Você está na página 1de 8

Implications of Research

in Cognitive Neuroscience for


Psychodynamic Psychotherapy

Drew Westen

(Reprinted with permission from Westen D: Implications of research in cognitive neuroscience for psychodynamic psychotherapy, in The
Oxford Textbook of Psychotherapy. Edited by Gabbard GO, Beck JS, Holmes J. Oxford, Oxford University Press, 2005, pp 443448)

Virtually all psychotherapies rely on some mixture of the following:


exposure to new or anxiety-provoking stimuli,
ideas, feelings, or behaviors;
efforts at understanding and reworking
problematic ways of thinking, feeling, and
behaving;
efforts at behavior change that may in turn
catalyze cognitive and emotional change (as
well changes in the behavior of others);
interaction with another person (or group of
people) who may provide a supportive environment, act in ways that disconfirm past
expectations about relationships, or offer new
ways of interacting.

focus.psychiatryonline.org

LEARNING, MEMORY, AND THE


EVOLUTION OF COGNITIVE
NEUROSCIENCE
This section briefly describes the evolution of
conceptions of learning, memory, and cognition of
relevance to contemporary theory and research in
cognitive neuroscience. It focuses on how earlier
research inspired contemporary approaches to
treatment. The chapter then examines implications of more recent developments for all forms of
psychotherapy.

CLASSICAL

AND OPERANT CONDITIONING

The first systematic approach to learning


emerged from the laboratories of Pavlov, Skinner,
and hundreds of other researchers who studied what
came to be known as classical and operant conditioning. For much of the first half of the twentieth
century, researchers from a behavioral tradition
argued that the most complex behaviors reflect a
handful of learning mechanisms shared by humans
and other animals that could be understood without reference to internal mental processes. The animal learns in classical conditioning to produce a
relatively automatic response when a previously
neutral stimulus (the conditioned stimulus) is
repeatedly paired with a stimulus that innately
(prior to learning) produces a similar response (a
conditioned response). The best known example
occurred in Pavlovs experiments, in which dogs
learned to salivate at the sound of a tone that tended

FOCUS

Spring 2006, Vol. IV, No. 2

INFLUENTIAL
P U B L I C AT I O N S

All of these processes rely on learning, memory,


and cognitive change, which suggests that relevant
developments in the basic sciences should be useful
in conceptualizing, reformulating, and adding to our
repertoire of psychotherapeutic interventions. The
extraordinary progress in cognitive neuroscience
(and the related, emerging field of affective neuroscience) in the last decade has as yet led to only to a
handful of studies directly relevant to psychotherapy
(e.g., research linking changes in brain to changes in
depression or anxiety responses; e.g., Brody et al.,
2001; Goldapple et al., 2004). However, basic science data generated thus far may have substantial
implications for the therapeutic practice, both by
supporting long-held clinical hypotheses about the
way neural networks function and by challenging
exclusive use of therapeutic practices that focus primarily on only a handful of systems that regulate
thought, emotion, and behavior.
This chapter begins with a brief description of
how psychologists, cognitive scientists, and cognitive neuroscientists have come to understand learning, memory, and cognition. (Although one could
profitably focus on the cellular level, given that all
learning ultimately involves changes in synaptic

connections, gene expression, etc., the focus here is


primarily on molar processes likely to translate
more directly into implications for psychotherapeutic interventions.) It then briefly describes potential
implications for psychodynamic psychotherapy.
(For an expanded presentation of some of these
ideas, see Westen, 2000a,b, 2002; Westen and
Gabbard, 2002a,b; Gabbard and Westen 2003.)

215

WESTEN

to precede presentation of meat. The animal learns


in operant conditioning to associate certain behaviors
with consequencesreinforcers and punishersthat
increase or decrease the likelihood of the behavior
recurring. In general, classical conditioning tends to
involve involuntary reactions, whereas operant conditioning involves voluntary behaviors that a person
or animal performs or inhibits to obtain or avoid
rewarding or aversive consequences.
The understanding of classical and operant conditioning led, in the 1950s and 1960s, to the development of behavior therapies aimed at altering
conditioned emotional responses and maladaptive
behaviors. Classically conditioned emotional
responses are involved in many forms of psychopathology, particularly in anxiety disorders
(e.g., in the startle responses and intense anxiety
and autonomic reactivity that occur when patients
with posttraumatic stress disorder encounter triggers that resemble in some way those present during a traumatic event). Some of the earliest
behavioral treatments emerged directly from
research on classical conditioning processes, as
researchers and clinicians developed exposure techniques to try to break associative links between
stimuli (or imagined stimuli, as in flashbacks of
traumatic events) and intense negative feelings,
particularly fear and anxiety (Wolpe, 1958).
Exposure means presenting the person with the
feared stimulus and preventing him or her from
escaping the initial feelings of anxiety or panic.
Over time, if the person cannot escape exposure,
the intense emotional reaction irrationally associated with an objectively nonthreatening stimulus
will generally wane if not extinguish entirely. Using
exposure to treat conditioned emotional responses
in anxiety disorders has been demonstrated to be
quite efficacious (Barlow, 2002).
Behavior therapists similarly learned to use principles of operant conditioning to treat a range of problems, such as maladaptive parenting strategies that
fostered rather than curtailed aggression. Most
behavioral treatments use both classical and operant
principles to promote behavior change. For example,
effective treatments of anxiety disorders tend to
address not only classically conditioned emotional
responses hut also the avoidance mechanisms
patients develop through operant conditioning to
escape frightening experiences (e.g., agoraphobic
avoidance of situations associated with panic
attacks). Thus, behavior therapists typically combine
exposure to threatening stimuli (aimed at extinguishing a classically conditioned response) with
response prevention (preventing the patient from
escaping the feared stimulus and hence extinguishing a response learned via operant conditioning).

216

Spring 2006, Vol. IV, No. 2

FOCUS

SERIAL (CONSCIOUS) PROCESSING OF


INFORMATION: THE COGNITIVE REVOLUTION
Although highly productive, the behaviorist
enterprise ultimately ran aground as the dominant
perspective in experimental psychology as
researchers increasingly recognized anomalies that
could not be understood without reference to mental processes. Buoyed by developments in artificial
intelligence (and the development of high-speed
computers), cognitive science began to displace
behaviorism in a scientific revolution that began in
the late 1950s (see Robins et al., 1999). Based on
the metaphor of the mind as a computer,
researchers developed a serial processing model of
cognitionthat is, a model in which information
passes sequentially (serially) through a series of
three memory stores (Atkinson and Shiffrin,
1968). This three-stage model, now sometimes
called the modal model (Healy and McNamara,
1996), provided the theoretical basis for cognitive
research for 30 years.
According to this model, following a brief initial
stage of sensory registration that retains information for a fraction of a second, information is held
in short-term memory, which can maintain
roughly seven pieces of information in consciousness for about 30 seconds (Miller, 1956). (The
move to 10-digit local phone numbers in the US in
the 1990s, necessitated by the proliferation of fax
and computer lines, has posed a challenge to the
limits of human short-term memory.) The next
stage is long-term memory, from which information, if properly processed (e.g., memorized in a
way that is meaningful), is retrieved as needed into
short-term memory. Although information may
remain in long-term memory indefinitely, in general, the more frequently and recently information
has been used, the easier it is to retrieve.
Researchers offered a number of theories and
metaphors to describe the way information is
stored in long-term memory. One emphasizes associative networks: pieces of information are associatively connected with one another, so that
activating one node (unit of information) on a network spreads activation to related nodes. Suppose a
participant in an experiment is presented
(primed) with the word bird and subsequently
asked to press a button as soon as she recognizes
each of a series of words. With priming, she will
respond more quickly to the word robin than to
the word butter. The reason is that bird and robin
are located along the same network of associations,
so that activating one spreads activation to the
other. Another way researchers have described the
organization of memory is in terms of schemas, pat-

T H E J O U R N A L O F L I F E L O N G L E A R N I N G I N P S Y C H I AT RY

WESTEN

THE

SECOND COGNITIVE REVOLUTION

In the last decade the modal model has undergone considerable evolution in four interrelated
respects, which probably constitute more of a revolution than an evolution in thinking. The first
change is a shift away from a serial processing

focus.psychiatryonline.org

model. In the modal model, stages of memory storage and retrieval occur sequentially, one at a time,
with most of the real work of cognition done by
bringing information into short-term memory.
Contemporary researchers, however, recognize that
most processing occurs outside of awareness, as the
brain processes multiple pieces of information in
parallel. Serial processing, in this view, is the task of
a specialized memory system, working memory (a
construct that evolved from the construct of shortterm memory, referring to a work space in which
the individual can consciously manipulate information; see Baddeley, 1995; Richardson, 1996).
A second and related shift is from conceiving of
memory as involving stores (places where memories are kept) to a view of memory and cognition
as involving multiple circuits or systems. For example, when a person sees an object, cortical circuits
involving the occipital and lower (inferior) temporal lobes are involved in breaking it into component
parts and comparing it with familiar objects, and a
second circuit running from the occipital lobes
through the upper (superior) temporal and parietal
lobes attempts to pinpoint its location in space. The
person is never aware of using different circuits to
identify an image and locate it in space, because
both circuits are part of a broader neutral circuit
that integrates the informationand does this so
quickly that the person has no phenomenological
experience of anything other than the immediate
recognition of having seen a squirrel running across
the road. This conception of memory systems is
bolstered by research showing that memory for
episodes (e.g., remembering what happened yesterday), memory about the emotional meaning of
stimuli (e.g., whether something has consistently
been associated with pain), memory for procedures
(e.g., playing a piece on the piano), and working
memory constitute neurologically distinct memory
systems. For example, memory for episodes requires
an intact hippocampus, but a person with hippocampal damage can still associate a stimulus with
an emotional response, even though he may have
no memory for having ever encountered it (e.g.,
Bechara et al., 1994). Working memory, in contrast,
is readily disrupted by lesions to the dorsolateral
prefrontal cortex, which is involved in deliberate
conscious thinking and decision making.
A third major shift has occurred with the recognition of the existence of two ways that memory can
be expressed, either explicitly (via conscious recall or
recognition) or implicitly (in behavior, independent
of conscious control). Explicit memory refers to conscious memory for ideas, facts, and episodes.
Implicit memory refers to memory that is observable
in behavior but is not consciously brought to mind

FOCUS

Spring 2006, Vol. IV, No. 2

INFLUENTIAL
P U B L I C AT I O N S

terns of thought that guide perception and memory. Thus, if an eye witness to an accident is asked
how quickly a car smashed into another car, she is
likely to estimate a higher speed than if asked how
quickly the car was going when it hit the other car,
because smash activates a schema that implies
high impact (Loftus et al., 1975).
This information-processing model offers a general view not only of memory but of thinking
that is, of the processes by which people
manipulate remembered information to solve
problems. According to the model that dominated
the field for 30 years (and remains the foundation
of many cognitive models of thought and decision
making, with some caveats; see Markman and
Gentner, 2000), when people want to make a decision, they use short-term memory to maintain current information, retrieve relevant information
from long-term memory, and perform various
operations on the information held there (Newell
and Simon, 1972; Klahr and Simon, 2001). Thus,
problem solving involves parsing a problem into an
initial state (how things currently are), a goal state,
and potential operators that might transform the
initial state into the goal state.
This way of thinking about cognition provided
the zeitgeist within which cognitive approaches to
psychotherapy developed in the 1960s (e.g., Ellis,
1962; A. T. Beck, 1967, 1995). Early cognitive
models of therapy tended to presume a serial model
of cognition, in which people feel and act based on
the thoughts that come into consciousness (or on
automatic thoughts that lie just outside the
periphery of awareness but can be readily retrieved
with proper cueing). An important goal of these
therapies is to change dysfunctional attitudes,
views of the self, and things people say to themselves that are associated empirically with negative
mood states such as depression and dysfunctional
behaviors such as bulimic binge-purge cycles.
Although the information processing models of the
1960s and 1970s were relatively silent about the
kinds of classical and operant learning processes
studied by behaviorists, in clinical practice by the
late 1970s cognitive-behavioral approaches began
to emerge that integrated behavioral techniques
with cognitive strategies designed to change dysfunctional thinking patterns.

217

WESTEN

(Roediger, 1990; Schacter, 1992, 1998). One kind


of implicit memory is procedural memory, which
refers to how to knowledge of procedures or skills,
such as how close to stand to another person or how
to respond when someone reaches out his or her
hand for a handshake. Another kind of implicit
memory involves associative memory. For example,
priming subjects with an infrequently used word
such as syncopate among a long list of words renders
them more likely a week later to respond with the
correct word when asked to fill in the missing letters
of the word fragment, S-----ATE. This occurs even
though they may lack any conscious recollection of
whether syncopate was on the list a week earlier
(Tulving et al., 1982). Essentially, the network of
associations still has some residual activation, leading to memory expressed in behavior but not in
conscious recollection.
A fourth shift involves a change in metaphor.
Cognitive psychologists in the late 1950s and early
1960s saw in the computer a powerful metaphor
for the human mind. Today, cognitive scientists are
turning to a different metaphor: mind as brain. In
this view, memory is not so much a matter of storing something somewhere in the brain and later
retrieving it (as in a computer file) than a process
by which an experience activates a set of neurons
distributed throughout the brain that can in turn
be reactivated by similar experiences or efforts at
recollection. In this view, memory is simply a
potential for reactivation of a set of neurons that
together constitute a representation. The notion of
using the brain as a metaphor for the mind may
seem today obvious if not tautological; however,
metaphors of mind have tended to follow understanding in other domains, particularly in the
physical sciences. Freud, for example, certainly
knew that what he referred to as mental processes
occur through the actions of brain processes, and
he developed some complex models of neural excitation that appear today to be remarkably prescient
in multiple respects (Freud, 1966; Pribram and
Gill, 1976; Westen, 1998). However, because
knowledge of the brain was so primitive, be turned
to metaphors from physics to explain how mental
dynamics function. Similarly, the information
processing theorists of the 1960s through 1980s
tended to draw their inspiration from computer
technology. Not until knowledge of the brain
expanded exponentially in the last two decades did
brain processes become potential metaphors for
mental processes.
The notion of mind as brain is central to connectionist, or parallel distributed processing, models of
perception, memory, and thinking (Rumelhart et
al., 1986; Kunda and Thagard, 1996; Smith,

218

Spring 2006, Vol. IV, No. 2

FOCUS

1998). Connectionist models suggest that most


information processing occurs in parallel, outside
of awareness, as multiple components of a thought,
memory, or perception are processed simultaneously. Representations are distributed throughout
the brain over many sets of neurons processing different aspects of a thought, perception, or memory,
rather than located in any particular part of the
brain. Knowledge lies in the connections among
these neural units or nodes, which, like neurons,
can either inhibit or activate each other. Cognitive
activity involves a process of constraint satisfaction,
in which the brain simultaneously and unconsciously processes multiple features of a stimulus,
attended to by different nodes or sets of nodes in a
network that provides constraints on the conclusions that can be drawn. The brain then draws the
best tentative conclusion it can based on the available data. In other words, it equilibrates to the
solution that provides the best fit to the data.
Thus, if a patient is crying, the clinicians interpretation of that crying as tears of pain or joy will
depend on auditory and semantic cues processed
simultaneously (in parallel).
Connectionist models have the advantage of
building in a way of modeling both the chronic
ways people tend to process information and
moment to moment changes in the way they view
important people and experiences in their lives (see
Barsalou, 1999). One of the virtues of connectionist models is their suggestion that representations,
such as a persons representations of significant others, are not static. Rather, the representation of a
significant other activated at any given point
depends on the context. Thus, the same person can
represent his wife as impossible to deal with at one
time but a source of loving support at another,
depending on aspects of his wife network activated by the current situation, his feeling state, and
so forth. At the same time, the chronic activation
of a way of seeing something or someonethat is,
the frequent activation of a set of neurons representing some aspect of that personwill create an
attractor state, a pattern of neural firing that is readily activated under particular circumstances. Thus,
a patient with a critical parent may be primed to
hear his therapists comments as criticisms because
a network representing self-being-criticized-byparental-figure is an attractor state that attracts the
brain to this interpretation. In this view, then, a
representation is not something stored in the
brain. It is a potential for reactivation of a set of neural units that have been activated together in the
past. Activating part of that network may reproduce much of the original experience (as in an
episodic memory, e.g., of a time the parent was

T H E J O U R N A L O F L I F E L O N G L E A R N I N G I N P S Y C H I AT RY

WESTEN

critical, or more directly in a flashback in posttraumatic stress disorder) or may influence the way the
person interprets current experiences.

IMPLICATIONS

FOR PSYCHODYNAMIC
PSYCHOTHERAPY

MAPPING

AND CHANGING IMPLICIT NETWORKS

Fundamental to all psychoanalytic forms of


treatment is the effort to map the idiosyncratic
associative networks that may be relevant to the
patients sources of distress. The goal, as first enunciated by Freud, was to give the patient more freedom to make conscious, explicit choices. Indeed,
in describing the process of open-ended, long-term
therapy to patients, it can be very useful to offer a
simple explanation such as the following:
Much at what we do reflects the way thoughts and
feelings have gotten connected in our minds. But we
have no direct access to those connections. So you find
yourself bingeing and then vomiting but dont really
know why youre doing it and cant find a way to
stop. In many ways, our task together is to map those

focus.psychiatryonline.org

The existence of unconscious or implicit networkswhich tend to he resistant to change


because they reflect longstanding regularities in the
persons experience and allow him or her to navigate the world in ways that feel predictable (even if
sometimes rigid, inaccurate, or otherwise maladaptive)provides perhaps the best empirical justification for long-term therapies. Deeply engrained
views of the self, otherscalled internal working
models of relationships in research on attachment
(see, e.g., Bowlby, 1973; Main et al., 1985; Fonagy
et al., 2002) and object representations in theory
and research on interpersonal functioning more
broadly in psychoanalysis (see, e.g., Greenberg and
Mitchell, 1983; Westen, 1990, 1991; Blatt et al.,
1997)may take months or years to identify in
their various manifestations. The same is true of
problematic ways of regulating emotions (defenses)
that are triggered automatically and may lead to a
cascade of internal and interpersonal events.
For example, patients with prominent passiveaggressive features are often unaware of both their
anger and the ways they put other people in
uncomfortable positionswhich in turn lead others to become angry at or avoid them. This, in turn,
makes the patient more angry and passive-aggressive. Consciously, these patients view themselves as
helpless victims of indifferent or mean-spirited others; unconsciously, they provoke precisely the
behavior that makes them feel mistreated. Breaking
into these kinds of self-sustaining spiralsinto
what Wachtel (1997) calls cyclical psychodynamics
can take a long time, because the patient cannot
report them. Such dynamics may become most
apparentand most workable as a treatment
issuewhen they show up in the therapeutic relationship (Luborsky and Crits-Christoph, 1998).

INFLUENTIAL
P U B L I C AT I O N S

Although psychoanalytic practice has largely


evolved from the consulting room independent of
experimental research, in many respects, recent
developments in the cognitive neurosciences have
breathed new empirical life into psychodynamic
forms of psychotherapy, bolstering the basic science
behind them even if the applied science (treatment
research) lags far behind. The second cognitive revolution documented perhaps the most central psychoanalytic hypothesis, and the one that
distinguished it from other approaches to the mind
and treatment for a century: that unconscious associative networks and unconscious procedures (e.g.,
defenses, motives) influence thought, feeling, and
behavior outside of awareness. The research evidence is now clear that much of the way people
view themselves and others is implicit or unconscious; that their brains are frequently triggered or
primed to behave or interpret events in certain
ways based on the implicit activation of networks of
which they have no awareness; that they can have
emotional reactions of which they are unaware; that
they can regulate emotions outside of awareness to
avoid painful feelings (what psychoanalysts call
defense); and that the same event can trigger contradictory thoughts, feelings, or actions consciously
and unconsciously (such as negative racial attitudes
in people who consider themselves free of racism, or
devalued views of self in patients who present with
grandiosity) (Westen, 1998).

connections in your head, so we can figure out whats


leading you to do things youd rather not do and to
begin developing new connections.

TECHNIQUES FOR EXPLORING ASSOCIATIVE


NETWORKS: FREE AND DIRECTED ASSOCIATION
Thus, contemporary research in cognitive neuroscience corroborates some central psychoanalytic
assumptions that have been the source of tremendous controversy for a century. At the same time,
this research also poses some important challenges
for psychodynamic psychotherapy and suggests
potential refinements in theory and technique (see
Westen, 2002; Westen and Gabbard, 2002a;
Westen and Gabbard, 2002b; Gabbard and Westen,
2003). For example, research in cognitive neuroscience suggests precisely why the psychoanalytic

FOCUS

Spring 2006, Vol. IV, No. 2

219

WESTEN

practice of exploring patients associations to symptoms, feelings, or eventsasking them what comes
to mindcan often be very useful: people cannot
report on their implicit networks, and they typically
invent plausible but often inaccurate explanations if
called upon to do so (e.g., when asked, Why do
you think you felt that way?; Nisbett and Wilson,
1977). On the other hand, this same body of
research suggests limits to free association as a therapeutic technique, on two grounds.
First, although free association can be essential in
exploring implicit networks, it may do very little to
illuminate or alter explicit (conscious) beliefs, procedures, or ways of behaving that operate through
the action of different neural networks. As argued
below, with limited therapeutic time (even for
patients treated more than once a week), attention
to implicit processes inherently comes at the
expense of attention to explicit processes, which
can also wreak havoc on a persons quality of life,
and nothing guarantees that even emotionally
important change in implicit expectations,
motives, feelings, or conflicts will alter conscious
habits of thought or behavior that have attained
functional autonomy over years or decades of use.
This recognition is precisely what led Aaron Beck
(1976) to develop cognitive therapy for depression.
Second, research in cognitive science suggests
that what is on a persons (unconscious) mind at
any time is a joint function of what is chronically on
his mind (much of which is likely, in fact, to reflect
concerns forged in childhood) and what is recently
on his mind, which may or may not be related to
the concerns that brought the patient to treatment.
In other words, the particular associations that
emerge in any analytic hour if the patient follows
what Freud called the fundamental rule of psychoanalysis (namely, to say whatever comes to
mind) may or may not prove useful to explore,
depending on what has been activated recently in
and out of the consulting room. Any given set of
associations reflects some combination of clinically
meaningful signal and clinically less meaningful
noise, and one cannot always distinguish the two.
Over time one would expect important material to
be reflected repeatedly in the patients associations,
as chronically activated networks influence the
patients thought, feeling, and behavior in the treatment. However, waiting for important material to
emerge, particularly in the context of therapeutic
interventions (particularly interpretations) that
shape subsequent associations, is likely to be an
inefficient process.
Patients can also avoid doing things associated
with anxiety, such as allowing themselves to fall in
love because doing so is associated with anxiety or

220

Spring 2006, Vol. IV, No. 2

FOCUS

fear of rejection. As a result, some of the most


important networks may never be activated to the
extent necessary for useful exploratory work until
the patient actually exposes herself to the feared
situation. Under such circumstances, the therapist
may do well to encourage the patient to approach
what she fears, to alter the patients associative
networks and/or to bring material to the fore in
the patients associative networks that are most
important in maintaining maladaptive patterns.
Freud himself noted that people do not get free of
their fears unless they confront them, and he
practiced a much more active mode of therapy
than practiced by subsequent generations of analysts for many years. As many psychoanalytically
oriented clinicians now recognize, good treatment
probably requires a balance of exploration and
exhortation at the service of further exploration
and behavioral change (see Wachtel, 1997;
Gabbard and Westen, 2003).
One useful way to employ associative techniques
in once- or twice-weekly psychotherapy, where one
does not have the luxury to explore whatever associations come to mind at any given time, is what
might be called directed free association, in which
the therapist targets particular thoughts, feelings,
or memories for further associative work (see
Westen, 2000b).
For example, one patient had a pattern of
becoming excited about some plan (e.g., spending
the evening out with friends) but then finding
himself depressed and unable to imagine that he
would really enjoy doing it. (For a sophisticated
cognitive-dynamic explanation of the way
patients shift between such states of mind, see
Horowitz, 1979.) As a result, the patient experienced few pleasures in life. After the pattern
became clear, I routinely asked him to imagine as
vividly as possible what he initially thought and
felt when he was excited about the plan or to picture the moment he found himself feeling
depressed and uninterested in pursuing it. I
would ask him to walk me through the episode or
image moment by moment or scene by scene, taking associations along the way, much as Freud
would have explored a dream, encouraging him to
report whatever he felt at the time and whatever
thoughts, feelings, images, or memories emerged
as he pictured the experience. Doing so led to a
series of associations and memories in which he
wished for something that subsequently fell
through and his corresponding fear of hoping for
enjoyment, as well as survivor guilt around a
mentally retarded sister who could never have
such pleasures and toward whom he felt a mixture
of love and (largely unacknowledged) resentment.

T H E J O U R N A L O F L I F E L O N G L E A R N I N G I N P S Y C H I AT RY

WESTEN

UNDERSTANDING

TRANSFERENCE PROCESSES

focus.psychiatryonline.org

Perhaps the most important lesson to be learned


from developments in the cognitive neurosciences
is that clinicians need to attend in psychotherapy to
both implicit processes (emphasized by psychoanalysis and to some extent by both cognitive and
behavior therapy) and explicit processes (emphasized by cognitive therapy). One cannot assume
that the same techniques likely to change explicit
thought processes will change implicit networks
and vice versa (Westen, 2000b; Gabbard and
Westen, 2003). Indeed, data from the cognitive
neurosciences suggest that implicit and explicit
processes often reflect neuroanatomically distinct
brain systems and that what registers implicitly and
explicitly can be very different.
Psychotherapists and treatment researchers need
to think more carefully about, and study empirically,
the tradeoffs inherent in attention to implicit and
explicit processes in psychotherapy. Every time clinicians explore the meaning of a self-critical statement,
they are choosing not to try to alter an explicit
process directly. Every time they explore the meaning of an anxiety symptom, they are only indirectly,
if at all, using exposure techniques that might he
applied therapeutically in much more direct ways to
alter the feeling state. Conversely, every time clinicians draw a patients attention to a self-critical statement as a way of trying to alter current mood or
address an explicit way the patient talks to herself,
and every time they use exposure techniques to try
to change an affective association, they are altering
the conditions that would allow optimal exploration
of its implicit meanings. What is exciting about
developments in cognitive neuroscience is that they
may help clinicians, theorists, and researchers begin
to address crucial issues such as this by calling attention to multiple systems that will likely require multiple types of therapeutic intervention.

INFLUENTIAL
P U B L I C AT I O N S

Data from the cognitive neurosciences may also


help shed new light on psychoanalytic constructs
such as transference. The connectionist notion of
representations as potentials for reactivationas
sets of neurons that have been activated in the
past and are hence more readily activated as a
unit in the futureoffers a mechanism to explain
the long-held psychoanalytic position that
patients are likely to express important conflicts,
defenses, motives, and interpersonal patterns in
their relationship with the therapist (Westen and
Gabbard, 2002b). To the extent that the therapeutic situation or relationship matches prototypes from the past, it is likely to activate similar
responses (for empirical evidence, see Andersen
and Baum, 1994; Luborsky and Crits-Christoph,
1998). It should therefore not be surprising if
important relational patterns emerge in a relationship in which the patient self-discloses and
becomes attached in an intimate but asymmetrical relationship with another who is trying to be
helpful, nurturant, and attentive primarily to his
or her needs. Inherent in the cognitive situation of
the therapeutic relationship is the likelihood that
the therapist will be experienced as an authority
figure, an attachment figure, or an object of love
or affection, which renders exploration of the
therapeutic relationship of particular use if the
patient is presenting with problems that include
interpersonal components (which is nearly always
the case).
Important dynamics are likely to emerge relatively quickly and persistently in patients with rigid
maladaptive patterns of interpersonal functioning,
cognition, and emotion regulation. For example,
narcissistic patients tend to manifest particular patterns when interacting with their therapists
regardless of the therapists theoretical understanding or technical approach. Empirically, clinicians of all therapeutic orientations report that
narcissistic patients need excessive admiration
from them, vacillate between idealizing and
devaluing them, and need to be special to the
therapist at the same time as being sadistic and
hostile and feeling criticized by the therapist
(Bradley et al., in press). Correspondingly, therapists of all theoretical orientations report similar
countertransference reactions to their narcissistic
patients: They tend to feel annoyed, manipulated,
used, criticized and as if they are walking on
eggshells with the patient, and correspondingly
frequently fight their impulses to be sadistic
themselves or to drift off during sessions (Betan et
al., in press).

CONCLUSIONS

REFERENCES
Andersen, S. M. and Baum, A. (1994). Transference in interpersonal relations:
inferences and affect based on significant-other representations. Journal
of Personality, 62, 45997.
Atkinson, R. and Shiffrin, R. (1968). Human memory: a proposed system and
its control processes. In: K. W. Spence and J. T. Spence, ed. The psychology
of learning and motivation, Vol. 2, pp. 89195. New York: Academic Press.
Baddeley, A. D. (1995). Working memory. In: M. S. Gazzaniga, ed. The cognitive
neurosciences, pp. 75564. Cambridge, MA: MIT Press.
Barlow, D. (2002). Anxiety and its disorders, 2nd edn. New York: Guilford Press.
Barsalou, L. W. (1999). Perceptual symbol systems. Behavioral and Brain
Sciences, 22, 577660.
Bechara, A., Damasio, A. R., Damasio, H., and Anderson, S. W. (1994).
Insensitivity to future consequences following damage to human prefrontal
cortex. Cognition, 50, 715.
Beck, A. T. (1967). Depression: clinical, experimental, and theoretical aspects.
New York: Harper and Row.
Beck, A. T. (1976). Cognitive therapy and the emotional disorders. New York:
International Universities Press.
Beck, A. T. (1995). Cognitive therapy: present, and future. In: M. J. Mahoney,

FOCUS

Spring 2006, Vol. IV, No. 2

221

WESTEN
ed. Cognitive and constructive psychotherapies: theory, research, and
practice, pp. 2940. New York: Springer.
Betan, E., Zittel, C., Heim, A., and Westen, D. (in press). The structure of countertransference phenomena in psychotherapy: an empirical investigation.
Unpublished manuscript, Emory University.
Blatt, S. J., Auerbach, J. S., and Levy, K. N. (1997). Mental representations in
personality development, psychopathology, and the therapeutic process.
Review of General Psychology, 1, 35174.
Bowlby, J. (1973). Separation, Vol. 2. London: Hogarth Press.
Bradley, R., Heim, A., and Westen, D. (in press). Transference phenomena in
the psychotherapy of personality disorders: an empirical investigation.
Unpublished manuscript, Emory University.
Brody, A., et al. (2001). Regional brain metabolic changes in patients with
major depression treated with either paroxetine or interpersonal therapy:
preliminary findings. Archives of General Psychiatry, 58(7), 63140.
Ellis, A. (1962). Reason and emotion in psychotherapy. New York: Lyle Stuart.
Fonagy, P., Gergely, G., Jurist, E. L., and Target, M. (2002). Affect regulation,
mentalization, and the development of the self. New York, NY: Other Press.
Freud, S. (1966). Project for a scientific psychology. In: J. Strachey, ed. The
standard edition of the complete psychological works of Sigmund Freud,
Vol. 1, pp. 281397. London: Hogarth Press.
Gabbard, G. and Westen, D. (2003). On therapeutic action. International Journal
of Psycho-Analysis, 84, 82341.
Goldapple, K., et al. (2004). Modulation of cortical-limbic pathways in major
depression. Archives of General Psychiatry, 61, 3441.
Greenberg, J. R. and Mitchell, S. (1983). Object relations in psychoanalytic theory. Cambridge, MA: Harvard University Press.
Healy, A. F. and McNamara, D. S. (1996). Verbal learning and memory: does the
modal model still work? Annual Review, 47, 14372.
Horowitz, M. (1979). States of mind: analysis of change in psychotherapy. New
York: Plenum.
Klahr, D. and Simon, H. A. (2001). What have psychologists (and others) discovered about the process of scientific discovery? Current Directions in
Psychological Science, 10, 759.
Kunda, Z. and Thagard, P. (1996). Forming impressions from stereotypes, traits,
and behaviors: a parallel-constraint-satisfaction theory. Psychological
Review, 103, 284308.
Loftus, E. F., Altman, D., and Geballe, R. (1975). Effects of questioning upon a
witness later recollections. Journal of Police Science and Administration,
3, 1625.
Luborsky, L. and Crits-Christoph, P. (1998). Understanding transferenceThe
core conflictual relationship theme method, 2nd edn. Washington, DC:
American Psychological Association.
Main, M., Kaplan, N., and Cassidy, J. (1985). Security in infancy, childhood, and
adulthood: a move to the level of representation. Monographs of the
Society for Research in Child Development, 50(1/2), 66104.
Markman, A. B. and Gentner, D. (2000). Thinking. Annual Review of Psychology,
52, 22347.
Miller, G. (1956). The magical number seven, plus or minus two: some limits on
our capacity for processing information. Psychological Review, 63,8197.

Newell, A. and Simon, H. A. (1972). Human problem solving. Englewood Cliffs,


NJ: Prentice-Hall.
Nisbett, R. E. and Wilson, T. D. (1977). Telling more than we can know: verbal
reports on mental processes. Psychological Review, 84, 23159.
Pribram, K. and Gill, M. (1976). Freuds project reassessed. New York: Basic
Books.
Richardson, J. T. E. (1996). Evolving concepts of working memory. In: J. T. E.
Richardson, R. W. et al., ed. Working memory and human cognition. New
York: Oxford University Press.
Robins, R. W., Gosling, S. D., and Craik, K. H. (1999). An empirical analysis of
trends in psychology. American Psychologist, 54, 11728.
Roediger, H. L. (1990). Implicit memory: retention without remembering.
American Psychologist, 45, 104356.
Rumelhart, D. E., McClelland, J. L., and PDP Research Group. (1986). Parallel
distributed processing: explorations in the microstructure of cognition: Vol.
1, Foundations. Cambridge, MA: MIT Press.
Schacter, D. L. (1992). Understanding implicit memory: a cognitive neuroscience approach. American Psychologist, 47, 55969.
Schacter, D. L. (1998). Memory and awareness. Science, 280, 5960.
Smith, E. R. (1998). Mental representation and memory. In: D. T. Gilbert, S. T.
Fiske, and G. Lindzey, ed. Handbook of Social Psychology, Vol. 1. pp.
391445. New York: McGraw-Hill.
Tulving, E., Schacter, D. L., and Stark, H. A. (1982). Priming effects in wordfragment completion are independent of recognition memory. Journal of
Experimental Psychology: Learning, Memory, and Cognition, 8, 33642.
Wachtel, P. (1997). Psychoanalysis, behavior therapy, and the relational world.
Washington, DC: American Psychological Association Press.
Westen, D. (1990). Towards a revised theory of borderline object relations:
contributions of empirical research. International Journal of
Psychoanalysis, 71, 66193.
Westen, D. (1991). Social cognition and object relations. Psychological Bulletin,
109, 42955.
Westen, D. (1998). The scientific legacy of Sigmund Freud: toward a psychodynamically informed psychological science. Psychological Bulletin, 124, 33371.
Westen, D. (2000a). Commentary: Implicit and emotional processes in cognitivebehavioral therapy. Clinical PsychologyScience and Practice, 7, 38690.
Westen, D. (2000b). Integrative psychotherapy: integrating psychodynamic and
cognitive-behavioral theory and technique. In: C. R. Snyder and R. Ingram,
ed. Handbook of psychological change psychotherapy processes and practices for the 21st century, pp. 21742. New York: Wiley.
Westen, D. (2002). Implications of developments in cognitive neuroscience for
psychoanalytic psychotherapy. Harvard Review of Psychiatry, 10, 36973.
Westen, D. and Gabbard, G. O. (2002a). Developments in cognitive neuroscience: I. Conflict, compromise, and connectionism. Journal of the
American Psychoanalytic Association, 50, 5398.
Westen, D. and Gabbard, G. O. (2002b). Developments in cognitive neuroscience: II. Implications for theories of transference. Journal of the
American Psychoanalytic Association, 50, 99134.
Wolpe, J. (1958). Psychotherapy by reciprocal inhibition. Palo Alto, CA: Stanford
University Press.

NOTES

222

Spring 2006, Vol. IV, No. 2

FOCUS

T H E J O U R N A L O F L I F E L O N G L E A R N I N G I N P S Y C H I AT RY

Você também pode gostar