergence of competition, and ultimately a "survival of the
fittest." With increasing sophistication among consumers-both patients and
reimbursement sources-those professionals, facilities, and programs that
consider and apply ethical considerations to their practice would seem the
most likely to endure beyond the momentary triumph of success in a new
market.
4.1 Selling Hope

Disseminating information to consumers of neuropsychological rehabilitation


services is important so as to educate referral sources about available ser-

14. Neuropsychological Rehabilitation

373

vices, about what constitutes an appropriate referral, and about what can
reasonably be expected from intervention. A professional representation of
services can be contrasted with "selling hope." In the extreme form of the
latter, some facilities may choose to obtain the greatest possible market share
of potential patients through direct sales to patients and their families. In some
of these facilities, a major portion of the operating budget is allocated to sales
and marketing. Families and patients are not necessarily educated consumers
of neuropsychological rehabilitation services, but they are particularly interested in any representation that offers hope. It is of course incumbent upon
professional staff to differentiate rehabilitation potential from false hope.
Market-driven and finance-driven facilities may also be tempted to diversify
services in response to consumer needs, even if their staff members may not be
qualified to provide the additional services. "Filling the void" before a
competitor does has become standard operating procedure for many healthcare corporations. This approach can also leave professionals with feelings of
conflict about what is expected of them in terms of provision of services and
professional training and repertoire. Sustained conflict of this kind will ultimately erode their professional careers, lead to high personnel turnover, and
have a negative impact on the field of neuropsychological rehabilitation.
4.2 The Insurance Company Referral

Many reimbursement sources are committed to utilizing resources that not


only may be able to enhance the patient's quality oflife and level of adaptation,
but that also may assist the carrier (by shortening the amount of time that the
carrier will be financially liable for the patient's care). These circumstances
would appear to be "the best of all possible worlds": The insurer is truly invested in the patient's recovery, and there is the potential to minimize financial
losses.
Because it offers guaranteed reimbursement, the insurance company referral
is very attractive to most providers. In these circumstances, however, the provider must keep separate the priorities of the insurance company and the clinical
needs of the patient. Clinically indiscriminate admissions will ultimately not
meet the needs of the referring insurance company: If the provider cannot
adequately serve the patient, the patient will not benefit in a way that enhances
independence (and therefore will not reduce the insurer's long-term liability).
Consequently, the patient will have been inappropriately served, and the credibility of the provider will have been questioned. Furthermore, admitting
patients for short-term financial gain will have negative long-term financial
implications for the provider. If the provider does not have the competence
(see Section 4.3) to offer the needed service, the patient should be referred
elsewhere, despite the immediate loss of financial gain to the provider. This is
not only the right choice on ethical grounds, it is also the right choice in terms
of the institution's long-term self-interest.

374

IV. Professional and Legal Issues

4.3 Determination of Competence and Representation of Services

No one profession encompasses the breadth of knowledge and experience


necessary to address all of the rehabilitative needs of brain-injury survivors.
Diagnosis and rehabilitation of the brain-injury survivor's neuropsychological
defects and functional adaptation requires a team of professionals. Each rehabilitation professional has his or her own characteristic area of specialization
(although areas of expertise often overlap). However, human behavior is determined by many, very dynamically interrelated factors, and the "responsibility"
for dealing with various aspects of a patient's functioning is frequently divided
along professional lines, rather than according to functional lines.
Who is competent to provide neuropsychological rehabilitation services?
Given the absence of any existing professional standards, the answer to this
question might necessarily rely on the domains of knowledge and experience
integral to the practice of neuropsychological rehabilitation. Neuropsychological rehabilitation certainly requires careful, case-by-case integration of a
team of professionals, including physicians, neuropsychologists, and occupational, physical, and speech therapists, among others. However, for rehabilitation to be referred to as "neuropsychological," neuropsychologists must have
significant responsibility not only for evaluating potential for benefit from
rehabilitation but also for determining clinical priorities, methods, and goals
for such efforts.
For neuropsychologists, competence in neuropsychological rehabilitation
might be considered to be the result of four levels of experience and training:
1. Academic preparation and supervised clinical experience in diagnosis and
treatment of psychopathology and other behavioral disorders, in accordance with the curriculum of approved doctoral programs in clinical
psychology;
2. Academic training in cognitive psychology, psychobiology, human
neuropsychology, functional neuroanatomy, and neuropathology;
3. Clinical experience in medical rehabilitation;
4. Experience in integrating neurological, social, and vocational histories
with findings from neuropsychological examination and formulating practical, obtainable, and theoretically sound interventive recommendations.

It is incumbent upon relevant professional associations to establish professional definitions and criteria for the practice of neuropsychological rehabilitation; on the basis of name alone, this would at least concern the Division of
Clinical Neuropsychology of the American Psychological Association. For the
moment, however, professionals might want to compare themselves to the
levels of competence suggested above when evaluating their preparedness to
engage in services pertaining to the rehabilitation of neuropsychological disorders. For non-neuropsychologists, a program of neuropsychological rehabilitation needs to include consultation with a qualified neuropsychologist.

14. Neuropsychological Rehabilitation

375

The risks of an individual's or organization's failure to consider professional


competence when addressing the issues inherent in treating brain damage outweigh any possible advantages that might be gained from utilizing incompetent staff. For example, over the long term, the credibility of the individual's or
facility's neuropsychological rehabilitation services will be questionedgenerically and specifically. Staff members' morale declines when they feel
incompetent, and interpersonal conflict emerges. Staff turnover aggravates
the situation, and quality of care can further decline. For reasons like these, a
judicious and conservative posture on competence seems warranted.
Also related to the professional competence questions is the difficulty of
adequately defining neuropsychological rehabilitation services to insurance
carriers. Terms like cognitive therapy and memory therapy do not denote specific
professions, nor do they describe services that are typically recognized by health
care insurance. Reimbursement is therefore difficult to obtain.
Some providers have chosen to use traditional nomenclature, such as occupational therapy for cognitive remediation. This may be appropriate if the service
is provided by a qualified occupational therapist and falls within the domain of
services that the professional organization recognizes as being within the competence of the profession of occupational therapy. On the other hand, if what
is being provided clinically is in fact "memory therapy," the use of the term
occupational therapy is a misrepresentation and is therefore a questionable ethical
practice.
Recognition of these services (e.g., memory therapy and cognitive therapy)
by the insurance industry depends on 1) results of research on the efficacy of
the services; 2) acknowledgement that such services arc as "medically necessary" as any other rehabilitation therapy; and 3) inclusion of the services as
part of the insurance policies' benefits. At the moment, it remains incumbent
on rehabilitation providers to justify their definition and representation of services. The definitions of more traditional services should be reconsidered;
the further the actual clinical service departs from this definition, the more
necessary it becomes to use new vernacular.
5 THE FUTURE OF NEUROPSYCHOLOGICAL
REHABILITATION: ACCOUNTABILITY

The future progress and development of neuropsychological rehabilitation (as


a constellation of services separate from mainstream medical rehabilitation)
will largely depend on directions in professional definitions and standards,
clinical research, and the responses of the insurance industry.
5.1 Professional Definitions and Standards

Definitions and standards for the practice of neuropsychological rehabilitation


will eventually be set-through legislation, or enactments of professional
organizations, or some combination of these. The rapid diversification and
growth of the numbers of professionals and services has historically resulted in

376

IV. Professional and Legal Issues

greater regulation of both. It would certainly be better if the development of


such definitions and standards were initiated by professionals (and their corresponding professional organizations), rather than by external pressures (such
as legal precedents or legislation). The associations representing professionals
in rehabilitation obviously have the most interest in proactively establishing
such definitions and standards, primarily for the protection of their constituents
and the welfare of the patients for whom they provide services.
Academic training programs will need to more thoroughly address the spectrum of knowledge needed by the professional. Postgraduate courses are also
needed, to refine the expertise of professionals who are already providing neuropsychological rehabilitation services. Ongoing advances in the neurobehavioral
sciences will necessitate reconsideration of many current practices and beliefs.
At some point, some form of certification or licensure (or both) will probably emerge-again, with the aim of protecting the consumer as well as the
provider.
5.2 Clinical Research

Although today neuropsychological rehabilitation research is substantially


more extensive than that of even five years ago, the need for more treatmentefficacy and program-evaluation research still outweighs the progress (e. g.,
r3, 4]). Services that attempt to train brain-injured patients to achieve some
specific "cognitive skills" will need to empirically demonstrate the overall
functional impact of the investment of time, money, and effort. Those programs that employ expensive day-treatment models must justify that the costs
are warranted in terms of clinical benefit. Many issues surrounding the empirical basis for admission and discharge criteria, prediction of recovery and
methods of intervention are yet to be resolved.
Given the recency of the field of neuropsychological rehabilitation, it would
seem important, on ethical as well as professional grounds, that clinicians
engage in empirical outcome or quality-control research as a part of their practice. Because federal mandates for establishing this line of research as a priority
are just emerging, and because resources are still very limited, individual providers should consider initiating such activities themselves. In a field that is just
beginning to stabilize, peer review-of research efforts as well as of clinical
progress-becomes an important component of the process.
5.3 The Insurance Industry

As professional sophistication grows and as naivete and fads are replaced by


research findings, some current approaches to neuropsychological rehabilitation (as well as some programs) will survive, and some will not. The insurance
industry is likely to close the knowledge gap quickly, and insurers have already
begun to show signs of becoming better consumers of neuropsychological
rehabilitation services. The insurance industry (health-care as well as liability
and workman's compensation divisions) will increasingly require the provider

14. Neuropsychological Rehabilitation

377

to be accountable for the determination of cost per benefit of various services


and programs. The practices of programs that have admitted patients based
solely on reimbursement variables have already alerted some insurance industry experts about potential economically driven abuses-both clinical and
professional-in neuropsychological rehabilitation.
6 SUMMARY

Pressure to be accountable is not undesirable; it can be productively put to use


as an impetus to improve the profession. I hope that, as a result of external
pressures and awareness of the professional issues surrounding neuropsychological rehabilitation, our field will evolve and that we will responsibly review
the ethics of the services we provide. Peer review, program evaluation, consideration by professional associations, and research can protect the future
credibility of neuropsychological rehabilitation-as well as the patients for
whom the concepts were derived.
REFERENCES
1. Meier, M. (1984). Report of the Division 40/INS Joint Task Force on Education, Accreditation, and Credentialing (Chair: M.J. Meier). Newsletter 40 2, 3--8.
2. Seretny, M.L., Dean, R.S., Gray, j.W. and Hartlage, L.c. (1986). The practice of clinical
neuropsychology in the United States. Arch. Clin. Neuropsychol. 1,5-12.
3. Meier, M.J., Benton, A.L. and Diller, L., eds. (1987). Neuropsychological RehabilitatiOl1. Churchill
Livingstone, London.
4. Prigatano, G.P., ed. (1986). Neuropsycholof?ical Rehabilitatioll After Brain Injury. Johns Hopkins
Press, Baltimore.

15. LEGAL ISSUES THAT COMMONLY CONFRONT BRAIN-INJURY


SURVIVORS AND THEIR FAMILIES

SIMON H. FORGETTE

1 INTRODUCTION

Medical treatment of brain injury as we know it today is relatively new. Ten to


20 years ago, emergency medicine was usually unable to keep people with
severe brain injuries alive. On the other hand, those victims with so-called
"minor" brain injuries (which nevertheless produced cognitive as well as other
types of impairment) frequently went undiagnosed. Even if deficits caused by
minor brain injury were recognized, many cases went untreated because of the
lack of rehabilitation programs that took into account the special needs of
these patients. Today, however, medicine acknowledges the epidemic proportions of brain injuries and continues to improve the efficacy of treatment
and rehabilitation for this population.
The law is not far behind in rendering its own aid to survivors ofhead/brain
injury. Many problems that survivors and their families face have legal solutions. However, the law does not have an answer-at least a good answerfor every problem.
This chapter has been written primarily for health care providers, survivors,
and their families. To assist survivors and family members, references have
been placed directly into the text. Where the name of a law case is given (such
as Smith vs. Jones), the citation following the case name will permit a librarian
at any county law library to direct someone to that case for review.
Because this chapter is meant to be of general use in the United States, its
scope is necessarily broad. However, attorneys who are new to the representation of brain-injured people may also find this chapter informative.
379

380 IV. Professional and Legal Issues

The text is not intended to be a substitute for competent legal advice in the
reader's own jurisdiction. Rather, the purpose of this chapter is to give the
reader an awareness and some foundational understanding of the legal issues
and problems that are commonly encountered by victims of brain injury.
Legal issues and problems will be addressed in the order in which they
usually arise for the survivors and their families.
2 WHO WILL PAY THE MEDICAL BILLS?

The National Head Injury Foundation recently adopted the following definition of traumatic head injury:
Traumatic head injury is an insult to the brain, not of a degenerative or congenital
nature but caused by an external physical force, that may produce a diminished or
altered state of consciousness, which results in impairment of cognitive abilities or
physical functioning. It can also result in the disturbance of behavioral or emotional
functioning. These impairments may be either temporary or permanent and cause
partial or total functional disability or psychosocial maladjustment. 1

Families are never prepared for the emotional, psychological, and financial
roller coaster that a family member's brain injury causes them to ride. In cases
of severe brain injury, the entire family is commonly so traumatized that, for
lengthy periods of time following the injury, no one is able to "take care of
business"-the business of exploring all potential sources for the payment of
medical bills.
Most individual health-care providers think first about treating the sick and
secondly about receiving compensation for their services. However, there is
nothing inappropriate about a health-care provider's exploring the possible
sources of medical-bill payment available to survivors. Many times, the
health-care provider is able to help direct survivors and their families toward
potential sources of funding. When this assistance leads to payment of those
bills, it also has the therapeutic effect of relieving some of the psychological
burden carried by the survivors and their families.

Note to Families: Health-care providers usually refer to the period of weeks or


months after an injury-during which medical treatment is rendered to stabilize the patient and maximize his or her recovery as quickly as possible-as the
"acute phase" of posttraumatic medical care. Sources of funding for such acute
care are discussed following this note. However, families should be aware that
funding for long-term rehabilitation or "postacute" care is much more dif...,
ficult to obtain. Families should be aware that many brain-injury rehabilitation
programs have employees who will assist the survivors and their families in
attempting to find funding. The National Head Injury Foundation (333 Turn1 Adopted by the Executive Committee of the Board of Directors of the National Head Injury Foundation, Feb.
22, 1986.

15. Survivors and Their Families

381

pike Rd., Southboro, MA 01772; phone: 508-485-9950) or its local state


chapter may also be able to provide families with information about the
funding of long-term care from government sources.
In most cases, one or more of the sources listed below (sections 2.1 through
2.10) will pay for acute medical treatment and/or other damages or loss.
2.1 Health Insurance

Health insurance may pay all or part of a survivor's medical expenses after a
"deductible" amount that the patient must pay first. Young survivors may not
be aware that they are covered under health plans that their parents have either
purchased directly or that provide coverage through an employer. Although it
is not common, a parent's health plan may have a conversion option allowing
the inclusion-on the payment of an additional premium-of children who
may not have been covered before a disability. Each health plan has its own
policy language and must be studied carefully to ensure that its potential is not
neglected; I advise consulting an attorney to do this. The first step, therefore,
is not to assume (as so many parents of teenage survivors who have moved out
of the home prior to injury have) that a parental health plan is inapplicable, but
rather to have any and all health plans in the family carefully reviewed by an
attorney knowledgeable in insurance law (not just by the family insurance
agent or an agent of the insurance company) to determine what benefits may
be obtained.
A family member should be chosen to be responsible for submitting medical
bills and keeping a simple record of who paid each medical bill. A "master list"
should be kept, reflecting 1) each medical bill, including the name of the
health-care provider and the amount due; 2) when and to whom the bill was
submitted for payment; and 3) when and how much of the bill was paid.
Copies of all medical bills should be kept by the responsible family member.
The best practice is to write a short letter to accompany each set of medical
bills when it is submitted to the insurance company and to keep a copy of this
letter for the family's record. This letter should be dated, should make reference to the policy number of the insurance plan, and should list (by name of
provider and amount due) the bills that are being submitted at that time.
An organized method of submitting medical bills for payment should help
eliminate some of the stress that comes from receiving "late notices" and from
trying to figure out, months later, whether a bill was paid.
2.2 Accident/Auto Insurance

In cases of traumatic brain injury, whether health insurance coverage exists or


not, health-care providers may suggest that the survivors or their families
consult an attorney who is experienced in personal-injury law-ideally, one
who is experienced in handling brain-injury cases. Health-care providers (and,
for that matter, the survivors and their families) should not attempt to evaluate
the merits of a personal-injury case without the help of an attorney. The

382

IV. Professional and Legal Issues

attorney may be able to find insurance coverage that applies to cases in which
the facts seem hopeless, as in the following example:
A boy fell from a rope swing and struck his head against a car that was dangerously
parked and that was later driven away without being identified. Nevertheless, under his
parents' automobile policy, which provided protection against uninsured motor vehicles, the boy was able to receive compensation for the head injury he sustained.
Because memories fade, evidence is lost, and accident scenes change, it is
important that an attorney evaluate the potential personal-injury claim as soon
as possible. Personal-injury cases and the health-care provider's involvement
in the legal system are discussed later in this chapter.
2.3 Government Benefits-Social Security

The hospital social worker is usually the first person consulted about what
government benefits are available. Frequently, the social worker will have
names and phone numbers of "contacts" within the state or federal agencies
that administer these benefits. Going directly to the local Social Security
Administration office is also productive for the family in ascertaining rights,
benefits, and obligations under the Social Security Act. (The address of the
nearest Social Security office can be obtained from the local post office or from
the telephone directory, in which it is listed under United States Government,
Department of Health, Education and Welfare, Social Security Administration.)
There are three main federal programs that provide direct and continuing
financial and/or vocational assistance to disabled people: 1) Social Security
Disability Income (SSDI) benefits, under Title II of the Social Security Act; 2)
Supplemental Security Income (SSI) payments, under Title XVI of the Social
Security Act; and 3) Mdicaid, under Title XIX of the Social Security Act.
The Social Security Act is found in the United States Code. If you wish to
review the federal statutes regarding these programs, go to your local county
law library and ask the librarian to show you the volumes containing the US
Code. For SSDI, ask to see Title 42 of the US Code, beginning at Section 401,
For SSI, look at Title 42 of the US Code, beginning at section 1381. For
Medicaid, check Title 42 of the US Code at Section 1396.
Social Security Disability Income (SSDI) benefits will be paid to a disabled
worker and his or her family if earnings are lost or reduced due to the worker's
disability. To qualify for SSDI payments, the disabled worker must have
worked a certain amount of time at a job that paid into the Social Security
system. The amount of work credit needed depends on the worker's age. For
purposes of SSDI payments, the assets of the disabled worker do not matter.
Persons are considered "disabled" if they have a physical or mental impairment that 1) prevents them from working, and 2) is expected to last for at least
12 months or to result in death. This definition of disability applies to both
SSDI and SSI benefits.

15. Survivors and Their Families

383

Supplemental Security Income (SS!) payments are made monthly to disabled


people who have limited income and assets. A person does not need any Social
Security work credits to get SSI payments. People who have never worked
may be eligible for SSI payments. A disabled person and his or her family may
qualify for both Social Security Disability Insurance benefits and SSI payments.Unlike SSDI, the amount one can receive from SSI depends upon the
applicant's income and assets. However, a car or a home are not generally
counted as assets that would disqualify a person from receiving SSI payments.
Medicaid is a Title XIX-funded (42 US Code 1396), state-operated, medical
assistance program. It is most commonly referred to as "Medicaid," although
some states have changed or modified the name. For instance, in California, it
is known as "MediCal."
Whatever the name, Medicaid provides full coverage-that is, there is no
deductible amount that must first be paid by the survivor. If the survivor
qualifies, eligibility applies from the date on which a written application is
filed with the Social Security Administration. The application should therefore
be submitted to the local Social Security Administration office as soon as
possible.
To qualify, the survivor must meet the disability criterion for SSI and SSDI
stated above. A person who qualifies for SSI payments is also automatically
qualified for Medicaid. The SSI eligibility requirements changed in 1985, and
it is now much easier for people with head injuries to qualify for SSIIMedicaid.
Over half of the states have a second discretionary program for the payment
of medical bills; this is sometimes referred to as the "medically needy" or
"medically indigent" program. If such a program exists in the survivor's state,
it may pay medical bills even if the survivor has more in the way of assets or
income than SSIIMedicaid eligibility woulGl permit. In cases in which the
survivor has a significant amount of medical bills and no private insurance to
pay them, SSIIMedicaid should always be applied for, and an inquiry should
be made about whether the state has a discretionary fund that might provide
assistance with medical bills.
2.3.1 Estate-Planning Tip

An inheritance can disqualify a survivor from receiving SSIIMedicaid benefits.


Family wills should be reviewed by an attorney and changed so that the
survivor's potential inheritance is placed directly in a trust that is sufficient for
the purpose of maintaining SSI eligibility. (See the discussion about trusts in
Section 4.5.3 of this chapter.)
2.3.2 SSI-SSDI Return-to-Work "Incentives"

Both SSDI and SSI programs provide so-called "incentives" for those disabled
workers who wish to attempt to return to work. However, under certain
circumstances, workers who attempt to return to work may find that they are
actually making less on the job than they would be receiving under SSI or

384

IV. Professional and Legal Issues

SSDI, and further that-if unsuccessful in the attempt-they must reapply for
Social Security and Medicare/Medicaid benefits. Survivors who are receiving
Social Security benefits and who anticipate attempting to return to work
should carefully review Social Security work incentives with an expert in the
field of Social Security law, in order to arrive at a plan that will both maximize
the chances of returning to gainful employment and minimize the risk that, if
the attempt to return to work is unsuccessful, all or part of the Social Security
benefits will be lost.
2.4 Government Benefits-Vocational Rehabilitation

Vocational rehabilitation is usually administered by a state agency under a


program funded by state and federal governments through a system of matching funds. The regulations regarding the administering of vocational rehabilitation programs vary between states, but generally the following rules apply:
1. To qualify, a person must have a physical, mental, or emotional disorder
that constitutes a vocational handicap for that person.
2. Priority is usually given for those who have severe disabilities.
3. A reasonable expectation must exist that gainful employment will be
obtained if vocational rehabilitation services are provided.
4. Anyone is eligible to apply for services and receive and evaluation without
cost, regardless of the extent of disability or amount of income or assets the
person may have.
As a result of the 1985 reauthorization of the Rehabilitation Act of 1973, the
emphasis in government-sponsored vocational rehabilitation has shifted from
rehabilitation workshops toward on-the-job training or retraining in the community. There is, therefore, some cause for optimism that the survivors of
brain injury will be able to receive part of their vocational rehabilitation back
in their own former job or in a similar job setting in the community.
2.5 Government Benefits-The Importance of a Written Application

Regardless of what patients or their families are told regarding eligibility (over
the telephone or in person by the Social Security Administration or the state
agencies that work in conjunction with it), a written application form should
immediately be filled out and submitted. Most benefits, if granted, are backdated to only a few days before the written application is filed-not to the date
of the injury. A telephone inquiry is not considered an application. If the person
filling out the application does not have all the information requested, the
application should still be filed, and the additional information should be
provided as soon as possible. A copy of the application should always be kept,
and all discussions with (and advice from) Social Security employees should be
confirmed in writing.

15. Survivors and Their Families

385

2.6 Denial of Government Benefits-When to Appeal

A written denial of Social Security benefits contains a relatively short deadline


within which the decision denying benefits must be appealed. As a general
rule, a denial of benefits should be appealed immediately. Employees at the
local Social Security office may be willing to refer a survivor and his or her
family to an attorney who is experienced in Social Security law. Such an
attorney may be willing to review the survivor's case and give an opinion
about its merits for a nominal fee.
2.7 Government Benefits-Workman's Compensation

If the survivor was injured while on the job, a workman's compensation claim
should be filed. In some states, workman's compensation coverage is provided
by private insurance carriers. In others, benefits are paid from a state Workman's Compensation Fund, which is administered by a state agency. In some
states, employers are allowed to be self-insured regarding on-the-job injuries,
rather than pay premiums to an insurance company or pay into a state Workman's Compensation Fund. In such cases, an injured worker would make a
claim directly against his or her employer.
Information about the mechanics of making a workman's compensation
claim in any given state should be obtainable from the survivor's employer.
Generally, people who are injured on the job and are covered by workman's
compensation may not sue their employers. However, a survivor may have a
valid personal-injury claim, in addition to his or her workman's compensation
claim, against people or entities (other than the employer) whose negligence
helped to cause the survivor's injury.
Work-related injury claims by certain types of employees-for example,
federal employees, longshoremen/harbor workers, seamen, and railroad
workers-must be made within the framework established by federal statutes
for such workers.
2.8 Government Benefits-Crime Victims

Many states have passed legislation that sets up a fund from which victims of
crime may receive compensation for damages such as medical expenses and
wage loss. If a survivor has been injured by another person's criminal act, the
local prosecutor's or district attorney's office will be able to provide information about any such legislation that exists in the survivor's jurisdiction,
including what steps should be taken to obtain compensation from such a
fund.
2.9 The Party Responsible for the Inquiry

All too often, the person responsible for the injury has little, if any, insurance.
The person responsible should nevertheless be investigated to determine if any
reachable assets are available. Frequently, the survivor's own insurance com-

386

IV. Professional and Legal Issues

pany will run an assets check on the person responsible for the injury, and this
information should be made available to the survivor.
2.10 Other Sources for Benefits

The foregoing are the most likely sources of medical bill payments. However,
the survivor, health-care provider, family, and attorney should also consider
other local public assistance programs that may be available, as well as federal
programs available to certain survivors, such as veterans' benefits (CHAMPUS), the Railroad Retirement Program, or the Civil Service Annuity
Program.
3 IS THE SURVIVOR COMPETENT TO HANDLE HIS OR HER OWN AFFAIRS?

Family members and health-care providers are often concerned about the
legal predicaments that survivors find themselves in. Whether the survivor is
purchasing a gold mine, is requesting to drive a motor vehicle (see Section 6.1
of this chapter), or is in the process of being divorced by a spouse who assures
the survivor that the survivor does not need to consult an attorney, other
family members and/or the health-care provider may question the survivor's
ability to make an appropriate decision. The question then becomes the extent
to which the family member or health-care provider should become involved
in the matter.
The need for a court-appointed legal guardian is obvious in the case of a
survivor who is comatose or is otherwise clearly incompetent to handle his or
her own affairs. The difficult cases are those involving survivors who look
normal, walk and talk normally, and are capable of independent living, but
who suffer from deficits (usually from so-called minor head injury) that impair
judgment. (See Section 3.3 of this chapter for considerations regarding the
appointment of a legal guardian for such survivors). Such survivors often will
not listen to the advice of family members. The family members may in turn
seek the help of a treating health-care provider. Although the health-care
provider cannot give legal advice, a treating health-care provider is in the best
position to explain to the survivor the deficits the survivor has and how they
may be affecting his or her judgment. In my opinion, there is nothing wrong
with the health-care provider's going one step further and urging that the
survivor consult with an attorney for legal advice about his or her immediate
legal situation.
3.1 What Is a Legal Guardianship?

Although the laws of each state vary, the following discussion should apply in
a general way to the law of most states.
A guardianship is a special relationship in which the court appoints a guardian to protect the legal rights and interests of a person who is not competent to
understand or manage his or her own affairs. Different states have different
names for this relationship. In most states, it is referred to as a "guardianship"

15. Survivors and Their Families

387

or a "limited guardianship." However, some states refer to it as a "conservatorship" or a "surrogate." It will be referred to here as a "guardianship."
A guardian may be charged by the court to take care of the person of the
brain-injured patient (guardian of the person) or the property, business, or other
financial r(ehts of the patient (guardian of the estate), or both. Parents are not
automatically the "legal guardians" of their adult children. Court action is
required to make a parent the legal guardian of a child.
Someone who is in a coma or who is clearly incompetent will require a
guardian of both person and estate. The duties of such a guardian would
include arranging for the transportation, housing, and medical treatment of
the incompetent person, as well as managing that person's property and assets.
Generally, the guardian may use these assets to provide for the incompetent
survivor's everyday living expenses and medical care. However, most transactions, including sale or lease of the incompetent survivor's property, require
court approval.
In addition, the guardian may bring any legal action, such as a personalinjury suit, on behalf of the incompetent survivor. If there are legal claims
against the survivor, the guardian is required to defend against such claims.
In most states, a guardian can be either a person or an entity, such as a bank,
trust company, or nonprofit corporation. Most states require the guardian to
be at or above the age of majority and of sound mind. Many states will not
allow a person who has been convicted of a serious crime to be a guardian.
Usually, the guardian is a member of the survivor's family. However, the
guardian should be emotionally able to handle the decisions that will need to
be made on behalf of the survivor. The guardian should have good business
sense, but he or she does not need to have any special knowledge or expertise
in the business field.
3.2 Creating a Legal Guardianship

For most people, the first step in setting up a guardianship is consulting with
an attorney. The attorney prepares a petition for the local court on behalf of
the proposed guardian. When the court receives the petition or request that a
guardian be appointed for a survivor, the court will (in some states) appoint a
disinterested person ("Guardian ad Litem") to investigate the circumstances
surrounding the request and to report back to the court as to whether the
guardianship is necessary. In essence, this disinterested third party (Guardian
ad Litem) acts as the eyes and ears of the court in determining whether the
circumstances warrant a guardianship.
In most states, the people petitioning for the creation of a guardianship are
required to pay the fees of the Guardian ad Litem for investigating the matter;
however, there may be exceptions for hardship cases. The patient and his or
her family should make a point of discussing the fees and costs of setting up a
guardianship during their first visit with the attorney they consult. Although
the attorney may not be able to determine exactly what the fees and costs of

388

IV. Professional and Legal Issues

setting up the guardianship will be, the attorney should be able to give an
accurate minimum and maximum range. In states that require an investigating
Guardian ad Litem, the most difficult cost to predetermine is how much the
Guardian ad Litem will charge. It may be a good idea to request that the court
put a limit on the amount of time that the Guardian ad Litem will spend in
investigating the case.
3.3 Will the Creation of a Legal Guardianship
Have an Adverse Effect on Rehabilitation?

Brain-injured people often retain a good deal of pride about themselves and
their capabilities, despite the fact that their functional abilities have decreased.
A finding of incompetency may be psychologically damaging to a survivor. In
this regard, it is extremely important to understand that, in most states,
guardianships can often be tailored to meet the needs of the survivor.
For example, a survivor who is capable of living independently, but who
has certain cognitive impairments that adversely affect his or her ability to
manage money, may not require a full guardianship. The court should be
petitioned to create a "limited guardianship" in which the legal guardian is
responsible only for the patient's financial transactions. The stigma of a finding
of "incompetence" would not normally be necessary for such a limited
guardianship.
Because a finding of incompetence can have an adverse effect on a braininjured patient's self-esteem and confidence-emotions so crucial to successful
rehabilitation-the family, the survivor's physician, and the attorney should
thoroughly explore the possibility of a limited guardianship.
3.4 When the Survivor Is Not Competent to Consent to Medical Treatment

Normally, with the consent of the incompetent survivor's guardian, healthcare providers may administer medical treatment that is not invasive. Necessary minor surgery is usually also permitted upon the consent of the guardian
(without further court approval). However, when the surgery or treatment is
not minor, a prudent guardian may want to request that the local court review
the situation and approve the surgery or treatment before it is administered.
This, of course, is particularly true in cases in which a survivor's treating
physician indicates that a less invasive means of treatment is a viable alternative
or in which the incompetent survivor does not want the surgery or the
treatment. Some states have statutes that require a guardian to obtain court
approval before authorizing certain types of medical treatment.
in determining whether certain treatment should be administered, courts
in many states use a method known as "substituted judgment." This means
that the court determines what the incompetent survivor would do, if the
survivor were competent. The court then makes a "substituted judgment" for
the incompetent survivor. The goal is not to do what the court believes is best,
or what most people would do under the circumstances, but rather to do what

15. Survivors and Their Families 389

this particular incompetent survivor would do ifhe or she were competent and
understood all the circumstances.
The court, therefore, sets out to consider all relevant factors that would
influence the survivor's decisions regarding medical treatment. These factors
may include the prognosis if no treatment were given, the prognosis if one
treatment were chosen over another, the risk of adverse side effects from the
proposed treatment, the severity of the treatment being proposed, the wishes
of family and friends (to the extent that those wishes would influence the
survivior), the survivor's religious or moral views regarding medical care, and
the ability of the survivor to assist with therapy after treatment.
The incompetent survivor's wishes are usually given substantial weight,
even if these wishes are made known while the survivor is incompetent. The
"substituted judgment" method of determining whether medical treatment
will be administered arose out of a line of cases beginning with In re Quinlan,
70 N.]. 10, 49, 355 A.2d 647 cert. den., 429 U.S. 922, 50 L.Ed.2d 289, 97
S. Ct. 319 (1976), which involved the discontinuation oflife-sustaining treatment
based upon what the court perceived the incompetent would want. Among
the many states that utilize the method of substituted judgment to determine
whether medical treatment will be given are Alaska, In re CD.M., 627 P.2d
607, (Alaska 1981); Tennessee (deciding whether to amputate an incompetent
women's gangrenous feet), State v. Northern, 53 S. W.2d 197 (Tenn. Ct. App.
1978); Massachusetts, Rogers v. Commissioner of Dep't. of Mental Health, 390
Mass, 489, 458 N.E.2d 308 (1983) and In re Roe, 383 Mass. 415, 421 N.E. 2d
40 (1981); Colorado, In re A. w.. 637 P.2d 366 (Colo. 1981); the state of
Washington (deciding in the case of an incompetent with a cancerous larynx
whether to surgically remove the larynx as recommended by the guardian, or
to utilize less invasive but less effective radiation treatment, which the incompetent person had expressed a preference for), In re Ingram, 102 W.2d 827, 689
P.2d 1363 (1984); and New York, In re Carson, 39 Mise. 2d 544,241 N. Y.S.2d
288 (N.Y. Sup.Ct. 1962).
In cases in which health-care providers feel that certain medical treatment is
necessary, they can usually proceed upon the consent of the incompetent
survivor's guardian. However, the more serious or invasive the treatment
being recommended is, the wiser it would be for the guardian to request an
order from the local court authorizing such treatment.
4 DOES THE SURVIVOR HAVE A PERSONAL-INJURY CASE?

4.1 The Personal-Injury Attorney As Part of the Rehabilitation Team

The personal-injury attorney often finds that rehabilitation for his or her client
may never begin unless the attorney is able to obtain benefits or compensation
that will pay for it. Most health-care plans will not pay for extended rehabilitation treatment for survivors who have sustained severe brain injury. Similarly, government benefits may not provide the breadth or extent of services

390

IV. Professional and Legal Issues

needed. If the patient has been injured through the negligence of others, it may
be up to the personal-injury attorney to obtain sufficient compensation to
begin or continue the rehabilitation process.
Health care providers and families need to know something about personalinjury law and the way a personal-injury case is handled, so that they will have
some foundation upon which to deal with the personal-injury attorney. The
remainder of this section will provide some basic information on those topics.
4.2 Genecal Types of Accident Insurance

The law of Torts (a civil law) enables a person to be compensated for injuries
and losses that were caused by a defective product, someone else's negligence,
or some other form of misconduct or failure to act. Damages are usually
divided into two categories: "special" damages that are capable of being
determined with a certain amount of exactness, such as medical expenses and
wage loss; and "general" damages, which are less precise, such as pain,
suffering, disfigurement, disability, and loss of earning capacity.
Each state has its own body oflaw regarding the extent to which these types
I
'of damages may be recovered by the injured person. In some states, a system
of so-called "no-fault" insurance has been implemented by the legislature.
This type of insurance assures the injured person of some partial economic-loss
compensation but places a prohibition-or at least a limit-on the amount of
"general" damages that can be obtained. Under "no-fault," partial compensation is usually paid to injured parties, regardless of who is at fault in the
accident. On the other hand, many states allow the injured person to seek full
compensation from the negligent party for all damages sustained. Under the
second system, if an agreement cannot be reached about the amount of compensation that is adequate, the matter is determined by a trial of the case, or
sometimes by arbitration.
"No-fault" systems are difficult to discuss generally without misleading
someone who lives in any given "no-fault" jurisdiction. Therefore, the following discussion applies only to jurisdictions in which "fault" is required
before liability for damages is imposed.
4.3 Referral to and Consultation with an Attorney

Health-care providers should have no reservations about referring a traumatically injured person (or the survivor's family) to an attorney for a consultation,
with the goal of protecting the patient's legal rights. Consider the following
factors in this regard:
1. In the case ofa severely brain-injured survivor, the survivor's family is
often so traumatized emotionally that members cannot calmly and rationally
take steps to immediately protect the survivor's legal rights.
2. Survivors who have mild to moderate impairment may not be able to
accept the fact that they are brain-injured, much less understand the legal
avenues of redress open to them.

15. Survivors and Their Families

391

3. Following most accidents, evidence is lost or discarded and witnesses'


memories fade. Prompt consultation with an attorney who can investigate the
matter is vital.
4. Statutes of limitations: As with any legal claim, personal-injury cases must
be brought within a certain period of time. This period of time varies, depending on the facts of the case and local law. A law that limits the time within
which a legal claim may be brought is usually referred to as a statute of
limitation. This means that if a lawsuit is not commenced (in accordance with
local law) before the limitation period set forth in the statute-whether one,
two, or three years or more-the court will refuse to hear the case. Again,
time is of the essence in obtaining a consultation with an attorney regarding
the merits of the survivor's case.
5. Attorney fees and costs: Most attorneys who specialize in personal-injury
cases will not charge a fee for the initial consultation with the survivor or his or
her family to discuss the case. Often the attorney can render a preliminary
opinion about the merits of the case, although further investigation may need
to be undertaken by the attorney before he or she can make a recommendation
about whether the case should be pursued.
Although attorneys are usually willing to handle such cases on an hourly fee
basis, most personal injury cases are taken on a contingent fee basis. This means
that the payment oflegal fees is contingent upon the recovery of compensation
for the client. If there is no recovery, there is no legal fee. If there is a recovery,
the attorney takes a percentage of it. Percentages vary from 25% to 50%,
depending upon the nature of the case and the chance of success. An attorney
fee of one third of the recovery in a personal-injury case is fairly standard. If
the recovery results from a settlement of the case, the client must approve of
the proposed settlement before the attorney can accept it.
The survivor will be responsible for costs incurred by the attorney in
investigating and pursuing the case, and the client should inquire at the initial
interview about the type and estimated range of costs the attorney anticipates
incurring (e.g., cost of obtaining medical records, travel expenses, consultation with experts who may assist the attorney in evaluating the case, filing
fees, court costs, etc.). Certainly the attorney can be directed not to incur costs
above a certain amount without first consulting with the client. Sometimes,
the attorney can render an initial opinion about the merits of the client's case
without incurring any costs.
The contingent fee benefits the client in a number of ways. For instance, a
client who cannot afford to pay an attorney's hourly fee on a monthly basis can
retain a competent attorney without paying any attorney's fees unless and until
a recovery is made. Further, the contingent fee arrangement provides the
attorney with a financial incentive to take on only those cases that appear
meritorious and likely to result in a recovery for the client. A percentage of
nothing is still nothing, and for this reason, an attorney is not likely to involve himself or herself and the client in a personal-injury case that does not
reasonably appear to have the potential for success.

392

IV. Professional and Legal Issues

4.4 Involvement of the Health-Care Provider in the Personal-Injury Case

There are as many ways of handling a personal-injury case as there are personal
injury cases. However, there are two basic approaches that most personalinjury attorneys choose, depending upon the circumstances:
1. Many attorneys prefer to monitor the client's medical condition while
investigating the facts of the injury-causing accident. Once the client's medical
condition has stabilized to the extent that the client is at least unlikely to get
any worse, the attorney will prepare a thorough presentation of the client's
case for the opposing insurance company and will attempt to negotiate a
settlement. If an agreeable settlement cannot be reached, a lawsuit is then
commenced.
2. The other approach involves filing a lawsuit immediately. This may be
done for any number of reasons, including an impending statute oflimitations
deadline, the need for formal legal process to conduct an adequate investigation, or simply the attorney's opinion that the opposing party or Insurance
carrier will not settle the claim in good faith.
In either approach, the attorney usually relies on the client to keep him or
her informed about medical progress. However, in brain-injury cases, relying
on the client can prove frustrating for the attorney. It is not uncommon for
survivors to forget many of their symptoms when visiting the doctor. Attorneys frequently urge their clients to keep a diary of their symptoms so that
they can be thorough in explaining their difficulties to health-care providers.
However, health-care providers must be willing to spend the extra time that
brain-injured patients require in obtaining a thorough history that will permit
accurate diagnosis. Because of their respective roles of healing and obtaining compensation, the health-care provider and attorney should periodically
"compare notes" in order to better understand their patient/client.
If the attorney attempts to negotiate a settlement prior to filing a lawsuit, the
health-care provider may have contact with the attorney on one or two
occasions, usually resulting in a medical report about the survivor's history,
diagnosis, prognosis, and recommended treatment. Prior to requesting this
report, the attorney will frequently request copies of the pertinent office notes,
records, and billing.
The provider may certainly charge a reasonable clerical fee for providing
records or billing and charge for the time spent consulting with the attorney
and preparing any requested report. However, it should be kept in mind that
the client will ultimately be responsible for paying for these "costs" of
pursuing the case, and the charges should be kept at a reasonable rate. A
suggested reasonable rate for consultation with the attorney or for reportwriting would be the rate that the health-care provider would normally charge
for the same amount of time spent in treating the patient. If, for instance, the
occupational therapist charges $50 per hour for treatment, it would be rea-

15. Survivors and Their Families

393

sonable for him or her to charge the attorney the same rate for an hour spent in
consultation.
There is nothing wrong with billing attorneys in advance for what they
request. Unfortunately, there appears to be a small minority of attorneys who
do not pay their bills. Charging attorneys in advance for services rendered
prevents this small group of attorneys from spreading their guilt by association to all attorneys.
Once a lawsuit is initiated, health care providers may be involved in three
ways: ongoing consultation with the patient's attorney, deposition, and trial
testimony.
During the course of "discovery," in which each party to the lawsuit
attempts to discover the facts known by the other party, two main discovery
tools are used. The first tool is known as "interrogatories." These are written
questions that are sent by one party to another, requiring that sworn answers
be provided in writing. The other discovery tool is the "deposition."
Usually, the health-care provider will not be involved in answering any
written interrogatories. However, the attorney may want to consult with the
health-care provider about answers that the attorney or the patient will provide to certain interrogatories about medical treatment.
The most common form of pretrial contact that the health-care provider will
have regarding the litigation will be the taking of the health-care provider's
deposition. A deposition is a party's opportunity prior to trial to question
(under oath) people who may have knowledge about the facts or damages
involved in the case. In a personal-injury case, the attorney for the party
alleged to have negligently caused the injuries will want to take the depositions
of key health-care providers in order to understand the survivor's injuries.
Such depositions normally take place in the provider's office, with the survivor's attorney present. A court reporter will also be present to record what is
said, and, if requested, to type a transcript of the questions and answers. The
"deponent" (person being asked the questions) will have an opportunity to
read over the transcript. Usually the deponent-witness should accept this
opportunity.
The health-care provider's notice that his or her deposition is desired may
occur when the provider receives a subpoena requiring attendance and production of pertinent medical records at the deposition. If the date given is
inconvenient, the health-care provider can certainly request, directly or through
the patient's attorney, that a more convenient time be set. Normally, these
requests will be complied with.
It is prudent for the survivor's attorney to consult with the health-care
provider before the provider's deposition. The survivor's attorney may be
charged for this time. However, the time spent during the actual deposition
should be charged to the attorney who has requested the taking of the
deposition.
If a settlement is not reached during the course of discovery, the case will be

394

IV. Professional and Legal Issues

resolved by trial. Particularly in large urban areas, scheduling a 4ealth care


provider's trial testimony can be extremely difficult for the attorney. Typically, he or she will appear in court with the client on the day appointed for trial,
only to discover that no judge is available to hear the case. Sometimes the
attorney and client will have to wait for several days before being assigned a
judge. This, of course, can play havoc with scheduled testimony based on
the original trial date. Frequently, the attorney can determine prior to the
scheduled trial date how the court calendar is progressing and can reschedule
the health-care provider's testimony accordingly. Note to health-care providers:
If your patient's attorney has not contacted you shortly before the date when
you have been scheduled to testify, you should contact him or her to find out
how firm that date appears to be.
4.5 Risks of Settlement without an AttorneyEvaluation, Subrogation, and
Protection of Government Benefits

Occasionally an insurance adjuster will make a settlement offer to a survivor or


his or her family before an attorney has been consulted. The settlement offer
may be significant and therefore tempting. Not infrequently, the survivor and
his or her family may ask the health-care provider for an opinion about the
adequacy of the settlement offer. Again, an attorney who is experienced in
personal injurylinsurance law should be consulted.
4.5.1 Evaluation

The attorney should first be questioned about what fee, if any, he or she will
charge for reviewing the case to determine the adequacy of the settlement
offer. If the attorney feels that the settlement offer is not adequate and offers to
represent the survivor in an attempt to obtain a more reasonable recovery, a
fee arrangement should be reached that protects the offer that the client has
been able to obtain without the help of the attorney. For example:
A survivor is offered $50,000 by an insurance company and is told that it is a "final
offer." The attorney feels that the offer does not adequately compensate the survivor
for his injuries. The attorney agrees to represent the survivor and to take no percentage
of the first $50,000 she recovers and 50% of any amounts she recovers in excess of
$50,000 or 33 1/3% of the entire amount recovered-whichever fee would amount to
less.

Even if the insurance company is offering the full amount of the insurance
coverage that it has available (i.e., its "policy limits") to the survivor or his or
her family, there are still several compelling reasons for consulting an attorney.
4.5.2 Subrogation

Insurance companies or government agencies that have paid medical bills or


other benefits to the survivor may have what is know as "a right of subroga-

15. Survivors and Their Families

395

tion." A right of subrogation is essentially a right to be paid back from money


that may be received from the party responsible for the injuries. For instance,
if a person is injured in an automobile accident, his or her own insurance company may pay the initial medical bills. If the injured person subsequently agrees
to a settlement with the insurance company for the party responsible for the
accident, his or her own insurance company would assert its right of subrogation-that is, its right to be paid back from this settlement for the medical bills
or other damages it has paid. However, equitable principles may apply that
would legally allow the survivor to keep the entire settlement amount without
paying back the survivor's insurance company.
Premature acceptance of a settlement offer may wipe out the survivor's right
to his or her own insurance coverage. For instance, any "uninsured or underinsured motorist" coverage that a survivor has-which will pay the survivor
for damages suffered if the party responsible has no insurance coverage (uninsured) or does not have enough (underinsured)-may not apply until the
responsible party's liability insurance coverage, if any exists, has been paid.
Settling with the responsible party usually involves signing a release of all
claims against that party. Once this has been done, the injured party's "uninsured or underinsured motorist" carrier may refuse to pay any benefits
because its right to reimbursement or right of subrogation from the responsible party was wiped out when the survivor signed the release of all claims. A
"worst case" example follows:
A survivor, Mrs. Jones, is seriously injured through the negligence of Mr. Smith, the
"responsible party." Mr. Smith has $100,000 in insurance coverage. Mrs. Jones has
$200,000 in underinsured-motorist coverage. In Mrs. Jones' state, her underinsuredl
uninsured motorist coverage does not apply until the responsible party's coverage is
exhausted. Mrs. Jones' medical bills ($45,000) have been paid by her health insurance
company, which has a subrogated interest. Mr. Smith's insurance carrier offers its
policy limit of $100,000, which is accepted by Mrs. Jones, who signs a release of all
claims against Mr. Smith. Subsequently, Mrs. Jones' health insurance company claims
repayment in full of its subrogated interest of$45,000 from this $100,000, which would
leave only $55,000. Mrs. Jones' underinsured-motorist insurance company claims that
its right to reimbursement from Mr. Smith for any benefits that it may be obligated to
pay has now been wiped out by the release signed by Mrs. Jones, and it refuses to pay
any of its underinsurcd-motorist coverage.

An attorney may still be able to resolve a situation like that of Mrs. Jones,
above, in favor of the survivor, but it will be an uphill battle that could have
been avoided by obtaining competent legal advice in the beginning.
4.5.3 Protection of Government Benefits

After a settlement offer has been made by the insurance company, another
compelling reason for consulting an attorney is to determine what effect acceptance of that settlement would have on any government benefits that are

396

IV. Professional and Legal Issues

being received or that may be received in the future. As discussed previously,


certain Social Security benefits are available only to those who have a limited
amount of assets.
A survivor's Supplemental Security Income (SSI) benefits and Medicaid
benefits can be wiped out by receiving a personal-injury settlement. On the
other land, Social Security Disability Income (SSDI) benefits should not be
affected by a personal-injury settlement.
It is not possible to overemphasize the importance of consulting with an attorney who is familiar with Social Security benefits and eligibility requirements
before a personal-injury settlement is accepted. This warning applies whether
or not the patient is receiving Social Security or other state-administered benefits at the time of settlement. If it is foreseeable that the brain-injured patient
may be applying for such benefits in the future because of his or her disability,
the effect of the personal-injury settlement on those benefits must be examined
before the settlement is accepted. Because many Social Security and stateadministered benefits are dependent upon the amount of a person's assets, the
patient is disqualified from receiving those benefits once he or she or his or her
guardian receives settlement proceeds in excess of the asset limit set by the Social
Security Administration.
In the "worst case" example given previously, Mrs. Jones may be left with
only $55,000 because she did not consult with an attorney before accepting a
settlement. In addition, if she were receiving SSI payments, the $55,000 would
probably be enough to disqualify her from those benefits. In short, the settlement, which initially appeared to be a "good deal," has had a tremendously
adverse effect for the survivor.
Once the survivor or his or her guardian has received the settlement proceeds, the proceeds cannot be given away in order to requalify for government
benefits. At present, the Social Security Administration requires two years to
elapse before the survivor would be able to requalify for benefits.
Currently, the Social Security Administration does authorize a way in which
personal-injury settlement proceeds may be received without causing the disqualification of the survivor for Social Security benefits. (See the Social Security
Administration-SSI Programs Operation Manual (POMS), Section 01120.105.)
This involves the creation of a trust.
In a trust, one party (the trustee) holds money or property for the benefit of
another party (the beneficiary). The trustee is given the legal right over such
assets but must invest or manage the assets prudently for the benefit of the
beneficiary.
Once an acceptable personal-injury settlement offer has been made, but
before the offer is accepted, the survivor and his or her family should explore
with their attorney the pros and cons of setting up a settlement trust that would
be acceptable to the Social Security Administration so that the trust principal
and income would not be considered assets or resources available to the patient
that would disqualify him or her from receiving Social Security benefits. If the
decision is made to create a trust, a guardianship (or limited guardianship)

15. Survivors and Their Families

397

should be set up at that point-if one has not already been created-so that
the local court can be petitioned to review the proposed settlement and trust
documents and enter a court order approving both.
The party or insurance company making the personal-injury settlement offer
to the survivor must then be directed to pay the settlement proceeds directly to
the trustee under the settlement trust. The settlement proceeds should not in
any way be transferred through the head-injured patient or his or her legal
guardian or attorney. If court review and approval is obtained, the funds will
usually be paid into the registry of the court, which will then disburse trust
proceeds and attorney's fees and costs directly. The Social Security regulations
cited above provide for the creation of such a trust, and the attorney must take
care to insure that the trust document meets Social Security standards.
Note to health-care providers: You may be aware that your head-injured
patient's bill is being paid in whole or in part by Social Security/Medicaid.
You may also be aware that your patient has a personal-injury claim that is
being pursued by an attorney. You should bring the information set forth in
this chapter to the attention of your patient and his or her attorney. If your
patient's attorney is already aware of this information, you have lost only a
few minutes of your time. If the attorney is not aware, and the settlement is
made without Social Security eligibility's being taken into account, your
patient may lose all of his or her Social Security (SSI-Medicaid) benefits.
5 IS THE BRAIN INJURED PERSON BEING DISCRIMINATED AGAINST?

For the head-injured population, questions of discrimination most frequently


arise in the areas of employment and education. This section cannot cover all
of the ways in which a survivor may be discriminated against. Rather, it discusses the areas in which discrimination is most likely to be found and refers
readers to the laws that are applicable to those situations. It is hoped that, after
reading this section, survivors and their families will be able to better understand the advice of counsel about the merits of any perceived discrimination
claim.
5.1 Discrimination in Employment

If survivors are able to perform their jobs, in spite of the impairments they
have suffered, they should not be terminated from their jobs or discriminated
against in any other way solely because they are brain-injured. Employers are
required to make reasonable accommodations for handicapped employees.
Federal law and the law of most states prohibits discrimination on the basis
of handicap alone. Section 504 of the Rehabilitation Act of 1973 (29 US Code,
Section 794) states in part:
No otherwise qualified handicapped individual in the United States, as defined in [29
US Code Section 760(7)], shall, solely by reason of his handicap, be excluded from the
participation in, be denied the benefits of, or be subjected to discrimination under any

398

IV. Professional and Legal Issues

program or actIvity recelvmg Federal financial assistance or under any program or


activity conducted by any Executive agency or by the United States Postal Service.

This law does apply to employment discrimination (see the United States
Supreme Court case of Consolidated Rail Corporation v. Darrone, 465 U. S. 624,
79 -L. Ed. 2d. 568, 104 S. Ct. 1248, 1984). However, at present, this federal
law applies only to programs, activities, and employers that receive federal
financial assistance. Congress is reviewing this situation now to determine the
appropriateness of this limitation.
A brain-injured person may commence a lawsuit under this law. However,
there is some doubt about the extent of monetary damages that the head-injured
person may recover. 2 In 1978, the Rehabilitation Act of 1973 was amended to
allow the award of attorney's fees to a prevailing party (other than the United
States) in a discrimination claim under that law. (A party may lose all but one
or two of its claims and still be considered the "prevailing party" for purposes
of attorney's fees on those claims that it won or substantially won on.)
The federal regulations that implement Section 504 of the Rehabilitation Act
of1973, as amended, can be found in the Code of Federal Regulations (45 CFR,
Section 84). These regulations define a handicapped person as "one who has a
physical or mental impairment which substantially limits one or more major
life activities" (45 CFR, Section 84.3 [j]). The handicapped person must
either have a prior record showing this impairment to be present or be
presently regarded as having the impairment. For purposes of this definition, a
"mental impairment" includes "any mental or psychological disorder, such as
mental retardation, organic brain syndrome, emotional or mental illness and
specific learning disabilities" (45 CFR, Section 84.3 [iJ, emphasis added).
Federal regulations, therefore, recognize organic brain syndrome as an impairment that would qualify as a handicap for purposes of federal laws prohibiting discrimination against the handicapped.
Because of the limitations upon the applicability of federal law (federal
financial assistance required), the survivor and his or her attorney should
review the applicability of state discrimination law. Most states have enacted
legislation prohibiting discrimination against the handicapped, and in many of
the states the procedural requirements and allowable damages may be more
realistic than proceeding under federal law. 3
For survivors who feel they have been discriminated against and don't know
where to begin, perhaps the best place may be the local office of the Federal
Equal Employment Opportunity Commission. Employees there will know
the name and phone number of the counterpart agency in the state govern2

For an easy-to-undcrstand explanation of this area of federal law, sec MatlUal

011

Employment Discriminatioll Law

alld Ci"il Rights ActiollS ill Federal COllrts by C. Richey (Federal Judicial Center, revised edition, 1984). Local law
libraries should have this book.
3 See the article by W.A. Harrington, "Construction and Effect of State Legislation Forbidding Job Discrimination on Account of Physical Handicap," in volume 90 of the American Law Reports, 3rd Series, page 393
(90 ALR 3rd 393). This article should be available in local law libraries.

15. Survivors and Their Families

399

ment and may be willing to provide the names of attorneys in the area who are
competent in discrimination law. Whether to pursue a claim of discrimination
through these administrative agencies or whether to seek immediate redress
through the civil court system is a matter for the survivor and his or her attorney to determine; such a decision will be based on the circumstances of the
case and the law as it exists in their jurisdiction. Pursuing a claim of discrimination through a federal or state administrative agency may limit the amount of
damages that could be obtained if the case were brought into court in a private
civil lawsuit. On the other hand, the time lost and expense incurred in pursuing a civil lawsuit may make an administrative hearing and determination of
the matter more attractive.
Few of the Federal Appellate Court decisions regarding Section 504 of the
Rehabilitation Act of 1973 involve head-injured people. This does not mean
that survivors of brain injury are not discriminated against. However, it may
indicate that many head-injured people are so severely disabled that they are
not able to be discriminated against with regard to employment. Whether the
survivor is comatose or is suffering symptoms following so-called minor head
injury (confusion, memory loss, difficulty with concentration, perceptual
problems, headache, fatigue, etc.), there is no greater handicap to employment
or education than brain injury. Handicaps such as spinal injuries, blindness,
amputation, or deafness usually do not affect cognitive functioning. People
with these handicaps can frequently find productive employment. However,
those people with even mild impairment in, say, the area of concentration and
short-term memory find that they have tremendous difficulty in performing
their work in a manner that is satisfactory to them and to their employer.
Where head injury is concerned, it may not be apparent, even to the survivor, that a handicap exists. For example, common head-injury symptoms
include irritability, impulsiveness, increased anger, and fatigue. These symptoms can turn an excellent employee into an undesirable one, from the employer's point of view. When a survivor is terminated from a job because of
these symptoms, it may not be recognized, even by the survivor at the time,
that he or she has been fired because of the handicap.
To maintain a claim of employment discrimination on the basis ofhandicap,
the handicapped person must show that he or she is "otherwise qualified" or
able to do the work required by the employer in spite of the handicap. This
requirement may be difficult for the survivor of head injury. Consider the
example of one who suffers from problems with concentration and short-term
memory loss, whether that person is a cook who has to remember orders, a
cab driver who has to remember addresses and directions, or an attorney who
listens to the facts given by clients, applies that law, and renders advice. Cognitive impairments may not permit such survivors to perform their jobs. If a
survivor is not qualified to perform his or her job, despite reasonable accommodations by the employer, it is unlikely that he or she would be successful in
a claim of discrimination based upon involuntary termination by the employer.

400

IV. Professional and Legal Issues

The key issue is whether a person's specific disabilities have been evaluated by
the employer, rather than the employer's having made a "general" decision
about what "all" people with head injuries can or cannot do.
For many, the impairments caused by mild to moderate head injury gradually
improve or even resolve during the months following trauma. A survivor and
his or her family and health-care providers should discuss the survivor's injuries and prognosis with the employer in an effort to arrange an employment
plan that would accommodate the needs of both employer and survivor, at
least on a short-term basis.
Survivors are frequently depressed by the impairments they have sustained
and the effect those deficits have on their ability to work. However, when
employers are willing to cooperate (and the law requires that reasonable accommodations be made), survivors must be encouraged and supported, by
both family members and health-care providers, to remain employed and to
formulate systems that may compensate for the deficits they may experience.
In this regard, a health-care provider can playa vital role by explaining a survivor's injury to his or her employer and explaining what the employer can do
to help.
As indicated earlier, under federal regulations, an activity, program, or
employer that receives federal financial assistance is required to make reasonable accommodations to the known physical or mental limitations of a headinjured person who is otherwise qualified to perform his or her job (45 CFR,
Section 84.12 [a]). Most states also require the employer to accommodate the
handicapped employee. Accommodations as simple as making blackboards or
notepads available so that the survivor can make lists of steps to take or tasks
to perform may enable the survivor to perform essentialjob functions. Accommodations can often overcome inconvenience or minor problems for a person
who is able to perform essential job requirements.
A significant number of people who sustain a head injury develop posttraumatic epilepsy or seizure disorders. These people mayor may not suffer
from severe cognitive impairments (or any other impairments) caused by the
brain injury. However, under federal law (and probably the law of most states),
a person suffering from posttraumatic epilepsy is "handicapped" (45 CFR,
Part 84, Appendix A, Subpart A-3).
Epileptics have been successful in asserting their right to remain employed
injobs in which they are otherwise qualified to perform the work. In a federal
case, Drennon v. Philadelphia General Hospital, 428 F. Supp. 809, E.D. Pa. (1977),
it was held that a person with epilepsy had a cause of action under the Rehabilitation Act of 1973 when she alleged that a city-owned-and-operated hospital
denied her employment as a technician in its laboratory solely because of her
epilepsy. In a state court setting, Foods, Inc. v. Iowa Civil Rights Comm., 318
N. W. 2d 162 (Iowa, 1982), it was held that discharging a cafeteria worker
because of a convulsive epileptic seizure constituted an unfair employment
practice in violation of Iowa state law. The court awarded reinstatement and

15. Survivors and Their Families

401

back pay after finding that the worker had been fired because of her epilepsy
and seizure incident and not because she was unable to do the work required.
The court further held that the employer was required to reasonably accommodate the worker's handicap by not requiring her to work with potentially
hazardous equipment but rather assigning her to her usual duties of cleaning
tables, washing dishes, and occasionally working at the serving line and cash
register. It is the employee's ability to do the essential tasks of the job, not the
fact of the seizure disorder, that is significant.
Whether the survivor has epilepsy or not, an excellent overview of legal
rights-both federal and state-can be found in The Legal Rights oj Persons
with Epilepsy, a manual published by the Epilepsy Foundation of America.
(Contact the Foundation's Legal Advocacy Department, 4351 Garden City
Dr., Landover, MD 20785; phone: 301-459-3700.)
In summary, if survivors are able to perform their jobs with reasonable
accommodations by the employers, they should not be terminated and may
have a case for discrimination if they are fired because of their handicap.
5.2 Discrimination in Education

The "Education for All Handicapped Children Act of 1975," as amended,


states as its purpose:
... to assure that all handicapped children have available to them ... a free appropriate
public education which emphasizes special education and related services designed to
meet their unique needs, to assure that the rights of handicapped children and their
parents or guardians are protected, to assist States and localities to provide for the
education of all handicapped children, and to assess and assure the effectiveness of
efforts to educate handicapped children. (See Title 20 of the US Code, Section 1400 [c].)

The federal regulations implementing this statute can be found in Volume


34 of the Code of Federal Regulations, specifically in Sections 104, 300, 302,
and 324. The Education for All Handicapped Children Act of 1975 was designed to provide funds for state efforts to create educational opportunities for
the handicapped.
At least one federal court decision has held that head-injured children are
handicapped and entitled to a free and appropriate public education under the
Education for All Handicapped Children Act. (See the case of Howard S. v.
Friendswood Independent School District, 454 F. Supp. 634 [S.D. Texas 1978].)
That case concerned a high-school student who had minimal brain damage
(from childhood meningitis rather than from blunt trauma) as well as emotional problems. His parents obtained an injunction that prevented the school
district from expelling their son and that required the school district to make a
detailed evaluation of the boy's special educational needs, develop an individual educational plan, and subsequently contract and pay for appropriate
educational services under that plan.

402

IV. Professional and Legal Issues

In addition to a claim under the Education for All Handicapped Children


Act, a handicapped child may also have an independent cause of action under
Section 504 of the Rehabilitation Act of 1973, which has been previously
discussed.
Parents who believe that their brain-injured child is not receiving an appropriate education tailored to that child's unique needs should contact the agency
in their state that deals with protection and advocacy for people with developmental disabilities. If the parents find that it is difficult to obtain the cooperation
of the school district involved, they should contact the state superintendent of
public instruction. The school district is obligated to provide the parents with
information on how to obtain free or low-cost attorney services regarding
their child's educational rights. (In this regard, see 34 CFR, Part 300, Subpart
A, 300, 506.)
The head-injury epidemic is a relatively new situation for educators. Usually
friction between parents and educators arises over the fact that the current
special-education programs are geared toward mentally retarded, emotionally
disturbed, or learning-disabled children. Programs tailored for these handicaps
usually do not fit the special needs of traumatically brain-injured students.
Parents and health care providers can help educators provide an appropriate
and meaningful education for brain-injured students. The National Head Injury
Foundation has published an excellent manual for educators that should be
obtained and read by both parents and teachers: An Educator's Manual- What
Educators Need to Know About Students With Traumatic Brain Injury. (This book,
published in 1985, is available from the National Head Injury Foundation,
Inc., 333 Turnpike Rd., Southboro, MA 01772; phone: 508-485-9950.)
Parents should take the initiative to arrange a meeting with the school official
responsible for preparing an education plan for their child and with their child's
physician and therapists so that the educator can better understand the special
needs of this particular student. It is hoped that educating the educators will
result in a program for the student survivor that is acceptable to the parentsthereby avoiding their having to resort to taking legal steps to obtain such a
program.
Currently, some health care providers and rehabilitation facilities that treat
head-injured survivors are developing specialized programs designed to ease a
head-injured student's transitions from hospital to schoo!. The National Head
Injury Foundation has a pediatric task force looking into the needs of headinjured adolescents. 4
6 DIFFICULT QUESTIONS THAT MAY ARISE DURING
THE COURSE OF BRAIN-INJURY REHABILITATION

6.1 Should the Survivor Be Driving?

Head injury can affect driving skills in many ways. Whether the impairments
are visual, neurological (seizures, loss of motor control), neuropsychological
4 For additional information. contact the National Head Injury Foundation, Inc.. 333 Turnpike Rd., Southboro,
MA 01772; phone: 508-485-9950.

15. Survivors and Their Families

403

(problems with attention, memory, spatial judgment, impulsivity, etc.), or


even due to side effects from medication prescribed, the survivor's driving
skills may fall below the level at which it is safe for him or her to operate a
motor vehicle on public roads.
Health-care providers, particularly those involved in rehabilitation programs, are concerned about public safety and also about the legal responsibility
they may have if a survivor they are treating causes a motor-vehicle accident.
A third party injured by the survivor may have a cause of action against a
health-care provider who failed to warn the survivor that he or she should not
be driving, or who told the survivor that it was all right to drive. To review
the law around the country on this topic, see the annotation by C. Sarno,
"Liability of Physician for Injury to or Death of Third Party Due to Failure to
Disclose Driving-related Impediment," located in Volume 43 of the American
Law Reports, Fourth Series, at page 153 (43 ALR 4th 153).
Reporting the survivor to the state licensing authority may violate the duty
of confidentiality owed to the patient. Even if the health-care provider is able
to clear this hurdle, many state licensing authorities are not set up to accept and
respond to information from health-care providers about their patients' driving
skills and abilities.
Most physicians and therapists are not able to exercise any control over the
survivors whom they treat. However, when there is a concern that survivors
may be driving, even though their impairments could make it unsafe for them
to do so, the health-care providers should advise the survivors not to drive and
should definitely make note of this advice in the survivors' medical charts.
Some rehabilitation programs write a letter to the head-injured survivors,
explaining why they should not drive and requesting that they obtain approval
from their treating physicians before driving. Copies of these letters are sent to
the physicians and to family members. When a family is assisting in a survivor's rehabilitation, the health-care provider may recommend to the family
that they do what they can to keep the survivor from getting behind the wheel.
In rehabilitation programs in which the clinic staff may have some control
over a survivor, the issue of driving may be discussed with the survivor and,
when appropriate, his or her family before accepting the survivor into the
program. The rehabilitation program may wish to make acceptance of the
survivor conditional upon the survivor's agreement to follow the program's
recommendations about future driving. These recommendations may include
undergoing a driving program oriented to the particular deficits noted in the
survivor's neuropsychological evaluation and-on successful completion of
such a program-being tested or retested by the state licensing authority. In
this situation, the survivor would agree to waive any right of confidentiality he
or she may have with regard to the rehabilitation program's furnishing the
state licensing authority with information about the survivor's status.
Even in formal rehabilitation programs, however, the problem of controlling
survivors remains. It is doubtful that a health-care provider or rehabilitation
program that has taken reasonable steps to caution a survivor against driving

404

IV. Professional and Legal Issues

would be liable for damages sustained by a third party, should the survivor
take it upon himself or herself to engage in driving activities that caused those
InJunes.
If the health-care provider and survivor communicate on an ongoing basis
about the survivor's return to driving status, the survivor may well be prevented from operating a motor vehicle prematurely.
6.2 What Can Be Done to Prevent
the Survivor From Abusing Drugs or Alcohol?

Drug or alcohol abuse usually does not become a legal problem until a survivor is caught breaking the law regarding these substances. Fortunately, most
survivors have no substance-abuse problems. However, substance abuse is a
significant problem for some survivors. It must be accepted that some of these
people will never stop abusing drugs or alcohol, no matter what preventive
steps are taken. On the other hand, others will benefit from preventive measures, adopted by families and health-care providers, that are calculated to
remove or minimize the opportunity for such abuse.
The following recommendations are derived from my experience with headinjured clients and their families, as well as from an excellent article entitled
"Chemical Abuse and Head Injury."s
It is certainly easier said than done, but filling up a survivor's spare time
with stimulating activities and relationships is usually the most successful way
to avoid substance abuse. As the authors of the article point out, "If there are
no voids, there will usually be no attempts to fill them with chemicals."
After a head injury, survivors may lose friends, even old friends; this is an
unfortunate but common occurrence. Social contacts and opportunities are
thereby reduced, and the job of filling up the survivor's spare time becomes
even more difficult. However, the family and survivor should work together
to find community services and other activities (e.g., community colleges,
YMCA-YWCA, and recreational services) that are adapted for the handicapped.
These will help to provide the stimulation and feeling of self-worth that the
survivor must have to avoid the temptation of drugs and alcohol.
An excellent starting point is the local chapter of the National Head Injury
Foundation. There is a chapter or affiliate ofNHIF in almost every state. These
local chapters in turn, usually have area support groups; the survivors and their
families can attend meetings and discuss their problems and concerns with
others who may have found solutions to the same problems. Depending upon
the level of organization in the local NHIF chapter or area support group, survivors and their families may be able to become directly involved in activities
and recreational pursuits that are organized by the chapter or group. This initial
contact may provide the foundation for survivors to begin interacting with
This article. by J. Falconer, Ph.D., and E. Tercilla, Ph.D., is available from Rehabilitation Psychology
Associates, 7140 Southwest Fir Loop, Suite 130, Tigard, OR 97223.

15. Survivors and Their Families

405

others and to begin reaching back out into the world towards the ultimate
goals of productivity and independent living.
Aside from assisting survivors in finding relationships and activities that will
keep them stimulated, families should do what they can to reduce the temptation of substance abuse. These steps may include limiting the amount of cash
available to a survivor (so that it would be difficult to buy drugs), removing
alcohol from the house, or even keeping the survivor away any preinjury friends
who are likely to abuse alcohol or drugs.
Another substance abuse problem is the misuse or abuse of prescribed
medication. Each survivor usually receives a number of different prescriptions,
from a number of different physicians, to combat a wide range of medical
complications. The survivor may not wish to take some of these medications
(such as seizure medication) because of the side effects they produce. On the
other hand, the combined effect of the medications being ingested may produce side effects that are not appreciated by the individual physicians who have
prescribed the medications separately. The dangers presented by these situations can be avoided if one physician is selected to oversee and approve all
medications prescribed for the survivor and if one responsible family member
is selected to "count pills" and (through careful monitoring) to make sure that
the survivor is taking all necessary medications.
The foregoing procedures require a great deal of work on the part of families
and may be especially difficult when precautionary steps against substance
abuse require families to keep survivors away from friends or favorite activities
that are likely to cause temptation. Health-care providers must support families
in taking these steps and must reinforce survivors through regular reminders
about the adverse effects that alcohol and chemicals can have, especially for
those with brain damage.
6.3 What Can Be Done When the Survivor
Becomes Involved in the Criminal Justice System?

Survivors, particularly those who have frontal lobe damage, often suffer from
impaired judgment or the inability to control their emotions and impulses.
I was once called by the distraught mother of a head-injured young man
who had been unable to obtain government benefits that would pay for therapy
and vocational rehabilitation. After robbing a gas station, the young man got
into his car and waited for the police to arrive and arrest him. He called his
mother from the police station and told her that he was sure that the government would now pay for the rehabilitation he required. Instead, he was initially
sent to the state mental hospital.
Stories like this are as sad as they are bizarre. Handicapped or impaired
people who commit crimes that they probably would not have committed if
the handicap or impairment were absent are generally referred to as "naive
offenders. "
Sometimes, for example, the crimes are sexual in nature. Whether the sur-

406

IV. Professional and Legal Issues

vivors are exposing themselves in public or molesting others, it is usually clear


that the people the survivors were-before the head injury-would not have
done such things.
In such cases, the first step for the survivor's family to take is to set up a
meeting with the survivor's attorney or public defender. Medical records and
reports should be provided to the attorney, along with a thorough explanation of the survivor's impairments and behavioral problems. An appointment
with key health-care providers should then be arranged to further educate the
attorney about the survivor's problems.
Armed with this information, the survivor's attorney will normally arrange
a meeting with the prosecuting attorney (and perhaps subsequently with the
judge assigned to the case) to explore both a deferment of prosecution and
a court-monitored plan of treatment that will, it is hoped, be of help to the survivor. This approach requires certain key elements that are not always present:
an aggressive defense counsel, a cooperative prosecutor, and an understanding
judge, not to mention the diligence of family members and the assistance of
health-care providers.
With some luck, the survivor may be able to avoid a conviction, or harsh
sentence, or both, if steps are taken that assure the court that the criminal conduct will not continue. Unfortunately, the same head-injury symptoms that
permitted the first crime to occur remain, and the family must redouble its
efforts in eliminating the cause or stimulus that may produce the criminal behavior. In second-offense or third-offense situations, judges and prosecuting
attorneys are much less understanding.
Following the crime, the young man who robbed the gas station to obtain
treatment spent several unproductive months in the state mental hospital.
However, the court and prosecutor were fully informed about his handicap;
luckily, they were sympathetic. Prosecution was deferred for several years,
with the understanding that the charges would be dropped if no further criminal activity occurred. In the meantime, a more liberal approach toward headinjured individuals by the Social Security Administration allowed the young
man to qualify for SSI and vocational rehabilitation. As of this writing, he has
not engaged in further criminal activity.
7 CONCLUDING NOTE

This chapter has outlined, in the order in which they usually arise, legal questions that commonly confront head-injury survivors and their families.
Family support is an important factor in head-injury rehabilitation. Of
course, many survivors will not improve substantially, even with strong family
support. On the other hand, many others could improve, but will not, because
family support or the support of others is lacking. To a certain extent, healthcare providers, guardians, case managers, and friends can help take up some of
this slack. Usually, however, the survivor returns to his or her family, and the
family has the opportunity to be the ultimate care-giver.

15. Survivors and Their Families

407

Family members, or those who have taken on the duties of supporting the
survivor, must dedicate themselves to be advocates for the survivor. This
applies whether the family is seeking rehabilitative care or legal assistance. A
survivor and those who support him or her should be careful in the selection of
an attorney, should require periodic status reports from the attorney selected
about the progress of the case, and should be as cooperative as possible in
supporting the attorney's efforts on their behalf.

INDEX

Abstract attitude, 277


Abstract reasoning, deficits in, 5;
rehabilitation of, 9
Abstract thinking, 2
Abuse, aggression and, 75; of alcohol and
drugs, 404-405
Academic-related activities, 225; example
of,235-236
Acceleration-deceleration movements, 18
Acceptance, of disability, 2, 6; emotional, of
brain injury, 197; of situation, 292
Accident, motorcycle, 16; motor vehicle, 2,
16
Accident insurance; see Insurance
Accountability, 375
ACTH (adrenocorticotropic hormone), 98
Acting out, 54
Activities of daily living (ADL), 214-215,
224-225,266; deficits in, 5;
rehabilitation of, 9; skills, 48, 195
Activity pattern, 232
Activity Pattern Analysis (APA), 225
Activity Pattern Indicators, 225
Activity, voluntary, 129
Acute phase of posttraumatic medical care,
380-381
Acute rehabilitation, 188; hospitals, 183
Adaptive behaviors, promotion of, 348-351
ADH; see Antidiuretic hormone

Adjustment, 79
Admission, conference, 175; criteria for,
376; to day program, 223-224; for
inpatient rehabilitation, 159; to
residential treatment, 193
Adolescents, younger, 2
Advocacy, by family, 313-314
Affect, 250-251
Afferent field, 148
Age, at which injury occurs, 5; relationship
to prognosis, 24; of survivor, 9
Aggression, 75-78, 166, 196; betaadrenergic blockers and, 94; conscious,
54; context of, 76; definition of, 320;
effects of medication on, 332; epilepsy
and, 54; integration of diagnostic
findings, 336-337; management of, 76;
neuropsychiatric diagnosis of, 332-335;
neuropsychological diagnosis of,
335-336; psychopharmacology and,
76-78; postictal, 76; postinjury factors,
330-332,333-335; posttraumatic
epilepsy and, 330; preinjury factors,
330,332-333; seizures and, 330; verbal
and physical, following brain injury, 249
(see also Aggressive behavior)
Aggressive behavior, antecedents of,
323-324; assessment of, 325-337;
baseline measurement of, 326"':329;
409

410

Index

consequences of, 324; contingency


management of, 324; decreasing, using
social learning principles and
techniques, 339-347; goal-directed
nature of, 320-321; management of,
during assessment phase, 337-339;
management of, following TBI,
317 - 360; neurophysiologic diagnosis of,
329-332; sample menu of, 325 (table);
social learning perspective and, 323325; social-behavioral diagnosis of,
325-329; stimulus controls associated
with, 321-323; target behaviors, 325326 (see also Aggression)
Agitation, 166; akathisia vs., 166
Agraphia, 68-69
Akathisia, 165-166; agitation vs., 166;
treatment of, 166
Akinesia, 62
Akinetic mutism, 56,161,163
Alcohol abuse, survivors and, 404-405
Alexia, 42, 69
"All or nothing" law, 130
Alpha-adrenergic systems, antidepressants
and,93
Alzheimer's dementia, 95
American Academy of Physical Medicine
and Rehabilitation, 159
American Congress of Rehabilitation
Medicine, 188
American Medical Association, Judicial
Council, 165
Amitriptyline, use with pathological
laughing or crying, 73
Amnesia, 48; anterograde, 71-72;
benzodiazepines and, 94; posttraumatic
(PTA), 69-72; reduplicative para-, 72;
retrograde, 70- 71
Analyzers, auditory, 132, 134, 140;
kinesthetic, 132, 134, 140; motor
analysis and synthesis, 132, 134, 140;
optic, 132, 134, 140
Anamnesis, 132
Anatomical reorganization theory, 6
Anoetic syndrome, 161
Anomia, category-specific, 68; modalityspecific, 68; semantic, 68; types of,
67-68; word-selection, 68
Anosmia, 29
Anosognosia, 78
Anoxia, 2, 47
Anterior choroidal artery, deformation of,
49
Anticholinergic profiles, of antidepressants,
93
Anticholinergics, akathisia and, 166
Anticoagulation treatment, 33
Anticonvulsants, 91; behavioral

complications of, 92; prophylaxis with,


52; side effects of, 53
Antidepressants, aggressive behavior and,
92; agitation and, 92; alpha-adrenergic
systems and, 93; anticholinergic profiles
of, 93; beta-adrenergic systems and, 93;
children and, 92-93; dopaminergic
systems and, 93; histaminic systems and,
93; neurotransmitter systems and, 93;
posttraumatic sleep disturbances and,
92; posttraumatic stress disorders and,
92; seizures and, 55; serotonergic
mechanisms and, 93
Antidiuretic hormone, inappropriate
secretion of, 37; lack of, 37
Anton's syndrome, 49
Anxiety, 245; following brain injury, 249
Anxiolytic effects of benzodiazepines, 93
APA interview, 225-226
Apallic syndrome, 56, 161
Apartment life, 189; programs, 190
Apathy, 62
Aphasia, 16,64; anomic, 64; global, 66;
nonfluent, 64; subcortical, 66;
transcortical, 64; transcortical motor,
67; transcortical sensory, 64-65;
Wernicke's, 64
Aphasias, 65 (table)
Aphasic syndrome, subcortical, 63
Aphonic, 131
Apnea, 17
Apraxia, 47; intentional, 130; kinesthetic,
130; kinetic, 130; spatial, 130
Aprosodia, 73- 74
Arterial-venous oxygen difference
(AVD0 2 ),22
Artifact theory, 6
Assessment, of family resources, 307 -308;
of personality, 10, 170; for residential
treatment, 193
Asthenia, 131
Atrophy, cerebral, difficulty of
distinguishing from hydrocephalus, 37
Attention, difficulties with, 2; impairments
of, 61; rehabilitation of, 9; variable, 54
Attention/concentration, deficits in, 5
Attitudinal issues, behavioral management
and,355
Attorneys, consultation with, 390-391; fees
and costs, 391; as part of rehabilitation
team, 389-390; personal-injury case
and, 390-397; referral to, 390-391
Australia, incidence of head injury in, 5
Auto insurance; see Insurance
AVD0 2 ; see Arterial-venous oxygen
difference
Aversive stimulus, removal of, 344
Awareness of deficits, 292

Index

Axonal injury, diffuse, 43


Axons, disruption of, 18
Barbiturate sedation, 24
Barthel score, 57
Basal cisterns, 22
Basal ganglia; see Ganglia, basal
Baseline measurement of aggressive
behaviors, 326-329
Basofrontal area, contusions of, 33
Behavior, disturbances of, 91; evaluation of,
170; neurological interventions, 6;
psychopathological, 123; selfdestructive, 168; as sequence of events,
112; therapy, 6
Behavior analysis, 309-313
Behavior-modification strategies, 170
Behavioral changes after brain injury,
243-244,248-249,252-253
Behavioral dyscontrol, 166
Behavioral interview, 329
Behavioral-management strategies, 170 (see
also Aggression, Aggressive behavior)
Behavioral neurology, 39; in head injury, 42
Behavioral problem, definition of, 167
Behavioral rehabilitation programs, 188
Behavioral taxonomy, 306
Behavioral units, 129
Benefits, government; see Government
benefits
Benzodiazepines, 77; adverse effects of, 94;
akathisia and, 166; anxiolytic effects of,
93; behavioral complications of, 92;
sedative effects of, 93; seizure activity
and, 93; tremor and, 93
Beta-adrenergic antagonists; see Betaadrenergic blockers
Beta-adrenergic blockers, 77; aggression
and, 94; akathisia and, 166;
contraindications, 94-95; dementia
and, 94; Korsakoff's syndrome and, 94;
side effects of, 95
Beta-adrenergic systems, antidepressants
and, 93
Beta-blockers; see Beta-adrenergic blockers
Birth control, 197
Bleeding, extradural, 29
Blindness, 29
Blood pressure, rise in, as sign of increasing
ICP,31
Bourdenko Neurosurgical Hospital, 137
Boxing, 17
Bradycardia, as sign of increasing ICP, 31
Brain, structures of, 129
Brain death, committee to define, 165
Brain injury, aggression and, 75; behavioral
neurology and, 39-90; behavioral
sequelae of, 243-244, 248-249,

411

252-253; day programming and,


221-237; definition of, 8; epidemiology
of, 157; incident of, 157, 183; legal
issues and, 379-407; long-term family
intervention and, 297-316;
neuropsychological investigation and,
127-153; neuropsychotherapy and,
241-269; pathophysiological basis for,
15-38; professional issues and, 363377; psychiatric perspective, 105-125;
psychopharmacological agents in
treatment of, 91-104; residential
treatment and, 183-219; structured
group treatment and, 271-295 (see also
Traumatic Brain Injury)
Brain Injury Special Interest Group, 159
Brainstem, involvement of, 48; reflexes, 54
Bromocriptine, 59, 163
Bryn Mawr Coma Emergence Scale, 56, 57
(table)
Burr-hole, for monitoring ICP, 28; temporal,
for decompression, 31
Calcium channel blockers, 95
Carbamazepine, 53, 77-78; behavioral
complications of, 92
Cardiopulmonary resuscitation (CPR), staff
training and, 209
CARF; see Commission on Accreditation of
Rehabilitation Facilities
Case management, 364-365
Catastrophic panic, 245
Catecholamines,98
Cell death, glial, 25
Central nervous system (CNS), 91
Cerebellar Purkinje cells, 47
Cerebral blood flow (CBF) studies, 166
Cerebral swelling, diffuse, 45
Cerebral vascular disorders, 2
Cerebral vasomotor reactivity, disturbed, 18
Certification for professionals, 376
Change, theory of, 185
Children, brain-injured, 169; stimulants and,
99; younger adolescents and, 2
Chlorpromazine, 77
Cholinergic agents, 95
Choosing appropriate program, 189
Circumstantiality, 54
Clinical neurorehabilitation; see
Neurorehabilitation, clinical
Clinician, experimental, 152
Clonus, 54
Closed-brain injury, 2
Coagulopathy,33
Cognitive evaluation and remediation, 170
Cognitive functioning, 250
Cognitive functions, relationship to
interpersonal functions, 273-276

412

Index

Cognitive organization, deficits in, 249


Cognitive psychology, 10
Cognitive remediation, 3, 6, 171-172, 195
Cognitive therapy, 375
Coma, 162; arousal programs, 163; definition
of, 56,161; emergence, 165; level of, 22;
relationship to prognosis, 24; vigil, 161
Coma-emerging stage, 157
Comatose, 2
Commission on Accreditation of
Rehabilitation Facilities (CARF), 160,
169
Communication, in neuropsychotherapy,
259-260
Community mobility, 225; example of, 229
Community reintegration, 214-215
Competence, fUlilctional, 292
Competence of services, 374-375
Competency of survivor, 386- 389
Comprehensive care plan, 173-174
Comprehensive outpatient rehabilitation
evaluation (CORE), 226
Computerized tomography; see CT
Concentration, difficulties, 2; rehabilitation
of, 9
Concept formation, rehabilitation of, 9
Conclusion, formulation of clinical
neuropsychological,142
Concussion; see Diffuse injury
Confabulation, 72, 258
Conference, admission, 175; discharge
planning, 176; team, 176
Conflict, 247; staff, 370
Confusional state, 58
Congestion, vascular, 18; venous, 18
Congestive heart failure, beta-adrenergic
blockers and, 94
Consciousness, 129; loss of (LOC), 17; and
PTA,70
Content, in neuropsychotherapy, 260-261
Context, definition of, 112
Contingent fee, 391
Contre-coup effect, 25
Contusions, anterior temporal, 47;
brainstem, 43; definition of, 25; focal
parenchymal, 47; frontopolar, 47;
frontotemporal, 45; hypothalamic, 37;
ICP elevations and, 24; orbitofrontal,
47; in Sylvian fissure, 25; in temporal
lobe, 33; in temporal poles, 25
Conversation, preliminary, 138; with
patient, 134
Cooperation of patient, 137
Corpus callosum, 22
Cost-benefit analyses, 184
Cost considerations, 160
Cost-containment, cognitive remediation
and,171
Costs, health-care, increase in, 184,364

Counseling, family, 170-171; residential


treatment, 196; vocational/academic,
172
Countertransference, 265-266
Course of treatment, residential treatment,
193
Cranium, 25
Crime victims, government benefits for, 385
Crimes, violent, 2
Criminal justice system, survivors and,
405-406
Crying, pathological, 73
CT, 1, 134; to identify DAI, 43; scans, 17,
169
Cushing response, 31
Cytoarchitectonic studies, 130
Data collection, behavior analysis and,
310-311
Day hospital, vs. day program, 221
Day programs, 221-237; advantages of, 222;
vs. day hospital, 221; evaluation of
patients, 224; goals of, 221; "home
base" for, 221; implementation of,
230-232; treatment planning, 224; when
to enter, 223-224
Deafness, word, 42
Deblocking, 147, 150
Defense, 247
Deinhibition, 147, 150
Deja vu, 53
Delirium, 58
Dementia, 15; beta-adrenergic blockers and,
94; posttraumatic, 37; as symptom of
hydrocephalus, 49
Dendrites, disruption of, 18
Denial, 78-79,113; as a defense mechanism,
113; in family, 164; impaired cognitive
function and, 113
Denmark, incidence of head injury in, 4,
16-17; rehabilitation programs in, 3
Department of Health, Education and
Welfare, 382
Depression, 74-75; carbamazepine and, 78;
diagnosis of, 75; epilepsy and, 54-55;
following brain injury, 249
Developmental epochs, 110
Developmental experiences, aggression and,
75
Developmental neuropsychology; see
Neuropsychology, developmental
Diabetes insipidus, 37
Diabetes mellitus, beta-adrenergic blockers
and, 94
Diagnosis-related groupings (DRGs), 364
Diagnostic qualification of defect, 151
Diazepam, 77
Differential diagnosi" 368-370
Differential reinforcement of alternative

Index

behaviors (DRA), 340


Differential reinforcement of low rates
(DRL), 340
Differential reinforcement of other
behaviors (DRO), 339-340
Diffuse axonal injury (DAI), 17,43
Diffuse cerebral swelling, 45
Diffuse injury; see Diffuse axonal injury
Diplophonia,64
Diplopia, 29
Disability, brain-injury, 157
Disability income, 364
Disability Rating Scale, 8
"Disabled," definition of, 382
Discharge, criteria for, 376; from residential
treatment, 197; planning conference,
176
Disconnection from sense of self, 113
Discrimination, education and, 401-402;
employment and, 397-398
Disturbed functions, 151
Dopamine, 98
Dopamine receptor agonists, 163
Dopaminergic systems, antidepressants and,
93
Double dissociation, 143
Drive, impairments of, 62
Driver's license, regulations about, 173
Driving, assessment of ability to, 173;
disclosure of impediments to, 403-404;
liability of health-care providers, 173,
403; survivors and, 402-404
Drug abuse, survivors and, 404-405
Dura-arachnoid, opening of, 28
Dysarthria, 43, 63
Dysprosodia, 67
Echolalia, 66
Edema, 45; brain, 15; diffuse, 18;
"malignant" brain, 24
Education for All Handicapped Children
Act of 1975, 401-402
Educational level of survivor, 9
EEG, 22; abnormalities, 48; unreliability for
predicting late seizures, 52
Ego functions, 276-278; 277 (table); 278
(model)
Electroencephalogram; see EEG
Emerging coma; see Coma
Emotional disturbances, 73-75; organic
bases of, 271-272
Emotional tone, 131
Emotionality, 2
Emotions, labile, following brain injury, 249
Employer, discrimination against survivor,
397-401; workman's compensation and,
385
Employment, aids for, 234; competitive, 232
(see also Work)

413

Engaged, 292
England, incidence of brain injury in, 5;
rehabilitation programs in, 3
Environment, restrictive, need for, 200
Environment dependency syndrome, 61-62
Epidemiology, of brain injury, 157
Epilepsy, aggression and, 54, 75, 330; as
clinical diagnosis, 54; behavioral
alteration in, 54; depression and, 54-55;
focal, 55; legal rights of persons with,
401; personality alteration and, 54;
posttraumatic, 29, 37, 51; psychosis and,
54-55; temporal lobe (TLE), 54,
330-331; violent behavior and, 75-76
(see also Seizures)
Episodic dyscontrol syndrome, 76
Ethical approaches to professional conduct,
372
Ethical issues; see Issues
Evaluation, day program, 224; of personalinjury case settlement offer, 394;
residential treatment, 193
Expectation bias, of observers, 327
Expenses, medical, 364
Experiencing-self; see Self
Extinction, of behavior, 341- 342
Extrapyramidal effects, as side effects of
medication, 77
"Falling to bits," feeling of, 245
Family, 185; advantages of day program to,
222-223; assessment ofresources,
307-308; counseling, 170-171;
definition of, 301-302; educating about
long-term recovery, 305-306; impact of
TBI on, 105,201; interventions, 6; legal
issues facing, 379-407; long-term
intervention, 308-313; reactions to
TBI, 298; relations following brain
injury, 249; residential treatment and,
201-207; role in treatment, 203; role of,
302-303; support for, 206; of survivor in
vegetative state, 163-164; as system,
301-302; training, model for, 304-314
Fatigue, 131; following brain injury, 249
Fear, ictal, 53
Feedback, 128
Fistula, carotid-cavernous, 37
5-hydroxy-indoleacetic acid (5HIAA),
lithium and, 96
Focal injury, 43
Follow-up, necessity for, 236
Forced normalization, 55
"Free" interval, 31
Freud, S., 242
Frontal lobes, contusions in, 25;
parenchymal damage, 48
Frontal poles, contusions in, 25, 33
Functional adaptation, 6, 185-186

414

Index

Functional relationships, 309


Functional systems, 128
Functional units in brain, 130; first, 130;
second, 130; third, 130
Gait, difficulty, as symptom of
hydrocephalus, 49; "frontal," 49;
"magnetic," 49
Ganglia, basal, 45
Gerstmann syndrome, 67
Geschwind syndrome, 54
Glasgow Coma Scale (GCS), 19,48,161
Glasgow Outcome Scale (GOS), 8, 25
Globus pallidus, 47
Government benefits, appeal, 385; Civil
Service Annuity Program, 386; crime
victims and, 385; denial of, 385;
importance of written application for,
384; Medicaid, 382-384; protection of,
in personal injury case, 395-397;
Railroad Retirement Program, 386;
Social Security, 382-384; veteran's
benefits, 386; vocational rehabilitation
benefits, 384; workman's compensation,
385
Graduate Medical Educational National
Advisory Committee, 159
Graduated work trials, 211-214
Group, activity, 119; cluster, 118; large, 117;
in rehabilitation, 117; small, 117;
therapeutic, 118; psychotherapy, 6, 10;
ways of functioning, 117 (see also Group
Therapy, Neuropsychotherapy,
Therapy)
Group exercise, 282-295; as cognitive
hierarchy, 293; coaching, 284; example
of, 283-291; feedback in, 287; followup, 290; instructions for, 283-284; as
optimal structure, 290-291;
preparations for, 238; as rehabilitation
algorithm, 291-292; setting for, 282; as
vehicle for public disclosure and
commitment, 291
Group psychotherapy, 6, 10, 118-119;
residential treatment and, 196
Group therapies, examples of, 176 (table)
Group therapy for brain-injury survivors,
model for, 279-282
Guardian ad Litem, 387-388
Guardianship, legal, creation of, 387-388;
definition of, 386-387; effect on
survivor's relrabilitation, 388; medical
treatment and, 388-389
Habilitation vs. rehabilitation, 246
Hallucinations, 53
Haloperidol, 77, 97
Handicapped, definition of, 398

Head Injury Task Force, 188


Head injury; see Brain injury
Head Trauma Program; see New York
University Head Trauma Program
Headache, as sign of increasing ICP, 31
Health insurance; see Insurance
Health-care providers, legal issues facing,
379-407
Hedonic response, 334
Hematomas, 24; extradural, 29, 31;
intracerebral, 29; intracranial, 15, 18,
29; and seizures, 51; subdural, 29, 33;
subdural, chronic, 33
Hemianopia, 16
Hemiparesis, 16
Hemorrhages, basal ganglia, 48; focal
parenchymal, 47; petechial, 18,22
Herniation, 48; central, 48; lateral
transtentorial, 30; transtentorial, 22;
uncal (iateralized), 48
Hippocampal region lesions; see Lesions,
hippocampal
Hippocampus, 45, 47
Histaminic systems, antidepressants and, 93
Hope, selling of, 372-373
Hospitals, acute rehabilitation, 183; day
programs in, 221
Housekeeping skills of survivor 225; example
of,229
Hovedcirklen, 4
Hydrocephalus, 33; dynamic, 50; identifying,
50 (table); with increased ICP, 49;
normal pressure (NPH), 49
Hyperarousability,2
Hyperemia, 22, 45; global, 24
Hypergraphia,54
Hyperphagia, 168
Hyperreligiosity,54
Hyperventilation, 24, 30
Hypoarousability,2
Hypometabolism. 47
Hypophonia, 64
Hypopituitarism, 37
Hyposexuality,54
Hypotension, as side effect of medication,
77; systemic, 18
Hypothalamic functions, disturbances of, 37
Hypothyroidism, beta-adrenergic blockers
and,94
Hypoxia, 15; global cerebral, 17; pulmonary,
18
Hypoxic injury, 45
I CP; see Pressure, in tracranial
Ictal events, 53; fear, 53
Identity, 251-252
Identity crisis, 113
Imitation behavior, 61

Index

Impairment, behavioral management and,


352
Impairments of mental functions, 272- 273
Impatience, following brain injury, 249
Incidence of brain injury, 157
Income, disability, 364
Incontinence, urinary, as symptom of
hydrocephalus, 49
Independent living programs, 189
Individual psychotherapy, 119;
neuropsychological, 241-269
Inpatient day hospital program, 221
Inpatient programs, 158; admission criteria
for, 159; interventions, 157-181;
postacute rehabilitation programs, 191;
rehabilitation, 157; services,
coordination of, 173-179; setting, 157
Insight, 78-79
Institution, policies and procedures, 168
Insurance, accident, 381-382, 390; auto,
381-382,390; health, 381
Insurance company, long-term liability of,
373; priorities of, 373; referral from,
373;
Insurance industry, 364, 376- 377
Insurance representative, as part of planning
team, 176
Intact functions, 151
Intelligence of survivor, 9
Intentional apraxia, 130
Intentional aspect, 129
Interference, proactive, 71
Interpersonal relationships, 2
Interval, lucid, 31
Intervention, methods of, 376;
neuropsychotherapeutic, 170; quality of
life, 164
Intimacy, difficulties with, 2
Intraaxonal processes, repair of, 25
Intracranial hematoma; see Hematoma
Investigation, Luria's neuropsychological,
131-147; preliminary, 140; selective,
141
Irritability, following brain injury, 249
Ischemia, 15; late and secondary lesions due
to, 18; primary, 17
Ischemic injury, 45
Isolation, TBI and, 111
Issues, ethical, 363; legal, 379-407;
professional, 9, 363-377
Kemsley Unit, 191,353
Kernohan's notch phenomenon, 49
Kindling, 52
Kinesthetic affcrence, 129
Kinesthetic apraxia, 130
Kinetic apraxia, 130
Kinetic organization, 129

415

Kluver-Bucy syndrome, 54, 92


Korsakoff's syndrome, beta-adrenergic
blockers and, 94
Lamina cribrosa, perforations of, 29
Language difficulties, 2
Laughing, pathological, 73
Lawyers; see Attorneys
L-dopa, 98,163; use with pathological
laughing or crying, 73
Learned helplessness, 245-246; factors that
foster, 299-300; model of, 298-299
Legal issues; see Issues
Legal Rights of Persons with Epilepsy, 401
Leisure activities, 9, 216, 225; cognitive
remediation and, 172; example of, 229
Leontiev, 127
Lergotrile, 163
Lesions, description of, 15; diffuse, 15;
dorsolateral, 47; effects of, 129;
expanding intracranial mass, 30; focal,
15; hemorrhagic, 17; hippocampal, 48;
interactional,S; localized, 17 (see also
Contusions); mesial frontal, 47;
orbitofrontal, 47; parasagittal, 45;
pathophysiological serial,S;
Level of consciousness, fall in, as sign of
increasing ICP, 31
Level of functioning, qualitative analysis of,
133
Levodopa; see L-dopa
Liability, insurance company and, 377
Licensure for professionals, 376
Lithium, 77, 96; aggression and, 96; agitation
and, 96; cerebrospinal fluid and, 96;
contraindications, 96; 5-hydroxyindoleacetic acid and, 96; metabolic
activity and, 96; side effects of, 96;
tryptophan and, 96;
Lobes, frontal; see Frontal lobes
Localized lesion; see Contusion
Locked-in syndrome, 58, 162
Loneliness, following brain injury, 249
Long-term family intervention, 297-316 (see
also Family)
Long-term liability, of insurer, 373
Long-term rehabilitation, 188, funding for,
380-386
Long-term treatment, 183
Lorazepam, 94
Loss of consciousness; see Consciousness
Lucency, periventricular, on CT scans, 51
Lucid interval, 31
Luria, A. R., 127,242; theory of
investigation, 127; theory of
rehabilitation, 147 -151
Luria's Neuropsychological Investigation
(LNI), 131-147; general procedures,

416

Index

134-137; phenomenological approach


of, 137-147; stages of, 134-137;
theoretical basis for, 131-133
Magnetic resonance imaging (MRI), 1, 134,
166
Mania, carbamazepine and, 78
Mannitol, 24, 30
Market, for neuropsychological
rehabilitation, 363
Mastery of behavior, 292
Maudsley Hospital, 330
Medicaid,382-384
Medical bills, 380-381
Medical care, acute phase, 380-381
Medical expenses, 364
Medical problems, primary, 166; secondary,
166
Medical services, standards for, in inpatient
institution, 169
Medication, "atypical," 167; psychogenic
properties of, 165; psychotropic, 167
Melodic Intonation Therapy, 42
Memory, deficits in, 5, 249; organization,
134; tasks, 142; therapy, 375
Mental control, impairments of, 61
Mental status examination, 39
Mesocortical dopaminergic pathways,
damage to, 166
Mesol,iulbic dopaminergic pathways, damage
to, 166
Metabolic activity, lithium and, 96
Methylphenidate, 78,163
Metoprolol,77
Missile injuries, 47, 51
Mixed injuries, 43
"Mobile mourning," 333
Mobility, in community, 225; example of,
229; physical impairments and, 106
Model Projects for Comprehensive Services
to Individuals with Traumatic Brain
Injury, 160-161
Modeling, of staff, 354
Morbidity, prevention of, 6
Motor restlessness; see Akathisia
Motor vehicle accident; see Accident
Motorcycle accident; see Accident
MRI; see Magnetic resonance imaging
Multiple concussion syndrome, 17
Muscarinic receptors, effects of medication
on, 95
Mutism, akinetic, 49; posttraumatic, 63;
transient, 63
Myoclonus, palatal, 54
National Head Injury Foundation (NHIF),
158,185,188,313,367,380-381,402,
404

National Institute on Disability and


Rehabilitation Research (NDRR), 161
Nausea, as sign of increasing ICP, 31
Negative reinforcement, 344
Neglect, unilateral, 58
Nested-skills approach, 231
"Neurasthenia," 37
Neuroanatomy, 10; staff training and, 209
Neurobehavioral sciences, 376
Neurodevelopmental disorders, 2
Neurogenesis, 19
Neuroleptic malignant syndrome, 77
Neuroleptics, 77, 97; aggression and, 97;
agitation and, 97; contraindications, 97;
side effects, 97 - 98
Neurology, behavioral; see Behavioral
neurology
Neuropeptides,98
Neuropsychiatric functioning, staff training
and,210
Neuropsychiatric interventions, 6
Neuropsychiatric literature, 165
Neuropsychological analysis, 152
Neuropsychological investigation, 127-153
(see also Luria's Neuropsychological
Investigation)
Neuropsychological rehabilitation,
accountability, 375-376; definition of,
363; ethical issues in, 372; future of, 375;
market for, 363; professional issues in,
363-377; programs, 7; social
developments in, 364-365; specialized
services, 366; staff issues in, 363-377
Neuropsychological Rehabilitation Program,
Oklahoma City, 192
Neuropsychological remediation, 170
Neuropsychological treatment, 4
Neuropsychology, developmental, 10
Neuropsychotherapeutic intervention, 170
Neuropsychotherapy, 241-269; current
perspectives, 242-243; final phase, 267;
historical perspective, 242; individual,
10; initial phase, 254-259; middle
phase,t259-267; model of treatment,
253-267; objectives, 261; strategies,
261; techniques, 257 -259; traumatic
brain injury and, 243
Neurorehabilitation, clinical, 10
Neurotransmitter systems, antidepressants
and, 93
Neurotransmitters, 98
New York University Head Trauma
Program, 271, 293
New York University Medical Center, 271
NHIF; see National Head Injury Foundation
Norepinephrine, 98
NPH; see Hydrocephalus, normal pressure
Nursing care, 169-170

Index

Nursing staff, training of, 170


Observing-self; see Self
Occupational therapist, 115, 171
Occupational therapy, 195
Ocular brainstem signs, 43
Oculomotor nerve, damage to, 29; signs of
compression of, 31
Office of Special Education and
Rehabilitation Services (OSERS), 313
Olfactory nerves, damage to, 29
On-the-job training, 384
Open-brain injury, 2
Orbital roofs, 25
Osmotherapy, intravenous, 30
Otorrhea, 28-29
Outcome, neurobehavioral, in childhood
and adolescence, 169
Outcome research, 7, 190; methodological
considerations, 190-191; review of,
191-193
Outpatient postacute rehabilitation
programs, 192
Outpatient programs, 157,366
Overcorrection, 345-346
Oxazepam, 77
Paced auditory serial addition test, 62
Palilalia, 64
Palsy, third-nerve, 48
Paragrammatism, 66
Paraphasic substitutions, 66
Parasomnia,161
Party responsible for injury, 385-386
P ASA T; see Paced auditory serial addition
test
Pathophysiology, 91
Patients' rights, 168
Pediatric patients, 169
Peduncle compression, 31
Peer review, 377
Perinatal or postnatal complications, 2
Perseveration, 59-61, 258; in amnesia, 71;
clonic, 59; intentional, 59-60;
"recurrent," 60; "stuck-in-set," 60;
Persistent vegetative state; see Vegetative
state
Personal business and consumerism, how
survivor manages, 225; example of,
229
Personal finances, how survivor manages,
225; example of, 229
Personal-injury case, 381-382, 385,
389-397; accident insurance and, 390;
attorney as part of rehabilitation team,
389-390; consultation with, 390-391;
involvement of health-care provider in,
392-394,397; referral to, 390-391;

417

risks of settlement, 394-397


Personal-injury claim; see Personal-injury
case
Personality, alteration and epilepsy, 54;
assessment of, 10, 170; changes after
brain injury, 243-244, 272; functioning
of survivor, 9; premorbid, of survivor, 9;
theory of, 10; structure of, 246-253;
system, 249-253; working model of,
246-253
Petrosal bone, fracture of, 29
Pharmacological agents, case-controlled
studies, 167; randomized studies, 167
Phenobarbitol, 52; behavioral complications
of,92
Phenytoin, 52; behavioral complications of,
92
Physiatrist, 169, 171
Physical therapy, 115, 195
Physician referral patterns, 159
Physostigmine, 95
Pituitary functions, disturbances of, 37
Plasticity, functional, 25
Positive practice, 345-346
Positive punishment, 346-347
Positive reinforcement paradigms, 350-351
Positive reinforcers, selection of, 349-350
Positron emission testing (PET), 166
Posterior cerebral artery, 18
Postinjury psychiatric symptoms, 124
Posttrauma recovery, patterns of, 8
Posttraumatic catatonia, 161
Posttraumatic epilepsy; see Epilepsy
Posttraumatic hydrocephalus; see
Hydrocephalus
Predicative speech, 142
Preferred-provider organizations (PPOs),
364
Pre mack principle, 351
Premorbid personality, aggression and, 75
Pressure, cerebral perfusion, 18; intracranial
(ICP), 15, 18; monitoring of, 1
Primary cortical zone, 130
Priming techniques, 71
Problem Behavior Rating Scale, 328
Problem identification, 306-307
Problem selection, 306-307
Problems, common, among inpatients with
TBI, 174 (table)
Professional issues; see Issues
Professional standards, 375-376
Professional training, 366-367
Professionals, role of, 176,303-304
Prognosis, in diffuse injury, 24
Program components of residential
treatment, 187-188
Program philosophy, 370-372
Program prescription, 174, 175 (table)

418

Index

Prolonged posttraumatic unconsciousness,


161
Propranolol,77
Pro social behaviors, promotion of, 349-350
Prosody, 73
Prosopagnosia, 42
Pseudo psychopathic syndrome, 47
Pseudoseizures, 54
Psychiatric diagnosis, aggression and, 75;
disorders, impact of, 123; perspective
and TBI, 105-125
Psychological processes, qualitative analysis
of,127
Psycho neuroendocrinology , 10
Psychopathological behavior, 123
Psychopathology, 124
Psychopharmacological agents, 91
Psychopharmacology, aggression and, 76-78
Psychosensory phenomena, 53
Psychosis, "alternative," 55; carbamazepine
and, 78; epilepsy and, 54
Psychotherapy, during residential treatment,
196; individual, 243 (see also Group,
Individual, Neuropsychotherapy)
Psychotropic medication, 167
PTA; see Amnesia, posttraumatic
Ptosis, 31
Pulmonary disease, beta-adrenergic blockers
and,94
Punishment, 342-347; positive, 346-347
Pupillary response, 22
Quadriparesis, 43
Quality of daily living, example of, 235-236
Quality of life interventions, 164
Rancho Los Amigos Scale (Rancho Scale),
57,164
rCBF,134
Reaction time, deficits in, 5; rehabilitation
of,9
Reafferentation, 128
"Real world," 116-117; discharge to, 184
Recovery, 150-151; curve, 5; of function,
10; inaccuracy of term, 6; methods for
assisting, 3; need for definition of, 3, 5;
prediction of, 376; theories of, 6
Recreational programs, residential, 196
Reduplicative para-amnesia, 72
Reduplicative phenomena, 72
Referral patterns, physician, 159
Regional Spinal Cord Injury Systems, 161
Regrowth, refinement of, 25
Rehabilitation, 150-151; acute, 188; goal of,
114; long-term, 188; Luria's theory of,
147; need for definition of, 5; process,
framework for, 115; programs,
neuropsychological, 7; as series of

opportunities, 114; workshops vs. onthe-job training, 384


Rehabilitation Act of 1973,384,398-401
Reinforcement, 339-341
Relationships, interpersonal, 251;
therapeutic, 256-257
Relearning, 25
Reliability of test results, 136
Remediation of skills, 6
Reorganization, of disturbed functions, 150;
first form of, 149; of functional systems,
149; intersystemic, 148, 150;
intrasystemic, 148, 150; second form of,
149; third form of, 149
Repair of intraaxonal processes, 25
Representation of services, 374-375
Research methods, clinical and preclinical,
10
Residential care programs, 188
Residential programs, types of, 189;
traditional,190
Residential treatment, 183-219; types of
programs, 186-190
Resistance, to communication, 261-263
Respiratory care, 169-170
Response-cost procedures, 344-345
Responsibility, 121
Restoration, physical, 172
Restraint, mechanical, 168; pharmacologic,
168
Restrictive environment, need for, 200
Retraction bulbs, 18
Retraining of skills, 6
Return to work incentives, 383-384
Rhinorrhea, 28-29
"Richmond screw," for monitoring ICP, 28
Rigidity, decerebrate, 22; decorticate, 22
Rimming, ventricular, on MRI scans, 51
Roles, professional, 176
Rusk Institute of Rehabilitation Medicine,
271
St. Andrew's Hospital, 191
Scotland, incidence of head injury in, 5
Secondary cortical zone, 130
Sedation, as side effect of medication, 77;
barbiturate, 24
Sedative effects of benzodiazepines, 93
Seizures, and aggression, 330
Seizures, anticonvulsants and, 91;
benzodiazepines and, 93; control of, 52;
"early," 51; and EEG, 52; frequency of,
52; late, 51; and missile injuries, 51 (see
also Epilepsy)
Self, experiencing-, 248; observing-, 248
Self, disconnection from sense of, 113
Self-appraisal, unrealistic, following brain
injury, 249

Index

Self-care, 225; example of, 228


Self-concept, 2
Self-destructive behaviors, 168
Self-esteem, 2
Selling hope, 372-373
Sensorimotor reactions, 134
Sensory stimulation, 163; techniques, 163
Serotonergic mechanisms, antidepressants
and, 93; neurotrans'mitter system, 77
Services, specialized, 366
Severity of injury, 5
Sexuality, 196-197
Shearing, diffuse, 18
Shearing lesion; see Diffuse injury
Sheltered workshops, 214
Shunt, ventriculoperitoneal, to relieve
hydrocephalus, 37
Shunting, complications of, 51
Side effects of medications, 77
Skull fractures, depressed, and seizures, 51
Slowness, 43
Social contact, following brain injury, 249
Social learning perspective, principles of,
323-325; impediments to application in
behavioral management program,
351-356; use in behavioral management
program, 323-356
Social life of survivor, 9
Social Security Act, 382
Social Security Administration, 382
Social Security Administration-SS! Programs
Operations Manual (POMS), 396
Social Security benefits, 382-384, 396-397
(see also Government benefits)
Social Security Disability Income (SSDI),
382-384,396-397
Social skills, training, 215-216
Social workers, 115, 171
Spatial apraxia, 130
Special education staff, 171
Speech, disorders of, 63; predicative, 142
Speech and language therapy, 195
SpeechlIanguage pathologist, 115, 171
Sphenoidal wing, 25
Staff, composition of, 207; major issues, 207;
psychological conflicts of, 370; in
residential treatment facility, 207 - 211;
support for, 210-211; training, 208, 217,
355-356,370 (see also Professionals)
Staff issues, behavioral management and,
353- 356 (see also Issues)
Statutes of limitation, 391
Steroids, 24
Stimulants, 98; children and, 99; depression
and, 99
Stress, 121-122; therapy staff and, 122
Structured group treatment, 271-295
Stuttering, 64

419

Subarachnoid bolt, for monitoring Iep, 28


Subcortical aphasic syndrome, 63
Subrogation, personal-injury case and,
394-395
Substance abuse, aggression and, 75
Substitutive compensation, 147
Suicide, 75,197
Supplemental Security Income (SSI),
382-384,396-397
Support, models of, 300; of staff, 210-211
Support groups, for families of survivors, 404
Survivors, impact ofTBI on, 105; isolation
of, 111; legal issues facing, 379-407;
missed opportunities of, 105; problems
of, 105 (see also Family, Traumatic
Brain Injury)
Swelling, brain, 24
Swelling, diffuse cerebral; see Diffuse
cerebral swelling
Sylvian fissure, contusions in, 25
Syncope, 54
Syndrome analysis, 132, 136
Tardive dyskinesia, 77
Target behaviors, adaptive, 348-349;
maladaptive, 325-326
TBI; see Traumatic brain injury
Techniques, therapeutic, 257-259
Temporal lobe, contusions of, 33
Temporal lobe epilepsy; see Epilepsy
Temporal poles, contusions in, 25
Tentorial edge, 18
Tertiary cortical zone, 130
Test validation, 133
Theory of change, 185
Therapeutic activities, 120, 123;
interventions, 119
Therapeutic contract, 255-256
Therapeutic dialogue, 280
Therapeutic recreation specialist, 115, 171
Therapeutic relationship, 256-257
Therapist, role of, 116
Therapy; see Group, Individual,
Neuropsychotherapy
Thioridazine, 77
Thought disorders, 69
Time-out procedures, 343-344
Title 42; see Social Security Act
Title II, 382; see Social Security Disability
Income
Title XIX; see Medicaid
Title XVI; see Supplemental Security
Income
TLE; see Epilepsy
Torts, law of, 390
Tracheostomy, 169,209
Transference, 263-265
Transitional living programs, 189

420

Index

Transmitter synthesis, 19
Trauma services, 159
Traumatic brain injury (TBI), 2; effects on
patients and family, 297; effects on
personality, 244-245; incidence of, 4;
need for services, 364;
neuropsychotherapy and, 241- 269; staff
training and, 209 (see also Brain injury)
Traumatic head injury, definition of, 380 (see
also Traumatic Brain Injury)
Treatment, long-term, 183; planning,
residential, 183-219; specialized
individual, 6 (see also Therapy)
Tremors, 54; benzodiazepines and, 93
Triazolam,94
Tumors, of brain, 2
Uncal herniation; see Herniation
Uncertainty, 121
Unconscious, 242, 247
Unconsciousness, 28
Unilateral neglect; see Neglect
United States, incidence of brain injury in, 4,
158, 183
United States Government, 382
Units, behavioral, 129
Utilization behavior, 61
Vascular congestion; see Congestion
Vasoparalysis, global, 24
Vasopressin, 98
Vegetative state, 2,15,56-57,162;
persistent, 56,161; services for families
of survivors in, 163-164
Ventilator programs, 169

Ventral pontine syndrome, 162


Ventricles, lateral, 22
Ventricular dilatation, ex vacuo, 49
Verbal expression, disorders of, 63-69
Vestibulocochlear nerve, 28
Videotapes, patients and, 215, 282, 290; staff
and, 338
Violent behavior, EEG and, 76; seizures
and, 76
Vision impairment, ipsilateral, 29
Visuospatial afference, 129
Vocational assessment, 211
Vocational counseling, 211
Vocational outcome, 7
Vocational programs, residential, 211-214
Vocational rehabilitation benefits, 384
Vocational specialist, 115, 171
Voluntary activity, 129; movements,
disturbances of, 130
Volunteer work, 233
Vomiting, as sign of increasing rcp, 31
Vygotsky,127
Wales, incidence of head injury in, 5
Wernicke's area, 42
Work, competitive, 232; as set of behaviors,
307; of survivor, 9; trials, graduated,
211-214; volunteer, 233
Work-related activities, 225; example of, 230
Work-related injury, 385
Workman's compensation, 364, 377, 385
Workshops, sheltered, 214
Zones, cortical, 130; primary, 130;
secondary, 130; tertiary, 130

S-ar putea să vă placă și