You are on page 1of 323

The Neuroscience

of Psychotherapy
Healing the Social Brain

• Third Edition •

Louis Cozolino

W. W. Norton & Company


Independent Publishers Since 1923
New York • London

A NORTON PROFESSIONAL BOOK

2
The Norton Series on Interpersonal Neurobiology
Louis Cozolino, PhD, Series Editor
Allan N. Schore, PhD, Series Editor, 2007–2014
Daniel J. Siegel, MD, Founding Editor

The field of mental health is in a tremendously exciting period of growth and conceptual
reorganization. Independent findings from a variety of scientific endeavors are converging in an
interdisciplinary view of the mind and mental well-being. An interpersonal neurobiology of human
development enables us to understand that the structure and function of the mind and brain are shaped
by experiences, especially those involving emotional relationships.
The Norton Series on Interpersonal Neurobiology provides cutting-edge, multidisciplinary views
that further our understanding of the complex neurobiology of the human mind. By drawing on a wide
range of traditionally independent fields of research—such as neurobiology, genetics, memory,
attachment, complex systems, anthropology, and evolutionary psychology—these texts offer mental
health professionals a review and synthesis of scientific findings often inaccessible to clinicians. The
books advance our understanding of human experience by finding the unity of knowledge, or
consilience, that emerges with the translation of findings from numerous domains of study into a
common language and conceptual framework. The series integrates the best of modern science with the
healing art of psychotherapy.

3
This book is dedicated to my family:
my mother’s courage, my father’s determination, and
the memory of my grandparents. Together they somehow
instilled within me the belief that all things are possible.

4
Contents

Preface to the Third Edition

Part I. Neuroscience and Psychotherapy: An Overview


1. The Entangled Histories of Neurology and Psychology
2. Building and Rebuilding the Brain: Psychotherapy and Neuroscience
3. Neural Integration in Different Models of Psychotherapy

Part II. How the Brain Works: The Legacy of Evolution


4. The Human Nervous System: From Neurons to Neural Networks
5. Multiple Memory Systems in Psychotherapy
6. Laterality: One Brain or Two?

Part III. Executive Functioning and Neural Integration


7. The Executive Brain: Directed Action and Inhibition
8. The Executive Brain: Navigating Space and Time
9. The Executive Brain: Discovering Others and Finding the Self
10. From Neural Networks to Narratives: The Quest for Integration

Part IV. Attachment and Connectedness


11. The Social Brain
12. Building the Social Brain: Shaping Attachment Schemas
13. The Neurobiology of Attachment
14. Altruism and Psychotherapy: Leveraging the Social Brain in the Service of Change

Part V. The Disorganization of Experience


15. The Anxious and Fearful Brain
16. Early Traumatic Stress: The Fragmentation of Self and Others
17. The Impact of Trauma: Biochemical Dysregulation and Neural Network Dissociation
18. The Self in Exile: Narcissism and Pathological Caretaking

Part VI. The Reorganization of Experience


19. The Evolutionary Necessity of Psychotherapy
20. Stimulating Neural Plasticity
21. The Psychotherapist as Neuroscientist
22. How People Change
Credits

5
References
Acknowledgments
Index

6
Preface to the Third Edition

Since the publication of the first edition of The Neuroscience of Psychotherapy (2002), the exploration of the
topography, functioning, and social nature of the brain has grown deeper, wider, and increasingly
sophisticated. There is a wonderful confluence of older theories being supported by new findings and new
theories coming along with an abundance of findings. We are gradually moving away from a simplistic neural
phrenology to an appreciation of the emergence of mind and conscious experience from complex systems
interactions within and between individuals. We are seeing the study of social functioning as an increasingly
central aspect of the field of neuroscience and the study of interpersonal neurobiology further penetrating the
standard educational practices of therapists and counselors across multiple professional disciplines.
It is especially gratifying to see researchers turn their attention to questions traditionally taboo in the
sciences: consciousness, emotion, attachment, love, and altruism are finding their place within the research
space. This openness to subjective experience is supporting the consilience of areas of study as diverse as
Buddhism, wisdom philosophy, and feminist philosophy. You will discover in this edition a much broader
exploration of executive functioning as I describe a model of three executive networks: a network of systems
with a range of specialties that allow us to react to danger, successfully navigate the world, establish and
maintain attachment, and create a sense of self and internal reality. This model is tied, in a later chapter, to
the long-term social and psychological impact of early trauma and the development of the default mode
network. You will also see a new chapter in this edition on the neurobiology of altruism and the possible
positive impact that altruistic behavior may have on the therapeutic process.

Lou Cozolino
Los Angeles, 2017

7
The Neuroscience of Psychotherapy
Third Edition

8
Part I
Neuroscience and
Psychotherapy: An Overview

9
Chapter 1
The Entangled Histories of
Neurology and Psychology
The seemingly irreconcilable dichotomies . . . with respect to mind vs. matter . . . become reconciled in a . . .
unifying view of mind, brain, and man in nature.
—Roger Sperry

How does the brain give rise to the mind? How does the mind impact the brain and by what means do they
interact with one another? These are difficult questions—so difficult, in fact, that the common reaction is to
focus on either the mind or the brain and act as if the other is irrelevant (Blass & Carmeli, 2007; Pulver,
2003). The problem with this approach is that it creates a barrier to understanding that the brain and mind
are a unified process (Cobb, 1944). Neurology and psychology are simultaneously pushed apart by academic
and professional politics and drawn together by their common psychobiological foundation. The entangled
histories of neurology and psychology reflect the push and pull of these opposing forces (Ellenberger, 1970;
Sulloway, 1979).
Freud started out as a rebel, a neurologist curious about the mind. I suspect that he was frustrated with the
mind-brain partisanship of medical school and longed to work with others who shared his interests. At the
age of 29, Freud won a traveling fellowship to spend the fall and winter of 1885 at the Salpêtrière Hospital on
the left bank of Paris. The choice of the Salpêtrière was based on the reputation of Professor Jean-Martin
Charcot. In Charcot, Freud sought a teacher who was well established, confident, and unafraid of the no-
man’s-land between mind and brain. One can imagine Freud’s excitement as he walked the streets of Paris on
his way to meet the great man, a possible kindred spirit.
Charcot specialized in patients suffering from what was then called hysteria. These patients had symptoms,
such as seizures or paralysis, that mimicked neurological illnesses but were without apparent physical cause. A
classic example is a condition called glove anesthesia, in which feeling is lost in one or both hands beginning at
the wrist. In these patients, the hands appear to take on symbolic significance; perhaps they have been used to
commit some taboo act that triggered overwhelming guilt or fear. It was believed that a conflict within the
mind was converted into a bodily symptom.
The 1880s were also a time when the ability of the subconscious mind to control behavior (as
demonstrated through hypnosis) burst into popular awareness. Charcot used hypnosis during clinical
demonstrations to illustrate his emerging theories about mind-body interactions. The months Freud spent at
Salpêtrière with Charcot had a profound effect on him. Freud came to believe that hidden mental processes
do indeed exert powerful effects on consciousness, and that hysterical symptoms result not from malingering
or feigning illness, but from the power of the unconscious mind encoded within the neural structures of the
brain. From this perspective, hysteria reflected the capacity of traumatic experience to reorganize the brain and
disrupt conscious experience.
Dissociative splits between consciousness and behavior led Freud to believe that the brain is capable of
multiple levels of conscious and unconscious awareness. In the decades to come, he would explore the use of
language, emotion, and the therapeutic relationship to reconnect them. Freud returned to Vienna in February
1886 and opened a clinical practice two months later. Despite his entry into the medical establishment, he
continued his rebellion later that year with the presentation of a paper on the existence of hysteria in males.
Deeply fascinated by the unconscious, Freud remained its most ardent explorer until his death in 1939.
In his writings, Freud expanded on Charcot’s thinking in many significant ways. He placed the
unconscious in a developmental context by tracing the genesis of hysterical symptoms to childhood
experiences. He came to believe that hysterical patients suffered from the unconscious emotional aftereffects
of repressed childhood memories. Furthermore, Freud connected the development of the individual to the
evolution of the species. Influenced by the ancient idea that we contain within us the biological history of our
primitive ancestors, he included the importance of instinctual drives such as sexuality, rage, and envy in his

10
developmental theories. Freud believed that beneath our civilized exteriors there exist more primitive beings,
which accounts for many of the conflicts and contradictions of “civilized” behavior.
Freud argued that in order to understand who and what we are, we need to accept the primal unconscious
elements of experience. He called this the id—the primitive and uncivilized life energy that we share with our
reptilian and mammalian ancestors. This concept was met with understandable hostility by Freud’s rational
contemporaries. At that time, physicians were pillars of European culture, highly invested in their superiority
over the animal kingdom and steadfast in their right to subjugate the “primitive” people of the world.
Needless to say, linking civilized humans to animals (to say nothing of his idea that children have sexual
desires) made Freud and his theories scandalous in respectable circles.

Freud’s Abandoned Project


We must recollect that all of our provisional ideas in psychology will presumably one day be based on an organic substructure.
—Sigmund Freud

In the late 1800s, the microscopic world of the nervous system was opened for the first time. Technical
improvements in the microscope and newly developed staining techniques led to the discovery of both
neurons and the synapses through which they communicate. The existence of synapses revealed that the
nervous system was not a single structure as was formerly believed. Rather, it is comprised of countless
individual structures. Furthermore, that humans shared these neurons with all other living creatures supported
the Darwinian theory of common ancestry. Around the same time, the work of Wernicke and Broca showed
that specific areas of the brain were responsible for different aspects of language. The dual neuroanatomical
notions of synaptic transmission and the localization of specific functions provided rich theoretical soil for new
ways of understanding the brain.
Inspired by Charcot, Darwin, and the opening of the microscopic neural world to investigation, Freud
wrote “Project for a Scientific Psychology” (as it came to be called) (Freud, 1968). In the “Project,” he
postulated that what we witness of conscious and unconscious behavior is organized by and stored within the
brain’s neural architecture. As part of this work, he drew simple sketches of interconnecting neurons to
represent human impulses, behaviors, and psychological defenses. These sketches depicted the interactions
among the drives, the organs of the senses, and mechanisms of inhibition. According to his colleagues, Freud
became obsessed with the idea of constructing a neurobiological model of the mind (Schore, 1997b). Despite
his enthusiasm, Freud realized that his desire for psychology to be based in an understanding of the nervous
system was far ahead of its time and at odds with prevailing religious and medical dogmas. For these and
other reasons, he suppressed the publication of the project until his death.
Perhaps Freud kept the project to himself because he feared that it would be relegated to the same sort of
obscurity as the case of Phineas Gage. Gage, a 19th-century railroad foreman, had a metal bar pass completely
through his head as a result of an accident, causing the destruction of the middle portions of his frontal cortex.
Although Gage had no specific motor or language deficits, those who knew him said that “Gage was no
longer Gage” (Benson, 1994). His emotionality, relationship abilities, and the quality of his experience were
all dramatically altered. Because Gage’s symptoms involved his personality and emotions, the publication
reporting his case received little attention for most of the 20th century. Not only was it outside the realm of
behaviors that neurologists felt comfortable addressing, but there was also a bias against relating human
personality to neurobiological mechanisms (Damasio, 1994). This particular area of the brain has since been
shown to be involved with judgment, planning, and emotional control.
Freud, the neurologist, became all but forgotten as his psychological theories moved further and further
from their biological roots. He chose instead to utilize the more palatable and accessible metaphors of
literature and anthropology to provide the primary vocabulary for psychoanalysis. Unfortunately, Freud’s shift
from the brain to metaphors of the mind opened psychoanalysis to all sorts of criticism. Metaphors such as
Oedipal and Electra complexes were seen as contrived fictions, shielding Freud’s work from scientific
evaluation. Perhaps Freud anticipated that psychoanalysis would eventually be integrated with neurobiology
when the time was right for a synthesis based in an equal partnership (Pribram & Gill, 1976). The time for
such an integration has arrived.
Respect for psychological processes has taken a strong hold within the scientific community and general
culture. We can discuss the mind and brain while avoiding a reduction of the mind to basic biochemical
processes. On the contrary, an appreciation for the structures and functioning of the brain by nonneurologists
is becoming the norm. It is in this spirit that we turn our attention to ways of thinking about the brain that

11
enhance our understanding of human experience. We begin with a model of the brain that provides a bridge
between the fields of neuroscience, evolution, and the origins of the unconscious.

The Triune Brain


He who joyfully marches in rank and file . . . has been given a large brain by mistake, since for him the spinal cord would suffice.
—Albert Einstein

In the 1970s, the neuroscientist Paul MacLean presented a theory that emphasized the conservation of more
primitive structures within the modern human brain (MacLean, 1990; Taylor, 1999). MacLean called his idea
the triune brain. Very much in line with the theories of Darwin and Freud, it provides an evolutionary
explanation that may account for some of the contradictions and discontinuities of human consciousness and
behavior.
MacLean described the human brain as a three-part system that embodies our evolutionary connection to
both reptiles and lower mammals such as mice and horses. Think of it as a brain within a brain within a brain,
with each successive layer devoted to increasingly complex functions and abilities. At the core is the reptilian
brain, relatively unchanged through evolutionary history, responsible for activation, arousal, homeostasis, and
reproductive drives. The paleomammalian brain or limbic system, which is central to learning, memory, and
emotion, wraps around the reptilian brain. The highest and outermost layer, the neomammalian brain or
cerebral cortex, organizes conscious thought, problem solving, and self-awareness (MacLean, 1985).
MacLean (1990) suggested that our three brains don’t necessarily communicate or work well together
because of their differing “mentalities” and the fact that only the neomammalian brain is capable of
consciousness and verbal communication. This idea provides fundamental connections among evolution,
neuroscience, and psychotherapy. What Charcot and Freud called dissociation and hysteria could well be the
result of inadequate integration and coordination among these different, cohabiting brains. MacLean’s
description of the nonverbal reptilian and paleomammalian brains unconsciously influencing processing in the
neomammalian brain roughly parallels Freud’s notion of the unconscious mind’s ability to influence conscious
experience.
The model of the triune brain serves the valuable function of providing a connective metaphor among the
artifacts of evolution, the contemporary nervous system, and some of the inherent difficulties in the
organization of human experience. This conservation of our evolutionary history alongside and interwoven
with our modern neural networks confronts the therapist with the unique challenge of simultaneously treating
a human, a horse, and a crocodile (Hampden-Turner, 1981).

Ah, If Only It Were So Simple!


The large brain, like large government, may not be able to do simple things in simple ways.
—Donald Hebb

A superficial reading of MacLean’s work might lead us to the idea that each layer of the triune brain evolved
independently and sequentially and that they all cooperate in a hierarchical fashion like a military chain of
command. This is clearly not the case. In reality, the reptilian and paleomammalian brains have continued to
evolve along with the neomammalian brain. Earlier structures are not conserved unchanged from past
generations, but also undergo a process of exaptation—the modification of earlier-evolving brain structures for
new applications in networks dedicated to alternative or more complex functions (Cacioppo & Berntson,
2004).
All three layers continue to evolve along with the emergence of ever more complex vertical and horizontal
neural networks that connect them. This conservation and modification of neural networks has led to an
amazingly complex brain capable of a vast array of functions from monitoring respiration to performing
mathematical computations. All of this complexity, grounded in millions of years of evolution, makes
understanding functional neuroanatomy brain-behavior-experience relationships quite a challenge.
An example from space exploration may prove useful in understanding the neuroanatomist’s dilemma.
When Apollo 13 approached the moon, difficulties with the air supply system left the crew with just a few
hours of oxygen (Lovell & Kluger, 1994). In the face of this crisis, scientists on earth removed nonessential
components from a mock spacecraft and constructed a new air supply system. Pieces of upholstery, plastic

12
bags, duct tape, and electrical wiring were used in innovative ways to serve new functions. The instructions on
how to build this makeshift device were then conveyed to the Apollo 13 crew. This scenario is much closer to
the crafting of the modern brain than imagining an engineer sitting down with a blank sheet of paper. An
engineer of the future presented with this bootstrapped air purification system would have a difficult time
figuring out what it was and why it was built the way it was. Although there are obvious differences between
the Apollo 13 scenario and natural selection, both are examples of a pragmatic adaptation with existing
materials to an environmental crisis.
The multiple roles played by the cerebellum offer a prime example of both neural conservation and
exaptation. The cerebellum is a primitive brain structure. At its core is the vermis, centrally involved in
balance. In fish, the vermis helps them to swim upright. In humans, it coordinates vestibular functioning,
allowing us to sit up and walk without falling over. During evolution, as our brains and bodies became more
complex, the cerebellum expanded to coordinate gross and then fine motor movements—a logical
development for a structure initially at the core of a fish’s ability to swim upright. In an interesting and
surprising twist, the later-evolving portions of the cerebellar lobes were utilized in the organization and
coordination of language, memory, emotion,and reasoning (Baillieux et al., 2008; Lupo et al., 2015;
Schmahmann, 1997).
It appears that the cerebellum’s ability to process, sequence, and organize vast amounts of sensory-motor
information was leveraged by the evolving brain as part of the neural infrastructure of higher cortical
processes. Just as balance and motor behavior require constant monitoring of posture and the inhibition of
unnecessary and distracting movements, so, in their own ways, do attention, concentration, memory, and
language. The same timing mechanisms involved in locomotion seem to have been conserved for the
sequential processing of thoughts and words. Although the cerebellum is considered a primitive brain
structure, its evolution involved vertical networking with most of the cortex, suggesting that vertical networks
that connect the horizontal layers of the brain may serve as clues to its evolutionary history (Alexander,
DeLong, & Strick, 1986; Cummings, 1993).
In addition to horizontal and vertical networks, evolution has also selected for increased differentiation
between the left and right hemispheres. Certain areas of the brain have become specialized for specific skills,
such as language and spatial abilities. Other areas, such as those in the prefrontal cortex, serve to organize and
control the activity of multiple other regions. Keep in mind that the brains of men and women also have many
differences and that the brain changes as we grow older (Cozolino, 2008). Many of these differences are
especially important to the processes of attachment and affect regulation that are so central to psychotherapy.
Neural networks relevant to psychotherapy exist throughout the brain—some are evolutionarily primitive;
others have developed more recently. Some are fully functional from birth, while others take decades to
mature. This is why an understanding of both evolution and development is vital in order to capture a fuller
picture of human experience.

The Interpersonal Sculpting of the Social Brain


It is difficult to give children a sense of security unless you have it yourself. If you have it, they catch it from you.
—William Menninger

The theory that ontogeny recapitulates phylogeny refers to the concept that the evolution of the species is re-
created in the gestation and development of each individual. To use MacLean’s terms, we pass through the
reptilian and paleomammalian stages before we develop into fully human beings. Although the theory of
recapitulation is in most ways incorrect (Gould, 1977), some interesting parallels exist between the process of
our evolution and our development.
At birth, the reptilian brain is fully functional, and the paleomammalian brain is primed and ready to be
organized by early experiences. The cortex, on the other hand, continues to slowly grow into the third decade
and matures throughout life. Thus, much of our most important emotional and interpersonal learning occurs
during our early years when our primitive brains are in control. The result is that a great deal of learning takes
place before we have the necessary cortical systems for explicit memory, emotional regulation, or perspective
taking. Consequently, these vital socioemotional learning experiences are organized and controlled by reflexes,
behaviors, and emotions outside of our awareness. Further, our early experiences are distorted by the biases
and limitations of our immature brains. To a great extent, psychotherapy owes its existence to these
problematic artifacts of evolution and development.
The slow development of the cerebral cortex throughout maximizes the influence of experience on the

13
structure and functioning of the brain. That so much of the brain is shaped after birth is both good and bad
news. The good news is that the individual brain is built to survive in a particular environment. Via epigenetic
processes, culture, language, climate, nutrition, and parents shape each of our brains in a unique way. In good
times and with good-enough parents, this early brain building will serve the child well later in life. The bad
news comes into play when factors are not so favorable, such as in times of war or in the case of parental
psychopathology (Benes, Taylor, & Cunningham, 2000). The brain is then sculpted in ways that assist the
child in surviving childhood but that may be maladaptive later in life. It is in these instances that a therapist
attempts to restructure neural architecture in the service of more adaptive behaviors, cognitions, and emotions.
Building the human brain is vastly complex. Rebuilding it is a difficult and fascinating challenge.
A portion of the brain called the anterior cingulate—centrally involved with maternal behavior, nursing,
and play—appears in the evolution of early mammals (MacLean, 1985). Before this, animals had to be
prepared to survive on their own at birth. A good example is newborn sea turtles that hatch from their eggs
high on a beach and make a mad instinctual dash toward the ocean. With the evolution of maternal care,
children are allowed to develop more slowly within a supportive and scaffolding environment. In the course of
evolution, primates have experienced increasingly long periods of maternal dependence. This luxury allows for
the evolution and development of more complex brains as well as an increasing impact of parenting and early
experiences on how the brain is built.
Konrad Lorenz (1991) found that geese imprint (bond to attachment figures) during a limited period of
time soon after birth. If baby geese saw Lorenz first after they hatched, they would follow him as if he were
their mother. Lorenz also found that when these geese reached sexual maturity two years later, they would
become attached to the kinds of geese they had been exposed to during their imprinting period. He even
noted that a baby goose, who had imprinted on him, fell in love with a human girl from the next town when
he reached sexual maturity, and the goose would fly there to see her. These early experiences seemed to be
permanently etched into the brains of Lorenz’s geese.
This principle of imprinting can be seen in humans in the more flexible and complex form of attachment
schema. The early interpersonal environment may be imprinted on the human brain by shaping the child’s
neural networks and establishing the biochemical set points in circuitry dedicated to memory, emotion, safety,
and survival. Later, these structures and processes come to serve as the infrastructure for social and intellectual
skills, affect regulation, and the sense of self.
Prolonged dependence in childhood and the differentiation of the two cerebral hemispheres has allowed
for the development of a neocortex so complex that we have become capable of spoken and written language,
self-consciousness, and the construction of both private and social selves. Although these abilities create
tremendous possibilities, brainpower does have its downside. We are now also capable of becoming anxious
about things that will never happen, depressed by imagined slights, and saddened by potential losses. Our
imaginations can simultaneously create exciting new worlds as well as the fears that prevent us from fully
living in them. It is obvious that despite the evolution of consciousness and rationality, our primitive
emotional brains and their early development continue to exert a great deal of influence over us.

Summary

Although Freud began his career attempting to create a brain-based psychology, the theories and technology
available to him did not allow him to carry out this project. Various ways of thinking about the brain (like
MacLean’s), although limited, provide models that bridge the gap between psychology and neurology.
Evolution’s legacy is a complex brain, vulnerable to a variety of factors that can disrupt the growth and
integration of important neural networks. The field of psychotherapy has emerged because of the brain’s
vulnerability to these developmental and environmental risks. So how can psychotherapists synthesize and
incorporate both the mind and the brain into their work? Chapter 2 presents a model of neural networks, how
they develop, and how we attempt to alter them during treatment. It is from this perspective that we will then
examine the relevance of the nervous system to our work as clinicians.

14
Chapter 2
Building and Rebuilding the Brain:
Psychotherapy and Neuroscience
I know of no more encouraging fact than the unquestionable ability of man to elevate his life by a conscious
endeavor.
—Henry David Thoreau

Although psychotherapy originally emerged from neurology, differences in language and worldview have
limited collaboration between the two fields. While psychotherapists developed a rich metaphoric language of
mind, neurologists built a detailed database of brain-behavior relationships. In the 21st century, neuroscience
is providing us with tools to explore what happens in the brain during early development and later in
psychotherapy. A return to Freud’s project of a biological psychology is finally at hand.
At the heart of the interface of neuroscience and psychotherapy is the fact that human experience is
mediated via two interacting processes. The first is the expression of our evolutionary past via the
organization, development, and functioning of the nervous system—a process resulting in billions of neurons
organizing into neural networks, each with its own timetable and requirements for growth. The second is the
contemporary shaping of our neural architecture within the context of relationships. The human brain is a
social organ of adaptation stimulated to grow through positive and negative interactions with others. The
quality and nature of our interpersonal relationships become encoded within the neural infrastructure of our
brains. It is through this translation of experience into neurobiological structures that love becomes flesh and
nature and nurture become one.
At the heart of psychotherapy is an understanding of the interwoven forces of nature and nurture, what
goes right and wrong in their developmental unfolding, and how to reinstate healthy neural functioning.
When one or more neural networks necessary for optimal functioning remain underdeveloped,
underregulated, or underintegrated with others, we experience the complaints and symptoms for which people
seek therapy. We can now assume that when psychotherapy results in symptom reduction or experiential
change, the brain has, in some way, been altered (Kandel, 1998).
How does psychotherapy change the brain? How is memory stored and how can the quality of experience
change? Before we can address these questions, we have to first get an idea of how the brain is organized and
how it performs some of its many functions. We will discuss the building and rebuilding of neural networks,
the role of enriched environments, and the role stress plays in changing the brain. We will also explore the
central role of the therapeutic relationship in this change process, as well as the importance of the expression
of emotion and the therapeutic use of language.

Neural Networks
A forest of these trees is a spectacle too much for one man to see.
—David Douglas

So far we have used the term neural networks in a general way; I would now like to get a bit more specific.
Neurons are the microscopic processing units that make up all parts of the nervous system. When we talk of
the frontal cortex, amygdala, or hippocampus, we are literally talking about large numbers of individual
neurons organized to perform a set of functions. The neurons within these systems need to be able to organize
and reorganize in such a way as to allow us to learn, remember, and act as we adjust to different situations.
Because each neuron is limited to either firing or not firing, the diverse capabilities of the nervous system
come from the complex interaction of individual neuronal signals.
A simplistic analogy is an old-fashioned billboard consisting of rows and columns of thousands of

15
lightbulbs. Although each individual bulb is limited to being either on or off, the pattern created by these
lights can spell out words, form images, and, through precise timing, create the illusion of movement. Digital
technology allows us to do the same things with millions of individual pixels on a screen. In a similar fashion,
patterns of neural firing come to represent specific information within the brain and throughout the nervous
system.

FIGURE 2.1
The Feedforward Neural Network
A depiction of sixteen neurons in a simple feedforward circuit

16
To accomplish the complexity required for behavior, neurons organize into neural networks. A neural
network can range from just a few neurons in a simple animal to millions of neural interconnections in brains
such as our own. Neural networks encode and organize all of our behaviors from basic reflexes, such as pulling
our hand away from a hot stove, to our ability to simultaneously comprehend the visual, emotional, and
political significance of Picasso’s Guernica. Neural networks can interconnect with multiple other networks,
allowing for interaction, coordination, and integration of functions. Because I refer to neural networks
throughout the chapters to come, it is important to keep a good visual image in our minds as we proceed.

FIGURE 2.2
The Feedforward and Feedback Neural Network
A slightly more complex model in which information is fed backwards and each neuron can communicate with all of its neighbors.

17
Figures 2.1 and 2.2 depict extremely simple neural networks, with each circle representing an individual
neuron. Starting with Figure 2.1, you will notice that the flow of information moves from left to right across
the four columns of neurons. On the left, some of the input neurons are firing in response to some stimulus (1
= firing/0 = nonfiring). In turn, their firing stimulates the activation of some set of neurons within the hidden
layers of processing, which leads to the firing of a set of output neurons, which then results in a particular
experience or behavioral reaction. Figure 2.2 represents a step toward a more accurate model, with
information flowing in both directions and an increased level of interaction among neurons. Each of the
connections will have either an excitatory or inhibitory effect on other neurons. This mosaic of firing patterns,
or the network’s instantiation, will determine which set of output neurons fire. Making things slightly more
complicated, instead of 16 neurons there are millions, each of which can be connected to thousands of others.
Instantiations are sculpted by experience, and they encode all of our abilities, emotions, and experiences
into one or more forms of memory. It is the consistency of these firing patterns that results in organized
patterns of behavior and experience. Once these neural patterns are established, new learning modifies the
relationship of neurons within these networks. At other times, new learning may occur when we shape one
neural network to inhibit the activation of another. When we talk of building and rebuilding the brain,
neurons are our basic building blocks and neural networks are the structures that we are sculpting.
Learning within neural networks occurs as a result of trial and error. Feedforward and feedback
information loops form complex patterns of excitation and inhibition among neurons within the hidden
layers. This process eventually leads to consistent and adaptive output. This is demonstrated in a soon-to-be
toddler who repeatedly tests and refines her balance, leg strength, and coordination with each new attempt to

18
walk. Her brain drives her to keep trying while recording her successes and failures within neural networks
responsible for balance, motor coordination, and visual tracking. In the same way, neural networks learn to
organize behaviors, emotions, thoughts, and sensations. The brain eventually creates a refined set of neural
activations that makes walking seem second nature.
I remember being surprised to find a table of random numbers in an appendix of my college statistics
textbook. At first, I thought this to be a waste of paper, assuming that anyone could generate random
numbers on their own. When I shared my thoughts with the professor, he assured me that much research had
gone into demonstrating that we are incapable of generating random numbers. He said that as hard as we
might try, we cannot avoid generating specific patterns of numbers. This finally makes sense to me based on
neural network organization: We are unable to engage in random actions because our behaviors are guided by
patterns already established through previous learning to which we automatically return. And while not being
able to generate random numbers is of little consequence to us in our day-to-day lives, the tendency to make
the same mistakes again and again is cause for a great deal of human suffering. This tendency to repeat
patterns of thought and behavior is what led the psychoanalyst Wilhelm Reich to say that people tend to
remain sick because they continue to find the same wrong solutions to the problems they hope to change.

Neural Network Growth and Integration


Plasticity then, in the wide sense of the word, means the possession of a structure weak enough to yield to an influence, but strong enough not to
yield all at once.
—William James

The growth and connectivity of neurons is the basic mechanism of all learning and adaptation. Learning can
be reflected in neural changes in many ways, including changes in the connectivity between existing neurons,
the expansion of existing neurons, and the growth of new neurons. All of these changes are expressions of
plasticity, or the ability of the nervous system to change in response to experience. Although the first two
forms of plasticity have been recognized in humans for decades, the birth of new neurons (neurogenesis) was
only recently discovered in regions involved with ongoing learning, such as the hippocampus, the amygdala,
and the frontal and temporal lobes (Eriksson et al., 1998; Gould, Reeves, Graziano, & Gross, 1999; Gould,
Tanapat, Hastings, & Shors, 1999; Gross, 2000).
Existing neurons grow though the expansion and branching of the dendrites that they project to other
neurons in reaction to new experiences and learning (Purves & Voyvodic, 1987). This process is reflected in
the connectivity among neurons in our simple schematic diagrams. Neurons interconnect to form neural
networks, and neural networks, in turn, integrate with one another to perform increasingly complex tasks. For
example, networks that participate in language, emotion, and memory need to become integrated in order for
us to recall and tell an emotionally meaningful story with the appropriate words, correct details, and proper
emotions.
Association areas within the cortex serve the roles of bridging, coordinating, and directing the multiple
neural circuits to which they are connected. Although the actual mechanisms of this integration are not yet
known, they are likely to include some combination of communication between local neuronal circuits and the
interactions among functional brain systems (Trojan & Pokorny, 1999). Changes in the synchrony of
activation of multiple neural networks may also play a role in the coordination of their activity and the
emergence of conscious awareness (Crick, 1994; Konig & Engel, 1995).

Genetic Inheritance and Gene Expression


Evolution consists of the gradual transformation of organisms from one condition of existence to another.
—Ernst Mayr

Now that we are living beyond the nature versus nurture debate, we can acknowledge that the growth and
organization of the brain reflects a complex yet subtle blending of genetic and environmental influences.
Toward this end, it is much more helpful to think of genes in terms of serving both a template and a
transcription function (Kandel, 1998). As templates, genes provide the organization of the uniform structures
of the brain that are generally unaffected by environmental influences, except in cases of mutations caused by
toxins during gestation. These structures and functions, such as the general layout of the nervous system and

19
basic reflexes, are inherited via our DNA and shared by all healthy members of our species. This is the aspect
of genetic inheritance traditionally thought of as nature.
On the other hand, the expression of many genes depends on experiences that trigger their transcription
(Black, 1998). Transcription genetics controls the subtler aspects of the brain’s organization, such as the
specific sculpting of later-developing neural networks and the levels of specific neurotransmitters available to
different brain systems. In fact, the majority of our cortex is shaped after birth in an experience-dependent
fashion through the transcriptional process. Nurture, therefore, influences brain development via the selective
activation of genes (gene expression) that shape the experience-dependent aspects of development.
Experience results in the expression of certain genes that trigger the synthesis of proteins that build neural
structures. Through genetic transcription, existing neurons grow different kinds of receptors, expand their
dendritic structures, and adjust their biochemistry. For example, although identical twins raised in the same
household may have identical genes for schizophrenia, only one may develop the disorder. This is believed to
be the result of the expression of different genes based on the unique interactions between each child and his
or her environment. The transcription function of genes allows for ongoing neural plasticity throughout life
and provides the basis for enriched experiences (like psychotherapy) to be of benefit. In a later chapter I
explore the links between maternal nurturance and the early building of the brain that result in different levels
of learning, emotional regulation, and attachment behavior.

The Role of Enriched Environments


It is always with excitement that I wake up in the morning. . . . It’s my partner.
—Jonas Salk

The brain is not a static organ; it continually changes in response to environmental challenges. Because of this,
the neural architecture of the brain comes to embody the environment that shapes it. You could also think of
our neural architecture as a tangible expression of our learning history. The early research on neural plasticity
began by exploring the impact of different types of environments on brain development. In these studies done
primarily with rats, enriched environments took the form of more diverse, complex, colorful, and stimulating
habitats, while impoverished environments were relatively empty monochromatic enclosures. It was found
that animals raised in enriched environments had more neurons, more dendrite connections among neurons, a
greater number of blood capillaries, and more mitochondria activity (Diamond, Krech, & Rosenweig, 1964;
Kempermann, Kuhn, & Gage, 1997, 1998; Kolb & Whishaw, 1998; Sirevaag & Greenough, 1988). These
findings demonstrate that a brain that is challenged comes to be more complex, active, and robust. Subsequent
research with humans has yielded similar results for individuals with higher education and more intellectually
challenging occupations.
For humans, enriched environments include the kinds of experiences that encourage us to learn new skills
and expand our knowledge. Higher levels of education, practicing skills, and continued engagement in mental
activities all correlate with more neurons and neural connections (Jacobs & Scheibel, 1993; Jacobs, Schall, &
Scheibel, 1993). Higher levels of education and reading ability have also been shown to correlate with a
diminished impact of dementia later in life (Schmand, Smit, Geerlings, & Lindeboom, 1997). Interestingly,
brain regions dedicated to certain skills can actually hijack cells in adjacent neural areas in order to develop
skills like playing an instrument or learning Braille (Elbert, Pantev, Wienbruch, Rockstroh, & Taub, 1995).
There is no doubt that the human brain grows in response to challenge and new learning. Psychotherapy can
be thought of as a specific type of enriched environment that promotes social and emotional development,
neural integration, and processing complexity. The way the brain changes during therapy will depend upon
the neural networks involved in the symptoms and focus of treatment.

Learning and Stress


Every stress leaves an indelible scar, and the organism pays for its survival after a stressful situation by becoming a little older.
—Hans Selye

Mild to moderate stress (MMS) activates neural growth hormones supportive of new learning (Cowan &
Kandel, 2001; Gould, McEwen, Tanapat, Galea, & Fuchs, 1997; Jablonska, Gierdalski, Kossut, & Skangiel-
Kramska, 1999; Myers, Churchill, Muja, & Garraghty, 2000; Pham, Soderstrom, Henriksson, &

20
Mohammed, 1997; Zhu & Waite, 1998). Thus, MMS may be utilized to enlist naturally occurring
neurobiological processes in the service of new learning. Although we typically use the term stress in animal
research, humans also demonstrate arousal in the form of curiosity, enthusiasm, and pleasure. Humans can
also be motivated to learn new skills and take on new challenges in order to relieve discomfort and stress.
These motivational states have all been recognized for their role in successful outcomes from psychotherapy.
Dissociation is a common result of the high levels of stress associated with traumatic experiences.
Characterized by a disconnection among thoughts, behaviors, sensations, and emotions, dissociation
demonstrates that the coordination and integration of these functions is an active neurobiological process.
Because all of these functions are seamlessly and unconsciously interwoven during normal states of awareness,
it is easy to overlook the fact that their integration is a central component of mental health.
The power of MMS to trigger neural plasticity is a key element in the success of psychotherapy or any
situation that requires new learning. As opposed to traumatic experiences, controlled exposure to MMS
during therapy enhances new learning and increases neural integration. As therapists, we intuitively work to
regulate stress and integrate neural networks, a process that is essentially the opposite of the dissociation
observed in reaction to trauma. Healthy functioning requires proper development and functioning of neural
networks organizing conscious awareness, behavior, emotion, and sensation.
As in early development, the repeated exposure to stress in the supportive interpersonal context of
psychotherapy results in the ability to tolerate increasing levels of arousal. This process reflects the building
and integration of cortical circuits and their increasing ability to inhibit and regulate subcortical activation.
Affect regulation, especially the modulation and inhibition of anxiety and fear, allows for continued cortical
processing in the face of strong emotions, allowing for ongoing cognitive flexibility, learning, and neural
integration.
In this process the therapist essentially plays the same role as a parent, providing and modeling the
regulatory functions of the social brain. As affect is repeatedly brought into the therapeutic relationship and
successfully managed, the client gradually internalizes these skills by sculpting the relevant neural structures
necessary for emotional regulation. As in childhood, the repeated cycle of attunement, rupture of the
attunement, and its reestablishment gradually creates an expectation of reconnection (Lachmann & Beebe,
1996). The learned expectation of relief in the future enhances the ability to tolerate more intense affect in the
midst of the stressful moment.
As a therapist, one of my primary goals is to shift my clients’ experience of anxiety from an unconscious
trigger for avoidance to a conscious cue for curiosity and exploration. One of my patients described it
metaphorically as using anxiety as a compass to help guide him to and through his unconscious fears.
Becoming aware of anxiety is then followed with an exploration and eventual understanding of what we are
afraid of and why we are afraid. The next step is to move toward the anxiety with an understanding of its
meaning and significance. In this way, anxiety becomes woven into a conscious narrative with the possibility
of writing a new outcome to our story. This process reflects the integration of cortical linguistic processing
with conditioned subcortical arousal in the service of inhibiting, regulating, and modifying maladaptive
reactions.
As we will see later when discussing the building of the social brain, biological and environmental factors
during childhood can result in long periods of dysregulation. Early deprivation or chronic stress increase the
chances of damage to the brain, deficits in memory and reality testing, and the prolonged utilization of
primitive defenses (Brown, Henning, & Wellman, 2005; Radley et al., 2006; Sapolsky, 1985). With increased
nurturance and support, stress hormone levels decrease; physical comfort and soothing talk with caretakers
helps the brain to integrate experience.

Emotional Tolerance and Affect Regulation


Few things are brought to a successful issue by impetuous desire, but most by calm and prudent forethought.
—Thucydides

Although we usually think of the cortex as a hard drive that stores information, another primary role of the
cortex is inhibition. Take for example the grasping reflex we are all born with. This powerful grip allowed our
ancestors to hold onto their mothers as they climbed trees and moved over land. During the early months of
life this grasping reflex is soon inhibited by descending cortical circuitry. The inhibition of this and other
reflexes allows for a cortical takeover of these functions during development. So we sacrifice the grasping
reflex for the finger dexterity necessary to manipulate digits, write, and use tools. Later in life, if we have the

21
misfortune of succumbing to dementia, this and other early reflexes begin to reappear as our cortex gradually
loses its inhibitory ability. In a similar fashion, our prefrontal cortex is shaped by experience to inhibit and
control subcortical functional activation, which eventually results in our ability to regulate our emotions,
impulses, and behaviors. For example, early attachment relationships establish the experiences that shape
these neural networks and allow us to regulate our emotional experience.
Assistance with experiencing increasing levels of positive and negative affect is a vital component of both
parenting and psychotherapy. The gradually increasing tolerance for stress builds our brains, expands neural
organization of emotional and cognitive integration, and creates networks of descending control to help
inhibit and regulate affect (Schore, 1994). Emerging from childhood with an ability to experience a range of
emotions and tolerate stress serves as a means of both brain growth and ongoing positive adaptation.
During our first few years, we have repeated experiences of going from a comfortable, regulated state to a
state of dysregulation. We become frightened, cold, wet, and hungry, and show our displeasure with facial
expressions, bodily postures, vocalization, and crying. In the presence of good-enough parenting, our signals
are attended to, the source of our displeasure diagnosed, and we are helped back into a regulated state. Across
thousands of these temporal-emotional experiences, we go from regulation to dysregulation to reregulation.
These experiences shape secure attachment and the expectation of positive outcomes. The summation of these
experiences, stored throughout our nervous system, becomes the sensory-motor-emotional background of our
experience.
In the absence of adequate assistance in regulating affect or making sense of emotions, the brain organizes
a variety of defensive coping strategies. These defenses vary in the degree to which they distort reality in order
to achieve their goal of reducing anxiety. This distortion is accomplished in circuits of unconscious memory
that control anxiety and fear (Critchley et al., 2000). The neural connections that result in defenses shape our
lives by selecting what we approach and what we avoid, what our attention is drawn to, and the assumptions
we use to organize our experiences. Our cortex then provides us with rationalizations and beliefs about our
behaviors that help keep our coping strategies and defenses in place, possibly for a lifetime. These neural and
psychic structures can lead to either psychological and physical health, or illness and disability.

Psychopathology and Neural Network Integration


In a structure as complex as the human brain a multitude of things can go wrong. The wonder is that for most people the brain functions
effectively.
—Seymour Kety

If everything we experience is represented by instantiations within neural networks, then by definition,


psychopathology of all kinds—from the mildest neurotic symptoms to the most severe psychosis—must also
be represented within and among neural networks. In line with this theory, psychopathology would be a
reflection of suboptimal development, integration, and coordination of neural networks. Patterns of
dysregulation of brain activation found in disorders such as depression and obsessive-compulsive disorder
support the theory of a brain-based explanation for the symptoms of psychopathology.
Difficulties in early caretaking, genetic and biological vulnerabilities, or trauma at any time during life can
result in the lack of integration among networks. Unresolved trauma can cause ongoing information
processing deficits that disrupt integrated neural processing. For example, dissociative symptoms following
trauma—reflecting the disconnection among networks of behavior, emotion, sensation, and cognition—
predict the later development of posttraumatic stress disorder (Koopman, Classen, & Spiegel, 1994;
McFarlane & Yehuda, 1996). Children victimized by psychological, physical, and sexual abuse have a greater
probability of demonstrating electrophysiological abnormalities in executive regions of the brain vital to neural
network integration (Ito et al., 1993; Teicher et al., 1997).
In general, psychological integration suggests that the conscious and cognitive functions of the executive
brain have access to information across networks of sensation, behavior, and emotion. A primary focus of
neural integration in traditional talk psychotherapy is between networks of affect and cognition. Dissociation
between the two occurs when high levels of stress inhibit or disrupt the brain’s integrative abilities among the
left and right cerebral hemispheres as well as between the cortex and limbic regions. The integration of the
left and right hemispheres can be disrupted while the circuits of the reptilian and paleomammalian brains can
be unlinked from the conscious neomammalian cortex. As we will see in later chapters, trauma can result in
disruption of neural integration associated with the range of symptoms seen in posttraumatic stress disorder.
This unlinking may not be an evolutionary accident. As valuable as language can be for humans, evolution

22
appears to have selected for the shutdown of language and a decrease in cognitive processing when confronted
with a threat. The resulting disruption of information processing may be the most common cause of neural
network dissociation. Cortical networks responsible for memory, language, and executive control (in its many
forms) become inhibited and underperform during times of overwhelming stress. The very way that the brain
has evolved to successfully cope with immediate threat appears to have created a vulnerability to longer-term
psychological distress: enter psychotherapy.
Within this model, psychotherapy is a means of creating or restoring coordination among various neural
networks. Research has demonstrated that successful psychotherapy correlates with changes in activation in
areas of the brain hypothesized to be involved in disorders such as obsessive-compulsive disorder and
depression (Baxter et al., 1992; Brody, Saxena, Mandelkern, et al., 2001; Brody et al., 1998; Schwartz,
Stoessel, Baxter, Martin, & Phelps, 1996). The return to normal levels of activation and homeostatic balance
results in reestablishing positive reciprocal regulation among the participating neural networks.

Psychotherapy and Neural Network Integration


The only thing they (neural connections) can do . . . is to deepen old paths or to make new ones.
—William James

A basic assumption of both neuroscience and psychotherapy is that optimal functioning depends upon
optimal levels of growth, integration, and complexity. On a neurological level, this equates to the integration
and communication of neural networks dedicated to emotion, cognition, sensation, and behavior and a proper
balance between excitation and inhibition. On an experiential level, integration is the ability to feel, think, and
live life, love, and work with a minimum of defensiveness. Growth and integration are optimized by a positive
early environment, including stage-appropriate challenges, support, and parents who are capable and willing
to put feelings into words. These factors lead to positive affect regulation, biological homeostasis, and a quiet
internal milieu allowing for the consolidation of the experience of subjectivity and a positive sense of self.
From the perspective of neuroscience, psychotherapy can be understood as a specific kind of enriched
environment designed to enhance plasticity, the growth of neurons, and the integration of neural networks.
The therapeutic environment is individually tailored to fit the symptoms and needs of each client so as to
change the underlying neural systems that promote them. If this is true, then all forms of therapy, regardless
of theoretical orientation, will be successful to the degree to which they foster appropriate neuroplasticity.
Further, the clinical and neuroscientific research both support that neural plasticity, growth, and integration in
psychotherapy are enhanced by:

1. The establishment of a safe and trusting relationship


2. Mild to moderate levels of stress
3. The activation of both emotion and cognition
4. The coconstruction of new and more adaptive personal narratives

Although psychotherapists do not generally think in neuroscientific terms, stimulating neuroplasticity and
neural integration is essentially what we do. We provide information to clients about our understanding of
their difficulties in the form of psychoeducation, interpretations, or reality testing. We encourage clients to
engage in behaviors, express feelings, and become aware of unconscious aspects of themselves. We dare them
to take risks. We guide them back and forth between thoughts and feelings, trying to help them establish new
connections between the two. We help clients alter their description of themselves and the world,
incorporating new awareness and encouraging better decision making. With successful treatment, the
methods being used are internalized so that clients can gain independence from therapy, and we do this all in
the context of a warm, supportive, committed, and consistent relationship. The same factors are at play across
psychodynamic, systems, and cognitive-behavioral approaches to treatment. They also apply to parenting, the
classroom, and the workplace because brains are brains.
The broad context in which these processes can successfully occur is one of increasing levels of affect
tolerance and regulation and the development of integrative narratives that emerge from the client-therapist
relationship. In the context of empathic attunement within a safe and structured environment, clients are
encouraged to tolerate the anxiety of feared experiences, memories, and thoughts. In this process, neural
networks that are normally inhibited become activated and available for inclusion in conscious processing
(Cozolino, 2015; Siegel, 1995). Interpretations in psychodynamic therapy, exposure in behavioral therapies, or

23
experiments in differentiation from a systems perspective all focus on this goal. Through the activation of
multiple cognitive and emotional networks, previously dissociated functions are integrated and gradually
brought under the control of cortical executive functions. Narratives coconstructed with therapists provide a
new template for thoughts, behaviors, and ongoing integration.

Pathways of Integration
It is the harmony of the diverse parts, their symmetry, their happy balance; in a word it is all that introduces order, all that gives unity.
—Henri Poincaré

Given that information flows simultaneously in multiple directions through many neural networks, optimal
neural integration likely involves maximizing the flow and flexibility of energy through the entire nervous
system (Pribram, 1991). Using this model, psychopathology can be caused by difficulties not just in a specific
region of the brain, but through the brain and body (Mayberg, 1997; Mayberg et al., 1999). Numerous
processing networks combine affect, sensation, behavior, and conscious awareness into an integrated,
functional, and balanced whole—the neural substrate for what Freud called the ego. The ego is essentially
shorthand for how the organization of the self comes to be expressed in dimensions such as personality, affect
regulation, coping styles, and self-image.
The primary directions of information flow relevant to psychotherapy are top-down (cortical to subcortical
and back again) and left-right (across the two halves of the cortex). Keep in mind that these information loops
need to communicate with each other as well as with many other processing systems. Top-down or bottom-
up integration would include MacLean’s linkup among the three levels of the triune brain and the unification
of the body, emotion, and conscious awareness. This is called top-down because these circuits form loops that
go from the top of our head down into the depths of the brain and back up again. Top-down integration
includes the ability of the cortex to process, inhibit, and organize the reflexes, impulses, and emotions
generated by the brain stem and limbic system (Alexander et al., 1986; Cummings, 1993). Frontal lobe
disorders often result in a disinhibition of impulses and movements normally under its control, including
obsessive-compulsive and attention deficit disorders. Within this category, I include what has been referred to
as dorsal-ventral integration, which connects cortical with limbic processing (Panksepp, 1998; Tucker, Luu, &
Pribram, 1995).
Left-right or right-left integration involves abilities that require the input of both the left and right
cerebral cortex and lateralized limbic regions for optimal functioning. For example, adequate language
production requires an integration of the grammatical functions of the left and the emotional functions of the
right. Left-right integration allows us to put feelings into words, consider feelings in conscious awareness, and
balance the positive and negative affective biases of the left and right hemispheres (Silberman & Weingartner,
1986). A balance between the left and right prefrontal cortices is also necessary for the proper balance of affect
and emotion. Alexithymia (the inability to put words to feelings) and somatization disorder (the conversion of
emotional conflicts into bodily illness) may reflect left-right dissociation (Hoppe & Bogen, 1977). There is
also evidence that depression and mania correlate with dysregulation of the balance of activation between the
left and right prefrontal cortices (Baxter et al., 1985; Field, Healy, Goldstein, Perry, & Bendell, 1988).
The right hemisphere is more highly connected with the body and the more primitive and emotional
aspects of functioning. The left hemisphere is more closely identified with cortical functioning, whereas the
right is more densely connected with limbic and brain stem functions (Shapiro, Jamner, & Spence, 1997). For
example, states of stress, anxiety, and fear result in increased activation in the right cortex and subcortical
structures (Rauch et al., 1996; Wittling, 1997). This bias is also relevant to the organization of social-
emotional attachment patterns, transference, and affect regulation (Minagawa-Kawai et al., 2008). Much of
the integration of top-down and left-right systems is mediated through the frontal-parietal executive system.
Due to the interconnectivity between left-right and top-down neural networks, examining integration
from either the vertical or horizontal dimension alone is overly simplistic. Studies of metabolic activity in
specific areas of the brain in pathological states reveal differences in both cortical and subcortical structures on
both sides of the brain. This research suggests that restoring neural integration requires the simultaneous
reregulation of networks on both vertical and horizontal planes. There are other dimensions of neural
integration we will get to later. For now, the vertical and horizontal planes are enough to keep in mind. Keep
in mind that although we are discussing brain functioning from the perspective of neural networks, an equally
meaningful discussion could focus on the impact of pharmacological agents on the modulation and
homeostatic balance of these networks (Coplan & Lydiard, 1998). This perspective helps us to understand

24
why both psychotherapy and medication can result in shifts of neural activity and symptom reduction and why
they may work better together than either one alone (Andreasen, 2001).
Neural network integration can also be accomplished through the activation of conscious language
production (top and left) with more primitive, emotional, and unconscious processes (down and right) that
have been dissociated due to stress or trauma. Depending on their theoretical orientation, therapists facilitate
the process of network integration by supplying all kinds of challenges. An analyst may use interpretations to
enhance awareness of inhibited, repressed, or dissociated thoughts and emotions. A cognitive-behavioral
therapist will expose a client to a feared stimulus combined with relaxation training, allowing normally
inhibited cortical circuitry to integrate with the subcortical circuitry that controls fear. Research across all
forms of psychotherapy supports the hypothesis that positive outcomes are related to utilizing both support
and challenge in the combined engagement of thought and affect (Orlinsky & Howard, 1986). Both the
quality of the interpersonal connection and creating the proper learning environment appear essential.

Psychotherapy and Parenting


Parents are like shuttles on a loom. They join the threads of the past with threads of the future and leave their own bright patterns as they go.
—Fred Rogers

We have talked a little about the parallels between positive parenting and successful psychotherapy; these
similarities reflect the commonality of the conditions required for building and rebuilding the brain. Mutual
eye gaze and escalating positive emotional interactions between parent and child stimulate the growth and
organization of the brain. In the future, we may discover scientific evidence that the interpersonal experience
of psychotherapy impacts the neurobiological environment of the brain in ways that stimulate neural plasticity
and neurogenesis. Although the various schools of therapy tend to accentuate their differences, the therapeutic
relationship itself may be the most powerful curative agent.
The warmth, acceptance, and unconditional positive regard demonstrated in Carl Rogers’s work embodies
the broad interpersonal environment for the initial growth of the brain and continued development later in life
(Rogers, 1942). Having spent a brief period of time with Dr. Rogers as a student, I can attest to the power of
his interpersonal style and therapeutic technique. I am sure he left many, including myself, with the fantasy of
being available for adoption.
Primary goals of parenting include providing a child with the capacity for self-soothing and the ability to
form positive relationships. This allows the child to face the challenges of life and benefit from healing life
experiences. The successful mastery of challenges throughout life leads to taking on even more complex
challenges that will promote increasingly high levels of neural network development and integration. When
internal or external factors prevent an individual from approaching challenging or stressful situations, the
neural systems will tend to remain underdeveloped or unintegrated.
In a review of hundreds of studies examining the outcome of psychotherapy, Orlinsky and Howard (1986)
looked for those factors that seemed to relate to success. They found that the quality of the emotional
connection between patient and therapist was far more important than the therapist’s theoretical orientation.
Patients who are motivated to change and who are also able to work collaboratively with their therapists do
better. Therapists’ professional experience was also positively related to success, as were the use of
interpretation, a focus on transference, and the expression of emotion. The continual involvement of both
cognitive and emotional processing during treatment seems essential for positive change.
Psychotherapy, like parenting, is neither mechanical nor generic. Each therapist-client pair creates a
unique relationship resulting in a particular course of treatment and a specific outcome. The importance of the
unconscious processes of both parent and therapist is highlighted by their active participation in the
coconstruction of new narratives of their children and clients. As we will see in research on attachment, each
parent’s unconscious plays a role in the creation of the child’s brain, just as the therapist’s unconscious
contributes to the therapeutic relationship and the outcome of therapy. This underscores the importance of
proper training and adequate personal therapy for therapists, as they will be putting their imprint on the
hearts, minds, and brains of their clients.

Summary

25
In this chapter, we have explored some initial concepts in the integration of psychotherapy and neuroscience
based on common principles within both fields. We have equated psychological health with optimal neural
network growth and integration. Both the brain and the self are built in a stepwise manner by experience. The
nervous system is made up of millions of neurons while human experience is constructed within countless
moments of learning. The psychological difficulties for which patients seek psychotherapy are a function of
inadequate growth and integration within and between these networks. The aspects of development that
foster positive brain development and those in therapy that promote positive change are emotional
attunement, affect regulation, and the coconstruction of narratives.
In Chapter 3, we turn our attention to major models of psychotherapy in use today. By examining these
theories and techniques, we will see how therapeutic strategies and principles have been shaped by underlying
principles related to the growth and integration of neural networks. It is my belief that the development of
psychotherapy has always been implicitly guided by the principles of neuroscience. All forms of therapy are
successful to the degree to which they have found a way to tap into processes that build and modify neural
structures within the brain.

26
Chapter 3
Neural Integration in Different
Models of Psychotherapy
The techniques of behavior therapy and psychotherapy have relied on the principles of brain plasticity,
generally without realizing it, for nearly one hundred years.
—Nancy Andreasen

Like other scientific discoveries, psychotherapy developed from a combination of trial-and-error learning,
intuition, and just plain luck. Each school of psychotherapy offers an explanation of mental health and illness
as well as why its strategies and techniques are effective. Fortunately, the effectiveness of an intervention does
not depend on the accuracy of the theory used to support it. For example, there was a time when
psychoanalysts attributed the success of electroshock therapy to the need of a depressed person to be punished.
The treatment worked and continues to work for a certain group of clients despite our lack of a solid
understanding of its mechanisms of action.
Although each approach to psychotherapy can be treated as absolute truth by its disciples, all modes of
therapy are actually heuristics. Heuristics are interpretations of experience or ways of understanding
phenomena. The value of a heuristic lies in its ability to organize, explain, and predict what we observe. You
probably noticed that each form of therapy organized and explains the same behaviors, thoughts, and
symptoms in somewhat different ways. Neuroscience is another heuristic, one that we are now using to
explain the mechanisms of action of psychotherapy, in other words, how and why therapy works.
It is my belief that neuroscience is a helpful heuristic that will lead us to a fuller understanding of the
psychotherapy process and may also serve as a rational means of selecting, combining, and evaluating
treatment modalities. In this chapter, we examine, in broad strokes, some of the primary approaches to
psychotherapy. These overviews are presented in order to provide a context in which to understand and
organize the neuroscientific concepts in the coming chapters. In taking a sample of general theoretical
approaches to psychotherapy, we will look for common elements among them and how these elements may
relate to neural network development and change. Remember, from the perspective of neuroscience,
psychotherapists are in the brain-rebuilding business.

Psychoanalytic and Psychodynamic Therapies


Being entirely honest with oneself is a good exercise.
—Sigmund Freud

Freud’s psychoanalysis, the original form of psychodynamic therapy, has spun off countless variants in its
century-long existence. Ego psychology, self-psychology, and schools of thought connected to names such as
Klein, Kernberg, and Kohut have all attracted considerable followings. Despite their differences,
psychodynamic forms of therapy share theoretical assumptions such as the existence of the unconscious, the
power of early childhood experiences, and the use of defenses that distort reality in order to reduce anxiety and
enhance coping.
The exploration of the unconscious and its connection to our evolutionary past may be Freud’s greatest
legacy. He remained true to Charcot by exploring the multiple levels of human awareness, and he designed
many techniques to bring the unconscious into conscious awareness. The power of trauma, especially during
childhood, and its ability to shape the organization of the mind were central to his work. Freud theorized that
early attachment and relational difficulties, neglect, or trauma result in developmental arrests or fixations that
delay or derail the adult’s potential to love and work. From the standpoint of neurobiology, most of Freud’s
work addressed the discontinuities and dissociations between networks of conscious and unconscious

27
processing. Freud focused on the role of overwhelming emotion as the cause of unintegrated neural
processing.
Freud’s psychic self contains the primitive drives (id), the demands of civilization to conform for the
benefit of the group (superego), and those parts of the self that attempt to negotiate the naturally occurring
conflicts between the two (ego). In its role as a diplomat in the fight between id and superego, the ego utilizes
many elaborate defenses to cope with reality. Ego strength, our ability to navigate reality with a minimum of
defensiveness, reflects the integration of neural networks of emotion and thought, and the development of
mature defenses. The more primitive or immature the defense mechanism, the more reality is distorted and
the more functional impairment occurs. Sublimation, for example, enables us to convert unacceptable
impulses into constructive and prosocial goals.
Mature defenses, like sublimation or humor, allow us to assuage strong feelings, keep in contact with
others, and remain attuned to a shared social reality. Less mature defenses, such as denial and dissociation,
result in greater distortion of reality and difficulties in both work and relationships. Defenses are often
invisible to their owners because they are organized very early in life, stored in hidden layers of neural
processing that are inaccessible to conscious awareness. What Freud called defenses can be seen as ways in
which neural networks have adapted to cope with emotional stress. People seek treatment when their defense
mechanisms cannot adequately cope with strong negative emotions, when symptoms become intolerable, or
when they are unsatisfied with their relationships.
Despite a conscious awareness that something may be wrong, the hidden layers of neural processing
continue to organize the world based on the prior experiences that shaped them. As we will see in later
chapters, the neural circuitry involved with fear has a tenacious memory and can invisibly influence conscious
awareness for a lifetime. Part of psychodynamic therapy is an exploration and uncovering of this unconscious
organization of experience. Freud’s projective hypothesis described the process by which our brains create and
organize the world around us. As the clarity of a situation decreases, the brain naturally generates its own
structure from memory and projects it onto the world. The way we organize and understand ambiguous
stimuli gives us clues about the architecture of the hidden layers of neural processing (how our unconscious
organizes the world). From the projective hypothesis came the invention of projective tests such as
Rorschach’s inkblots, free association, and an emphasis on the importance of dreams as the “royal road to the
unconscious.”
As part of the projective hypothesis, psychodynamic therapists often provide minimal information about
themselves, allowing the client to project onto them implicit unconscious memories from past relationships.
This form of projection, transference, results in the client placing expectations and emotions from earlier
relationships on the therapist, which allows the client’s implicit beliefs to be experienced and worked through
firsthand. It is through this transference that early relationships for which we have no conscious recollection
are brought fully into therapy. Freud felt that the evocation and resolution of the transference was a core
component of a successful analysis. In Freud’s words, only transference renders “the invaluable service of
making the patient’s buried and forgotten love emotions actual and manifest” (1975, p. 115).
Resistance represents aspects of implicit memory presented by the client that the therapist must decipher.
Early experiences of rejection, criticism, or neglect from parents result in shame, which can evolve into a
child’s negative self-image. The resultant self-criticism (superego) manifests in disrespect for anyone who
shows the child love or respect. An example of this is expressed in the Groucho Marx line, “I’d never join a
club that would have me as a member.” In therapy, this may manifest as a strong distrust of the therapist’s
intentions or his or her ability to be of help.
Interpretations are one of the psychodynamic therapist’s most important tools. Sometimes called the
therapist’s scalpel, interpretations attempt to make the unconscious conscious. Based on observations of all
levels of the client’s behavior, the therapist attempts to bring the processing of the hidden layers to the client’s
attention. Repeated and skillful attention to unconscious material via interpretations, confrontations, and
clarifications results in a gradually expanding awareness of unconscious processes and the integration of
dissociated top-down and right-left processing networks.
Accurate, successful interpretations are sometimes accompanied by feelings of disorganization, anger, or
depression. This is because when defenses are made conscious and when they are exposed for what they are,
they lose their effectiveness, leading to a disinhibition of the emotions that they have been successfully
defending against. In other words, the networks containing the negative emotions become disinhibited and
activated. For example, if intellectualization is being used to avoid the shame and depression related to early
criticism, recognition of the defense will bring these feelings and related memories into awareness, forcing
people to experience those painful emotions.
Emotions play a central role in the success of psychodynamic therapies. The neural networks that organize

28
emotions are often shaped to guide us away from thoughts and feelings for which we were punished or
abandoned. Unconscious anxiety signals continue to shape our behavior, leading us to remain on tried-and-
true paths and avoid situations that trigger our unremembered past. An emphasis on the evocation of emotion
and cognition is an important contribution of psychoanalysis and reflects fundamental underlying
neurobiological processes of health and illness.
Across psychodynamic forms of therapy, conscious awareness is expanded, emotions are explored, and the
expression of repressed or inhibited emotions is encouraged. Feelings, thoughts, and behaviors are repeatedly
juxtaposed, combined, and recombined in the process of working through. The assumptions and narratives
from the past are edited based on new information, and those about the present and future are reevaluated.
The overall goal is combining emotion with conscious awareness and rewriting the story of the self. These
processes, when successful, enhance the growth, integration, and flexibility of neural networks and human
experience.

Rogerian or Client-Centered Therapy


The curious paradox is that when I accept myself just as I am, then I can change.
—Carl Rogers

Against the dominant background of psychoanalysis, Carl Rogers (1942) emerged with a form of therapy he
referred to as “client-centered.” In stark contrast to a theory-based analysis of the client, Rogers emphasized
creating a relationship that maximized the individual’s opportunity for self-discovery. Rogers’s approach
gained rapid acceptance in the nonmedical community, and by the 1960s client-centered therapy became the
dominant form of counseling (Gilliland & James, 1998).
When different approaches to therapy are compared for effectiveness, the general agreement is that the
perceived quality of the client-therapist relationship has the highest correlation with reported treatment
success. Some have gone as far as saying that the curative element is the therapeutic relationship itself, rather
than any specific techniques. Rogers would agree, for he believed that the curative aspects of therapy were the
therapist’s warmth, acceptance, genuineness, and unconditional positive regard for the client. His emphasis on
interpersonal congruence, emotional resonance, and empathic attunement foreshadowed the later emergence
of object relations and intersubjective approaches to psychotherapy (Kohut, 1984; Stolorow & Atwood, 1979).
Over the last century, the therapist attributes suggested by Rogers and what we have come to think of as
the best possible attitudes for optimal parenting have become essentially identical. Rogerian principles lead to
a minimized need for defensiveness and shame while maximizing expressiveness, exploration, and risk taking.
Rogers was likely describing the best interpersonal environment for brain growth during development and
neural plasticity in psychotherapy when he stated that client-centered therapy “aims directly toward the
greater independence and integration of the individual rather than hoping that such results will accrue if the
counselor assists in solving the problem. The individual and not the problem is the focus. The aim is not to
solve one particular problem, but to assist the individual to grow, so that he can cope with the present problem
and later problems in a better-integrated fashion” (1942, p. 28).
During my training in client-centered therapy, I was struck by the power of Rogers’s approach. I found it
immensely difficult to maintain his supportive stance, and I often struggled to keep myself from directing my
clients, giving advice, or pushing them to change. To my astonishment, I found that providing clients with a
supportive relationship led to insights on their part that mirrored the interpretations I struggled to suppress.
Clients often expressed a mixture of sadness and appreciation when they realized how much they longed to be
listened to without fear of judgment and shame.
What might be going on in the brain of someone in client-centered therapy? In the Rogerian interpersonal
context, a client would most likely experience the widest range of emotions within the ego scaffolding of an
empathic other. The activation of neural networks of emotion makes feelings and emotional memories
available for reorganization. Rogers’s nondirective method activates clients’ executive networks and their self-
reflective abilities. Supportive rephrasing and clarification of what clients say may also enhance executive
functioning. This simultaneous activation of cognition and emotion, enhanced perspective, and the emotional
regulation offered by the relationship may provide an optimal environment for neural change. Clients,
scaffolded by the therapist’s support and stimulated by his or her words, can then work to rewrite their stories.
We know that social interactions early in life result in the stimulation of both neurotransmitters and neural
growth hormones that participate in the active building of the brain. By re-creating a positive parenting
relationship, it is likely that the empathic connectedness promoted by Rogers actually stimulates biochemical

29
changes in the brain that are capable of enhancing new learning. For example, studies with birds have
demonstrated that their ability to learn songs is enhanced when exposed to live singing birds versus tape
recordings of the same songs (Baptista & Petrinovich, 1986). Other birds are actually unable to learn from
tape recordings, showing that they need positive social interactions and nurturance in order to learn (Eales,
1985). We will see later how maternal contact and nurturance in rats also protect the brain from the damaging
effects of stress (Meaney, Aitken, Viau, Sharma, & Sarrieau, 1989; Plotsky & Meaney, 1993).
Studies such as these demonstrate that social relationships have the power to stimulate the neural plasticity
required for new learning. The interpersonal and emotional aspects of the therapeutic relationship, called a
nonspecific factor in the psychotherapy outcome literature, may be the primary mechanism of therapeutic
action. As we will see in a later chapter, these nonspecific factors are, in fact, quite specific, as early maternal
care has been linked to increased neural plasticity, emotional regulation, and attachment behaviors. In other
words, those who are nurtured best survive best within a positive and safe environment. Unfortunately, the
social isolation created by certain psychological defenses reinforces the rigidity of neural organization as the
client avoids the interpersonal contexts required to promote healing. In these instances, the therapeutic
relationship may serve as a bridge to once again connect with others.

Cognitive Therapies
It’s not what happens to you, but how you react to it that matters.
—Epictetus

Cognitive therapies highlight the centrality of a person’s thoughts, appraisals, and beliefs in guiding his or her
feelings and actions. They emphasize that negative thoughts, skewed appraisals, and erroneous beliefs create
psychological problems. Cognitive therapy focuses on the identification and modification of dysfunctional
thoughts with the ultimate goal of improved affect regulation (Beck, Rush, Shaw, & Emery, 1979; Ellis,
1962). The primary targets of cognitive-behavioral therapy (CBT) have been depression, anxiety, obsessive-
compulsive disorder, phobias, and panic disorders.
Depressed patients tend to evaluate their world in absolute terms, take details out of context, and
experience neutral comments and events as negative. Common depressive thoughts include the expectation of
failure despite many past successes and thoughts that one is alone despite being surrounded by friends and
family. In cognitive therapy, the patient is educated about these common distortions and encouraged to
engage in reality testing and self-talk designed to counteract negative reflexive statements.
In anxiety disorders, fear comes to organize and control patients’ lives. High levels of anxiety inhibit and
distort rational cognitive processing. Cognitive interventions with these clients often include educating them
about the physiological symptoms of anxiety such as a racing heart, shortness of breath, and sweaty palms.
These clients are taught that feelings of dread are secondary to autonomic symptoms and should not be taken
as seriously as they feel. A focus on understanding normal biological processes usually redirects the client away
from catastrophic attributions that serve to increase anxiety.
With clients suffering from phobias or posttraumatic stress disorder, psychoeducation is combined with
exposure and response prevention, during which the client faces the feared stimulus (e.g., venturing outside or
thinking about a negative event) without being allowed to retreat to the safety of home or a state of denial.
Exposure is gradually paired with relaxation training used to aid in the downregulation of physiological
arousal. This process combines increased cortical processing (thought) with subcortical activation (emotion) to
allow for integration with cortical circuitry in order to permit habituation, inhibition, and eventual extinction
via descending cortical networks.
How does this translate into what is going on in the brain during cognitive therapy? Research has
demonstrated that disorders of anxiety and depression correlate with changes in metabolic balance among
different brain regions. For example, symptoms of depression correlate with activation imbalance within the
prefrontal cortex—lower levels of activation in the left and higher levels in the right (Baxter et al., 1985; Field
et al., 1988). This supports the hypothesis that mental health correlates with the proper balance of networks.
Symptoms of obsessive-compulsive disorder correlate with changes in activation in the medial (middle)
portions of the frontal cortex and a subcortical structure called the caudate nucleus (Rauch et al., 1994).
Posttraumatic flashbacks and states of high arousal correlate with higher levels of activation in right-sided
limbic and medial frontal structures. Importantly, high arousal also correlates with decreased metabolism in
the expressive language centers of the left hemisphere (Rauch et al., 1996).
Of all the different types of therapy, specific links have been found between successful CBT and changes

30
in brain functioning. As described in Chapter 2, changes in brain functioning and symptomatology in both
obsessive-compulsive disorder and depression have been found after successful psychotherapy (Baxter et al.,
1992; Brody, Saxena, Mandelkern, et al., 2001; Brody et al., 1998; Schwartz et al., 1996). These findings
strongly suggest that therapists can utilize cognition to alter the relationships among neural networks in a way
that impacts their balance of activation and inhibition. In striving to activate cortical processing through
conscious control of thoughts and feelings, these therapies enhance left cortical processing, inhibiting and
regulating right-hemispheric balance and subcortical activation. The reestablishment of hemispheric and top-
down regulation allows for increases in positive attitudes and a sense of safety that counteract the depressing
and frightening effects of right-hemisphere and subcortical dominance (Ochsner & Gross, 2008).
Although CBT is carried out in an interpersonal context of collaboration and support, it places far less
emphasis on the therapeutic relationship than Rogerian and psychodynamic approaches do. The inherent
wisdom of this approach with depressed and anxious patients lies in the fact that disorders of affect need
activation of cortical executive structures. Given that emotions are contagious, a deeper emotional connection
might result in the therapist attuning to dysregulated states and sharing in the patient’s depressed, anxious,
and panicky feelings. While emotional attunement with these feelings is helpful, it has been my experience
that after the working relationship is established, challenging thoughts and encouraging new behaviors can
often be far more beneficial to the therapeutic process than empathy alone. The structured aspect of CBT may
protect both therapist and patient from the power of negative affect.

Systemic Family Therapy


We must not allow other people’s limited perceptions to define us.
—Virginia Satir

There is increasing evidence that neural networks throughout the brain are stimulated to grow and organize
by interaction with the social environment. Early relationships become encoded in networks of sensory,
motor, and emotional learning to form what dynamic therapists call inner objects. These inner objects have the
power to soothe, arouse, and dysregulate, depending on the quality of our attachment experiences with
significant others. These unconscious memories organize our inner worlds when we are with others and when
we are alone. Thus, we constantly experience ourselves in the context of others.
This is one reason systems therapists question the validity of diagnosing and treating people in isolation.
They believe that in our day-to-day experience we simultaneously exist in two realities: our present families
and our multigenerational family histories. This perspective is especially relevant when working with children
who have yet to form clear ego boundaries between themselves and their family. Some adult patients who have
not successfully individuated also demonstrate unclear boundaries between their own thoughts and feelings
and those of family members. Regardless of age, however, the basic principles are the same.
Murray Bowen, a prime contributor to systems thinking, presented a model that is compatible with an
exploration of the underlying neuroscience of psychotherapy. His perspective is based on the recognition that
a family provides both emotional regulation and a platform for differentiation. He defines differentiation as
the development of autonomy—a recognition of the balance between the needs of self and others.
Differentiation involves the regulation of anxiety and a balanced integration of affect and cognition. Bowen
would say that anxiety is the enemy of differentiation. That is, the more frightened people are, the more likely
they are to dissociate and the more dependent and primitive they become in their interaction with others
(Bowen, 1978).
When this regression occurs, family members try—consciously and unconsciously—to shape the family in
a manner that reduces their own anxiety. The alcoholic needs the problem to go unmentioned while the
family needs to put on a good front for the outside world. Dysfunctional family patterns such as this one
sacrifice the growth and well-being of one or more members, most often the children, to reduce the overall
level of anxiety in the family. The cognitive, emotional, and social world of an alcoholic family is shaped by
the avoidance of feelings, thoughts, and activities that expose their shameful secret to conscious awareness and
the outside world. The development of the children becomes distorted by the adaptations necessary for their
survival within the pathological system. Unfortunately, the roles and rules of the family designed to decrease
anxiety maintain the pathologies of some and create new pathologies in others.
Over time, the dysfunction becomes embedded in the personality and neural architecture of everyone in
the family, and they collude to maintain the system because they now need to maintain the status quo in order
to feel safe. These experiences become embedded in their neural architecture and are carried forward into

31
adult relationships. As a result, many of us re-create the dysfunction from our family of origin in our choice of
partners and the families that we create as adults. Each family’s problems are determined by the
multigenerational, unconscious shaping of both neural structure and behavior. The functioning of brains and
family dynamics reflects how they have been organized. The dysfunctional brain, like the dysfunctional family,
is shaped by the avoidance of thoughts and feelings, resulting in the dissociation of neural systems of affect,
cognition, sensation, and behavior, as well as a lack of human differentiation.
As in other forms of psychotherapy, the goal of systems therapy is to integrate and balance the various
cortical and subcortical, left- and right-hemisphere processing networks. This process requires a decrease in
anxiety from high to low or moderate levels. High levels of affect block thinking, whereas moderate levels
enhance neuroplastic processes, which in turn support cognition and emotion. In essence, Bowen is
highlighting the fact that the simultaneous activation of cognition and emotion leads to neural integration
(1978). Increased differentiation of individuals within a family will decrease the overall rigidity of the system.
This process also allows family members to become more responsive to the needs of others and less reactive to
their own inner conflicts.
The first step in systems therapy is to educate the family about these concepts and to explore the history of
both sides of the family through the past few generations. In the context of systems theory and family history,
presenting problems are reframed and made more understandable. Uncovering family secrets and reality
testing around existing myths and projections allow for cortical processing of primitive and unconscious
defenses. The process of family therapy involves a series of experiments with increasing levels of
differentiation. Communication skills, assertiveness training, and exercises in new forms of cooperation can all
increase cortical involvement with previously reflexive or regressive emotions and behaviors. Often the person
with the symptoms needs to take more responsibility, while pathological caretakers must learn to accept
nurturance. Each member of the family needs to achieve a balance between autonomy and interdependence.
Ultimately, psychological, interpersonal, and neural integration are different levels and manifestations of the
same process.

Reichian, Gestalt, and Somatic Therapy


I am not in this world to live up to other people’s expectations, nor do I feel that the world must live up to mine.
—Fritz Perls

Wilhelm Reich, one of Freud’s early disciples, felt that memory and personality are shaped and stored not just
in the brain but throughout the entire body. Because of this, Reich not only paid careful attention to his
clients’ musculature, posture, and breathing but also encouraged them to express themselves physically during
analysis. By beating their fists, stomping their feet, and using exaggerated breathing techniques, they
attempted to release normally inhibited emotions. Reich highlighted the importance of the therapist’s
interpretation of the nonverbal messages of the body, making them available for conscious consideration. His
theories led to the development of Rolfing (which uses deep body massage to evoke and process memories)
and Gestalt therapy (which focuses on drawing attention to nonverbal aspects of communication and
increased self-awareness).
Reich (1945) believed that the major focus of psychotherapy should be the analysis of the character,
something he saw as similar to Freud’s notion of ego. While Freud focused on verbal communication, Reich’s
major contribution was to draw more attention to the nonverbal and emotional aspects of the therapeutic
interaction. He contended that the problems people bring to therapy are embedded in their character armor,
shaped during development as an adaptation against real or imagined danger. Character armor forms as a
result of misattunement, neglect, or trauma at the hands of caretakers. This armor is preverbal and organizes
during the first years of life. According to Reich, early defenses take shape at all levels of the nervous system,
become encoded in our entire being, and, like the air we breathe, are utterly invisible to us. The defenses
identified by Reich reflect emotional memories from early preverbal experiences that are stored in sensory,
motor, and emotional networks of early memory. Because character armor is invisible to its owner, the
therapist’s job is to make the client aware of its existence, expression, and meaning. Current somatic therapies
have followed these principles and expanded into many techniques.
Gestalt therapy is a unique expression of Reichian theory that is particularly relevant to the notion of
neural integration. Gestalt, a German word meaning “whole,” reflects the orientation of bringing together an
awareness of conscious and unconscious processes; in other words, seeing the whole picture. Gestalt therapy’s
charismatic founder, Fritz Perls, used the term safe emergency for the experience that psychotherapists strive to

32
create in treatment (Perls, Hefferline, & Goodman, 1951). A safe emergency is a challenge for growth and
integration in the context of guidance and support. It is also a wonderful way to describe an important aspect
of good parenting. Therapists create this emergency by exposing clients to unintegrated and dysregulating
thoughts and feelings while offering them the tools and nurturance with which to integrate their experiences.
Safety is provided in the form of a supportive and collaborative therapeutic relationship, often in the context
of a group. The emergency is created by an unmasking of defenses, making unacceptable needs and emotions
conscious, and by bringing into awareness dissociated elements of consciousness.
The stories a client tells about his or her problems are often interpreted as self-deceptions, keeping from
awareness those feelings that are relevant to healing but less acceptable. Unconscious gestures, facial
expressions, and movements that are first brought to awareness. They then need to be exaggerated and finally
given a voice with the purpose of understanding and integrating experience. The therapist points out
contradictions, such as making positive statements while shaking the head “no” or smiling while talking about
a painful experience. These contradictions are explored as indications of emotional conflicts to be brought into
conscious awareness, the automatic, nonverbal, and unconscious processes primarily organized in right-
hemisphere and subcortical neural networks.
Gestalt therapy emphasizes the exploration of projection as an avenue for discovering aspects of the self
that have been difficult or impossible to accept. In the popular empty-chair technique, patients alternately play
the role of different parts of themselves to fully articulate the different sides of inner conflicts. The Gestalt
therapist believes that maximizing awareness of dichotomous aspects of the self will result in increased
maturation and psychological health. This process depends on the integration of the neural networks
responsible for each of these functions.

Common Factors
My work as a psychoanalyst is to help patients recover their lost wholeness and to strengthen the psyche so it can resist future dismemberment.
—C. G. Jung

In reviewing these different psychotherapeutic modalities, a number of principles emerge that unify the
various therapeutic schools. The first is that psychotherapy values openness, honesty, and trust. Each form of
psychotherapy creates an individualized experience designed to examine conscious and unconscious beliefs and
assumptions, expand awareness and reality testing, and encourage the confrontation of anxiety-provoking
experiences. Each perspective explores behavior, emotion, sensation, and cognition in an attempt to increase
awareness of previously unconscious or distorted material. The primary focus of psychotherapy appears to be
the integration of affect, in all its forms, with cognition and conscious awareness.
Intellectual understanding of a psychological problem in the absence of increased integration of emotion,
sensation, and behavior does not appear to result in change. All forms of treatment recognize the need for
stress, from the subtle disruption of defenses created by the compassion of Carl Rogers to the confrontation
with feared stimuli in exposure therapies. There is a recognition that the evocation of emotion coupled with
conscious awareness is most likely to result in symptom reduction and socioemotional growth. Whether it is
called symptom relief, differentiation, ego strength, or awareness, all forms of therapy are targeting dissociated
neural networks for the purpose of increasing integration.
When theories of neuroscience and psychotherapy are considered side by side, a number of working
hypotheses emerge. First, given that the human brain is a social organ, safe and supportive relationships are
the optimal environment for social and emotional learning. Empathic attunement and a nurturing focus on the
client by the therapist provide the context in which growth and development occur. By activating processes
involved in secure attachment, empathic attunement creates an optimal biochemical environment for neural
plasticity.
Second, we appear to experience optimal development and integration in the context of a mild to moderate
level of arousal or what we might call optimal stress. Optimal stress will create the most favorable
neurobiological environment for neural plasticity and integration. Although stress appears important as part of
the activation of circuits involved with emotion, states of mild to moderate arousal are ideal for consolidation
and integration. In states of high arousal, sympathetic activation integration functions. The ebb and flow of
emotion over the course of therapy reflects the underlying neural rhythms of growth and change.
Psychodynamic therapies alternate confrontations and interpretations with a supportive and soothing
interpersonal environment (Weiner, 1998). The systematic desensitization of CBT pairs exposure to feared
stimuli with psychoeducation and relaxation training in the presence of a coach and ally (Wolpe, 1958).

33
Bowen’s (1978) family systems approach focuses on pairing anxiety reduction with experiments in increasing
levels of independent and differentiated behavior. All forms of successful therapy strive to create safe
emergencies in one form or another where learning can occur based on active neuroplastic processes.
A third hypothesis is that the involvement of affect and cognition appears necessary in the therapeutic
process in order to create the context for integration of neural circuits with a high vulnerability to dissociation.
It has been said that, in psychotherapy, “understanding is the booby prize.” It is a hollow victory to end up
with a psychological explanation and new labels for problems that remain unchanged. On the other hand,
catharsis without cognition does not result in integration either. Neural integration parallels an increased
ability to experience and tolerate thoughts and emotions previously inhibited, dissociated, or defended against.
Affect regulation may be the most important result of the psychotherapeutic process across orientations,
because it allows us to be connected with the naturally occurring salubrious experiences in life.
Repeated alternating and simultaneous activation of networks will aid in their integration. Repetitive play
in children and the phrase “working through” in therapy best reflect this process. This concept parallels the
principle that “neurons that fire together, wire together” (Hebb, 1949; Shatz, 1990). The simultaneous
activation of neural circuits allows them to stimulate the development of connections within association areas
to coordinate and integrate their functioning.
Fourth, the coconstruction of narratives between parent and child or therapist and client provides a broad
matrix supporting the integration of multiple neural networks. Autobiographical memory creates stories of the
self capable of supporting affect regulation in the present and the maintenance of homeostatic functions into
the future. Autobiographical memory maximizes neural network integration by organizing vast amounts of
information across multiple functions. Thus, language, narrative, and autobiographical memory are important
tools for both neurological and psychological development.

Sam and Jessica


The deepest principle in human nature is the craving to be appreciated.
—William James

Being human means communicating with others via touch, eye contact, tone of voice, and words. Through
our interactions, we have the power to impact one another at every level. One of my most powerful
experiences of the truth of this fact did not take place in a seminar or consulting room, but rather at the home
of a friend. I had volunteered to watch his two young children for a few hours while he ran some errands. I
had known Jessica, four, and Sam, six, all of their lives. I was someone on an outer ring of their universe, an
attractive combination of familiar and new and completely unprepared for what was about to happen. The
minute their father left, they shifted from low to medium to high gear, and I found myself in the midst of a
frenzy of excitement.
Toys began flying out of closets and storage containers; games were begun and tossed aside; videos were
started, stopped, and replaced—a succession of Indian princes, mermaids, lion kings, ladies, and tramps. After
what felt like hours, I glanced at my watch to find only 15 minutes had passed! Four more hours at this pace?
I wasn’t sure that I could survive. I kept trying to refocus Sam and Jessica’s activity, but to no avail. At one
point, as we dashed from bedroom to den to living room, I sank to the floor in the hall and propped myself up
against the wall. When they realized that I wasn’t right behind them, they ran back to find me.
They stood panting, one on either side of me, wondering what new game I had concocted. My suggestion
that we sit and talk for awhile passed unnoticed. After a few seconds, Sam looked at his sister and yelled,
“Show Lou how you burp your dolly!” Both of them let out a scream and Jessica soon returned with an
adorable squishy doll. As I reached for the doll to hold and admire it, Jessica threw the doll on the floor face
first and drove her fists into its back. As Jessica and Sam took turns crushing the doll into the carpet, I
watched in horror, completely identifying with the doll. I had to hold back my urge to save the poor thing
from its vicious attackers.
I quickly reminded myself that I was feeling sorry for a ball of cotton and that I should turn my attention
back to the children. I also realized that rescuing the doll would be scolding Sam and Jessica for their
behavior, which I did not want to do. I struggled to make sense of what was happening and asked myself if
there might be some symbolic message in the way they were treating this doll. Jessica and Sam had
experienced a great deal of stress in their brief lives in the forms of severe physical illness, surgery, drug
addiction in the family, and an understandably overwhelmed support system. The frantic activity I was
witnessing may have reflected the accumulated anxiety from all they had gone through, mixed with normal

34
childhood exuberance. But how might knowing this be helpful to these two beautiful children?
As I reflected on these things, I was hit by the notion that perhaps the doll represented both Sam and
Jessica. This doll needed to be burped. It needed the help of an adult to alleviate its discomfort and regain a
sense of comfort and equilibrium. Perhaps Sam and Jessica were showing me that when they needed to be
comforted, they were met with more pain or, at the very least, insufficient understanding and warmth. Might
their behavior be a message? “Please, we need nurturance and healing!” Their world seemed chaotic and
unsafe, a whirlwind; these were the same feelings that they had created within me during the last half hour.
Was their behavior a form of communication?
They had each taken a number of turns burping the doll, and I suspected that their attention would soon
turn to me. What to do or say? I didn’t want to burp the baby their way, and my thoughts about what was
happening would be meaningless. I could feel my anxiety growing when finally, they both turned to me and
cried in unison, “Your turn!” I hesitated. The chant of “Burp the baby, burp the baby” began to rise. I looked
at both of them and said, “I know another way to burp a baby. Here’s how my mom burped me.” A cheer
went up, and I suspect that they assumed that I was going to set the doll on fire or put it in the microwave.
I gently picked up the doll and brought it to my left shoulder. Rubbing its back in a circular motion using
my right hand, looking down at it with tenderness, I quietly said, “This will make you feel better, little one.”
A silence fell over the hallway. I looked up to find Jessica and Sam transfixed, as if hypnotized. Their eyes
followed the slow circles of my hand, heads tilted like puppies. Their bodies relaxed, their hands limp at their
sides, calm for the first time.
After following the movement of my hand for about 30 seconds, Jessica looked up at me and softly asked,
“Can I have a turn?” “Of course you can,” I told her. At first I thought she meant that she wanted a turn
burping the baby. But then carefully, almost respectfully, she took the doll from me and placed it on the floor
with its back against the wall. She stepped over to me, climbed over my crossed legs and put her head on my
shoulder where the doll’s head had been. She turned to me and almost inaudibly said, “I’m ready now.” As I
rubbed Jessica’s back, I felt her growing more and more limp as she melted into my shoulder and chest. I half
expected Sam to tear her off, climb on himself, and turn it into a wrestling match. When I looked over to
him, I could see that he was in the same posture and state of mind that he had been in watching me burp the
doll. He eventually looked up at me and asked, “Can I have a turn?” Before I could answer, Jessica lifted her
head slightly and told him, “In a minute.”
After a while, she gave up her spot on my shoulder, and Sam had his turn being burped. It felt wonderful
to hold them in this way and to give them something they seemed to need so badly. After a few turns for each
of them, we went into the den, curled up on the sofa, one of them under each of my arms, and watched a
movie. Actually, I watched the movie—they dozed off after only a few minutes. While my eyes followed the
frenetic animation on the screen, my breathing paced theirs, and I shared the peace they seemed to be feeling.
I marveled at how they managed to communicate their pain and confusion by creating these same feelings
in me. Emotion is truly contagious and a powerful source of human connection. By having them set the initial
pace of our play, I told them that I respected their way of coping. Through the use of the doll, they
communicated that when they needed soothing, their anxiety was often met with more of the same. When I
burped their doll in a caring and loving way, I showed them that I was capable of soothing them if they were
feeling bad. By asking me to burp them, they told me I was trusted. In falling asleep, they said, “We feel safe,
and we know you will watch over us while we rest.” While none of this was spoken, the communication was
clear.
Our interactions with the doll changed Sam and Jessica’s state of mind and body as well as my own. I
believe that it not only impacted their attitudes and behaviors that afternoon, but it may have also changed
their brains in some small but perhaps permanent way. I could see this reflected in their faces and hear it in
the tone of their voices; something fundamental had changed that affected their entire beings. I provided
them with a metaphor through which they could reorganize their experience, have their needs met, and
regulate their emotions. Together, the three of us coconstructed a new narrative for them to use as a way of
soothing themselves and each other.
Were this process to be repeated enough times, their brains could reorganize around this metaphor of
nurturance and holding and enhance communication between networks of cognitive and emotional
processing. Perhaps Sam and Jessica could internalize a model of self-holding and nurturance that would help
them navigate future challenges. This kind of interaction is at the heart of all forms of psychotherapy,
regardless of philosophy or technique. All forms of therapy have their own versions of integrative metaphors,
serving to reorganize neural networks and alter human experience, hopefully, for the better.

35
Summary

In this chapter we have discussed some of the basic principles connecting the historical and conceptual
connections between psychotherapy and neuroscience. Four common factors related to the nature of social
relationships, optimal stress, the activation of affect and cognition, and the coconstruction of narratives
emerge from the review. In the chapters to come, we will explore the components and organizing principles of
the nervous system. These basic concepts will help us understand the neural mechanisms of the building and
rebuilding of the brain.

36
Part II
How the Brain Works:
The Legacy of Evolution

37
Chapter 4
The Human Nervous System:
From Neurons to Neural Networks
All functions of mind reflect functions of brain.
—Eric Kandel

Studying the human brain is a daunting task. In fact, the human brain is so vastly complex that it would take
tens of thousands of pages to do it justice. But how much do we really need to know about the brain to help us
in our work as therapists? My belief is that a basic understanding of the nervous system, without getting lost
in the details, can be very helpful. With this as our goal, we will move through a thumbnail sketch of the basic
structures, functions, and development of the nervous system. Keep in mind that this quick look at the
nervous system will be biased toward the aspects of human experience and behavior that are most relevant to
psychotherapy.

Neurons
It is impossible, in principle, to explain any pattern by invoking a single quantity.
—Gregory Bateson

The basic unit of the nervous system is the neuron, which receives, processes, and transmits signals via
chemical transmission and electrical impulses. There are an estimated 100 billion neurons in the brain, with
between 10 and 100,000 synaptic connections each, creating limitless networking possibilities (Nolte, 2008;
Post & Weiss, 1997). Neurons have fibers called axons covered with myelin, an insulator that enhances the
efficiency of communication. Because neurons myelinate as they develop, one way of measuring the maturity
of a neural network is to measure its degree of myelinization. Multiple sclerosis—a disease that breaks down
myelin—results in a decrease in the efficiency of neural communication, negatively impacting cognition,
affect, and movement (Hurley, Taber, Zhang, & Hayman, 1999). The white matter of the brain is white
because myelin is white (or at least light in color). The unmyelinated gray matter consists primarily of neural
cell bodies.
When a neuron fires, information is carried via an electrical charge that travels down the length of its axon.
Neurons communicate with one another across synapses (the spaces between neurons) via chemical messengers
called neurotransmitters. The combination of these two complementary processes creates the brain’s
electrochemical system. Many neurons develop elaborate branches, called dendrites, which form synaptic
connections with thousands of dendrites from other neurons. The relationships formed among these dendrites
organize the complex networking of the nervous system that encodes our memories, emotions, and behaviors.

Glia
Complex, statistically improbable things are by their nature more difficult to explain than simple, statistically probable things.
—Richard Dawkins

Although the focus of neuroscience research is usually on neurons, they only make up half the volume of the
cerebral cortex. The other half is made up of approximately one trillion cells known as glia. One reason we
know so much more about neurons is that they are approximately 10 times larger than glial cells. However, it
has long been known that glia play an important supportive role in the construction, organization, and
maintenance of neural systems (Frühbeis, Fröhlich, Kuo, & Krämer-Albers, 2013; Ge et al., 2012; Sha et al.,

38
2013; Shao et al., 2013). More recently, it has become apparent that they are also involved in synaptic
organization, neural network communication, and neural plasticity (Allen & Barres, 2005; Pfrieger & Barres,
1996; Sontheimer, 1995; Vernadakis, 1996). Neural plasticity refers to the ability of neurons to change the way
they are connected to one another as the brain adapts to the environment through time.
Astrocytes, the most abundant kind of glia, have been shown to participate in the regulation of synaptic
transmission and appear to be involved in the coordination and synchronization of synaptic activity (Fellin,
Pascual, & Haydon, 2006; Newman, 1982). There now appears to be glial as well as neural transmission.
There is also the distinct possibility that astrocytes both shape and modulate synapses (Halassa, Fellin, &
Hayden, 2007). Through evolution, the ratio of glial cells to neurons has steadily increased, leading some to
believe that our expanding cognitive sophistication is, in part, related to the participation of astrocytes in
information processing (Frühbeis et al., 2013; Nedergaard, Ransom, & Goldman, 2003; Oberheim, Wang,
Goldman, & Nedergaard, 2006). We revisit this in a later chapter when we discuss Einstein’s glial cells and
his exceptional imaginal abilities.

Neurogenesis
What we teach today is part biology and part history . . . but we don’t always know where one ends and the other begins.
—J. T. Bonner

Neurogenesis, the birth of new neurons via cell division, occurs in the lower regions of the ventricles, the fluid-
filled cavities within our brains. Some fish and amphibians that demonstrate ongoing neurogenesis possess
nervous systems that continue to grow in size throughout life (Fine, 1989). During evolution, it appears that
primates may have traded much of their capacity for neurogenesis to continue building existing neural
networks in order to retain past learning and develop expert knowledge. In other words, if, instead of being
replaced, neurons are retained and continually modified through the branching of their dendrites in reaction
to new experience, more refined learning may occur (Ming & Song, 2011; Purves & Voyvodic, 1987).
Neurons do not appear to have a life span, but they die off either as a function of neural pruning during
development (apoptosis) or because their biochemical environment becomes inhospitable. High levels of
cortisol, a lack of blood flow, or the buildup of harmful free radicals can all lead to neuronal death.
The traditional wisdom concerning neurogenesis in vertebrates, and especially primates, has been that new
neurons are no longer created after early development (Michel & Moore, 1995; Rakic, 1985). Despite
considerable evidence to the contrary, this dogma held sway through most of the 20th century. However,
research continues to demonstrate that new neurons are formed in the brains of adult birds (Nottebohm,
1981), tree shrews (Gould et al., 1997), rats (Akers et al., 2014), primates (Gould, Reeves, Fallah, et al.,
1999), and humans (Gould, Reeves, Graziano, et al., 1999; Ernst et al., 2014; Ming & Song, 2011). Further,
neurogenesis is regulated by environmental factors and experiences such as stress and social interactions (Eisch
& Petrik, 2012; Fowler, Liu, Ouimet, & Wang, 2002).
Humans have maintained the ability to create neurons in areas involved with new learning, including the
hippocampus, the amygdala, striatum, and the cerebral cortex (Eriksson et al., 1998; Ernst et al., 2014; Gould,
2007; Gross, 2000). The importance of these discoveries and the abandonment of the old dogma cannot be
underestimated. Nobel prize–winning neuroscientist Eric Kandel referred to Nottebohm’s discovery of
seasonal neurogenesis in birds as one of the great paradigm shifts in modern biology (Specter, 2001).

Neural Systems
I believe in God, only I spell it Nature.
—Frank Lloyd Wright

As the brain develops and matures, neurons organize in more and more complex networks for successful,
adaptive functioning. The two most basic divisions of the nervous system are the central nervous system (CNS)
and the peripheral nervous system (PNS). The CNS includes the brain and spinal cord, whereas the PNS is
composed of the autonomic nervous system and the somatic nervous system. The autonomic and somatic nervous
systems are involved in the communication between the CNS and the sense organs, glands, and the body
(including the heart, intestines, and lungs).
The autonomic nervous system has two branches, called the sympathetic and parasympathetic nervous

39
systems. The sympathetic system controls the activation of the nervous system in response to a threat or other
basic drives. The parasympathetic system balances the sympathetic system by fostering conservation of bodily
energy, immunological functions, and repair of damaged systems. A third system, referred to as the smart
vagus, operates in parallel to the parasympathetic branch of the autonomic nervous system and is dedicated to
fine-tuning bodily reactions, especially in social situations (Porges, 2007). These three systems will be of
particular interest in later chapters when we discuss attachment and the effects of stress and trauma.
Although MacLean’s formulation of the triune brain is seen as too simplistic by most neuroscientists,
many still recognize the tripartite division of the cerebral cortex, the limbic system, and the brain stem. Each
layer is thought of as having different responsibilities. The brain stem—the inner core of the brain—oversees
the body’s internal milieu by regulating temperature, heart rate, and basic reflexes. The structure and functions
of the brain stem were shaped during our genetic history, and they are fully formed and functional at birth.
The reflexes we see in a newborn who grasps her mother, suckles her breast, and knows to hold her breath
when put under water are genetic memories that were retained from our tree-dwelling ancestors are all
controlled by the brain stem.
The outer layer of the brain, the cerebral cortex, is first organized by, and then comes to organize, our
experiences and how we interact with the world. As we grow, the cortex allows us to form ideas and mental
representations of ourselves, other people, and the environment. In contrast to the brain stem, the cortex is
experience dependent, which means that it is shaped through countless interactions with our social and
physical worlds. In this way, we grow to adapt to the particular and unique physical and social niches into
which we are born.
The two halves of the cerebral cortex have gradually differentiated during primate evolution to the point
where each hemisphere has developed areas of specialization, which is referred to as laterality. Language is the
best-understood example of lateral specialization. The two cerebral hemispheres communicate with each other
primarily via the corpus callosum, which consists of long neural fibers that connect the left and right cortices.
Although the corpus callosum is the largest and most efficient form of communication between the
hemispheres in adults, there are a number of smaller cortical and subcortical interconnections between the two
halves of the brain (Myers & Sperry, 1985; Sergent, 1986, 1990).
The cortex has been subdivided by neuroanatomists into four lobes: frontal, temporal, parietal, and
occipital (Figure 4.1). Each is represented on both sides of the brain and specializes in specific functions: the
occipital cortex includes the areas for visual processing; the temporal cortex for auditory processing, receptive
language, and memory functions; the parietal cortex for linking the senses with motor abilities and the
creation of the experience of a sense of our body in space; and the frontal cortex for motor behavior, expressive
language, and directed attention. The term prefrontal cortex is often used to refer to the foremost portion of
the frontal lobe. Two additional cortical lobes, the cingulate and insula cortices, are gaining increasing
recognition as distinct and important areas of the cortical-subcortical interface. They are involved in the
integration of inner and outer experience and linking the rest of the cortex with somatic and emotional
experience.
Between the brain stem and the cortex lies a region referred to as the limbic system, which is centrally
involved with learning, motivation, memory, and emotion. Because this book focuses on development and
psychotherapy, you will notice repeated references to two limbic structures. The first is the amygdala, a key
component in neural networks involved in attachment as well as the appraisal and expression of emotion
throughout life (Cheng, Knight, Smith, & Helmstetter, 2006; Phelps, 2006; Strange & Dolan, 2004). The
other is the hippocampus, which organizes explicit memory and the contextual modulation of emotion in
collaboration with the cerebral cortex (Ji & Maren, 2007).

FIGURE 4.1
The Four Lobes of the Cerebral Cortex
The four lobes of the cerebral cortex as seen from the left side of the brain.

40
Neurotransmitters and Neuromodulators
Brains exist because the distribution of resources necessary for survival and the hazards that threaten survival vary in space and time.
—John Allman

Recall that, within the nervous system, neurons communicate with each other via chemical messengers called
neurotransmitters. Different neural networks tend to utilize different sets of neurotransmitters, which is why
certain psychotropic medications impact different symptoms. Chemicals that serve as neurotransmitters
include monoamines, neuropeptides, and amino acids. Neuromodulators (e.g., the hormones testosterone,
estrogen, cortisol, and other steroids) regulate the effects of the neurotransmitters on receptor neurons. Amino
acids are the simplest and most prevalent neuromodulators. Glutamate is the major excitatory amino acid in the
brain and is central to neural plasticity and new learning (Cowan & Kandel, 2001; Malenka & Siegelbaum,
2001). Interactions with one of its primary receptors, N-methyl-D-aspartate (NMDA), regulates long-term
potentiation and long-term depression, thereby shaping the relationship between neurons believed to drive
learning (Liu et al., 2004; Massey et al., 2004; Zhao et al., 2005).
The monoamines—including dopamine, norepinephrine, and serotonin—play a major role in the regulation
of cognitive and emotional processing (Ansorge, Zhou, Lira, Hen, & Gingrich, 2004). All three are produced
in different areas of the brain stem, and they are carried upward via ascending neural networks to the cortex.
Dopamine, produced in the substantia nigra and other areas of the brain stem, is a key neurotransmitter in
motor activity and behavioral reinforcement. Too much dopamine can result in mood changes, increased
motor behavior, and disturbed frontal lobe functioning, which, in turn, can cause depression, memory
impairment, and apathy. Parkinson’s disease results from damage to the substantia nigra and a consequent loss
of dopamine. Many believe that schizophrenia is caused by too much dopamine, which overloads sensory
processing capabilities, and may lead to hallucinations and delusions.
Norepinephrine, produced in the locus coeruleus and other brain regions, is a key component of the
emergency system of the brain and is especially relevant for understanding stress and trauma. High levels
result in anxiety, vigilance, symptoms of panic, and a fight-flight response. Norepinephrine also serves to
enhance memory for stressful and traumatic events. Serotonin, generated in the raphe nucleus, is distributed
widely throughout the brain and plays a role in arousal, the sleep-wake cycle, and the mediation of mood and

41
emotion (Fisher et al., 2006). The action of popular antidepressant medications such as Prozac and Paxil
result in higher levels of available serotonin in the synapses and higher levels of neurogenesis (Encinas,
Vaahtokari, & Enikolopov, 2006).
The group of neurotransmitters known as neuropeptides includes endorphins, enkephalins, oxytocin,
vasopressin, and neuropeptide-Y. These compounds work together with neuromodulators to regulate pain,
pleasure, and behavioral reward systems. Endorphins tend to modulate the activity of monoamines, making
them highly relevant for understanding psychiatric illnesses. Endogenous endorphins (endorphins produced by
the body) serve as an analgesic in states of physical pain. They are also involved with dissociation and self-
abusive behavior, as we discuss in Chapter 16. The relationship between the monoamines and neuropeptides
is vitally important to the growth and organization of the brain.

Glucocorticoids/Cortisol
We’re lousy at recognizing when our normal coping mechanisms aren’t working. Our response is usually to do it five times more, instead of
thinking; maybe it’s time to try something new.
—Robert Sapolsky

Cortisol, the most important glucocorticoid, is often referred to as the “stress hormone.” It is produced in the
adrenal glands in response to dangerous situations as well as a wide variety of everyday challenges. The term
glucocorticoid comes from the fact that it was first recognized for its role in glucose metabolism. With further
study, however, many more functions of cortisol were uncovered. Glucocorticoid receptors (GRs) are found in
almost all of the tissues of our bodies. At normal levels and over short periods, cortisol enhances memory,
mobilizes energy, and helps to restore homeostasis after stressful situations. Glucocorticoids stimulate
gluconeogenesis and the breakdown of lipids and proteins to make energy available to us for emergencies.
Cortisol evolved to be useful for periods of brief stress which, when resolved, allow GRs to signal the
adrenal glands to shut down production. Prolonged cortisol release, on the other hand, can weaken the
immune system by preventing T-cell proliferation. In fact, the synthetic form of cortisol is called
hydrocortisone and is used to treat inflammation and allergies by inhibiting natural immunological responses.
Sustained high levels of cortisol disrupt protein synthesis, halt neural growth, and disturb the sodium-
potassium balance to the point of neural death. Early and prolonged stress has been correlated with memory
deficits, problems with affect regulation, and reduction of volume in brain regions including the hippocampus
and amygdala (Buchanan, Tranel, & Adolphs, 2004).
It is believed that sustained high levels of glucocorticoids early in life can have a negative impact on brain
development and make a child more vulnerable to subsequent stress. Maternal behavior in rats stimulates the
development of GRs in the brains of their pups, resulting in enhanced feedback to the adrenal glands to shut
down cortisol production. This is one of the underlying neurobiological correlates associating greater amounts
of maternal attention with resilience and positive coping later in life. The production and availability of these
neurochemicals shape all of our experience, from bonding to cognitive processing to our sense of well-being.
Regulation of these neurochemicals to control psychiatric symptomatology is the focus of the field of
psychopharmacology (Gitlin, 2007; Stahl, 2008).

Genetics and Epigenetics


I am convinced that it will not be long before the whole world acknowledges the results of my work.
—Gregor Mendel

At the forefront of the science of genetics is Abbot Gregor Mendel, who, in the garden of his ancient abbey,
discovered many of the principles of inheritance that still hold true. It turns out that his discoveries with pea
plants apply to animals and humans because the underlying mechanisms of heredity are similar for all complex
life forms. As you probably remember, his findings included dominant and recessive genes and the principles of
segregation and independent assortment.
With the benefits of modern technology, Mendel’s observations of the natural world were later understood
to be the effects of template genetics, or the way in which genes and chromosomes combine to pass along
traits from one generation to the next. We now know that our genetic information is coded in four amino acid
bases (adenine, thymine, guanine, and cytosine) that flow from DNA to messenger RNA (mRNA) to protein.

42
Although this understanding was a huge leap forward in our knowledge of the underlying processes of genetic
transmission, it accounts for only about 2% of genetic expression. The scientific term for the other 98% of
genetic material was “junk,” once thought to be accumulated debris of natural selection. It turns out, however,
that some of this junk actually plays an important role in guiding introns and exons, which help determine
whether specific elements of the genetic code get expressed or lie dormant.
Biologist C. H. Waddington coined the term epigenetics by combining the words genetics and epi, Greek
for over or above. Epigenesis describes the transformation of cells from their original undifferentiated state
during embryonic development into a specific type of cell. Thus, epigenetics is the study of how our genotype
is orchestrated into our phenotype. Understanding the elements of epigenetics may help us grasp why
identical twins with the same genes may differ in phenotype, that is, why one develops schizophrenia and the
other does not.
This gets us back to the old nature-nurture debate and the question: What do we inherit, and what do we
learn from experience? Our best guess is that almost everything involves an interaction of the two. While we
inherit a template of genetic material (genotype), which genes get expressed (phenotype) is guided by
experience. Experience can include anything from toxic exposure to a good education; high levels of sustained
stress to a warm and loving environment; feast to famine. Thus, many more genes are involved with the
regulation of what is expressed than with the direct synthesis of protein. So while template genetics may
primarily guide the earliest formation of the brain during gestation, the ongoing regulation of gene expression
by experience directs its long-term development in an ongoing adaptation to the social and physical worlds.
Epigenetics is a term used to describe this change in the phenotypic expression of genes in the absence of a
change in the DNA template.
An example of this process of particular relevance to emotional development and psychotherapy is the
impact of early stress on the adult brain. Meaney and his colleagues (1991) have shown that early
environmental programming of neural systems has a profound and long-lasting effect on the hypothalamic-
pituitary-adrenal (HPA) axis, which regulates an individual’s responsivity to stress. Research with rats has
demonstrated that the stress of early maternal deprivation downregulates the degree of neurogenesis and the
response to stress during adulthood (Mirescu, Peters, & Gould, 2004; Karten, Olariu, & Cameron, 2005).
Just as important for clinicians is that these processes are reversible later in life. As therapists, we attempt to
reprogram these neural systems via a supportive relationship and the techniques we utilize during treatment.
In other words, we leverage epigenetics to change the brain in ways that enhance our clients’ psychological
well-being.

Views of the Brain


When considering the abilities and complexities of the brain, one is struck by the incredible efficiency and splendor expressed in gray and white
matter.
—Julian Paul Keenan

Throughout most of the history of neurology, the human brain was only examined after injury or death. The
location of brain damage found on autopsy was linked to the nature and severity of the patient’s clinical
symptoms during life. Brain development was studied by examining and comparing the brains of humans at
different ages for size, the number of neurons, synapses, and dendrites, the degree of myelinization, and other
aspects of neural maturation.
Newer techniques allow us to examine brain structure in living subjects. Through the use of computerized
tomography (CT) and magnetic resonance imaging (MRI), we are able to see two- and three-dimensional
pictures of the living brain. Both of these techniques provide a series of cross-sectional images of the brain
through its many layers. CT scans do this via multiple X-rays. MRI scans utilize radio waves and a magnetic
field to study the magnetic resonance of hydrogen molecules in the water present in different brain structures.
In determining brain-behavior relationships, these measures need to be evaluated on the basis of whether they
are causes or correlates of the disorder being studied (Davidson, 1999). In their present practical applications,
radiologists learn to read these images for the presence and locations of tumors or lesions in order to assist
surgeons in their work. These scans have become an indispensable tool in neurology.
The functioning of the brain can also be measured in many ways. Clinical and mental status exams, tests of
strength and reflexes, and neuropsychological assessment all require a patient to perform physical or mental
operations that are tied to known neurobiological systems. These clinical tests are supplemented by a number
of laboratory tests that measure different aspects of brain functioning. The electroencephalograph (EEG)

43
measures patterns of electrical activity throughout the cortex. There are characteristic brainwave patterns in
different states of arousal and stages of sleep. Epilepsy or the presence of tumors will cause characteristic
alterations of normal electrical functioning, allowing EEGs to be used as diagnostic tools. EEGs can also be
used to measure brain development because neural network organization is characterized by the replacement
of local erratic discharges with more widespread and constant wave patterns (Barry et al., 2004; Field &
Diego, 2008b; Forbes et al., 2008).
The most exciting new tools in neuroscience are the various brain-scanning techniques providing us with a
window to the brain in action. Positron emission tomography (PET), single-photon emission computed tomography
(SPECT), and functional magnetic resonance imaging (fMRI) measure changes in blood flow, oxygen
metabolism, and glucose utilization, which tell us about the relative activity of different regions of the brain.
Using these techniques, neuroscientists can now explore complex activation-deactivation patterns of brain
activity in subjects performing a wide range of cognitive, emotional, and behavioral tasks (Drevets, 1998).
Most of these newer scanning techniques are still somewhat experimental, and the methodological standards
regarding their use and interpretation continue to evolve. These methods, and those yet to be developed, will
vastly enhance our understanding of the brain. As they grow increasingly accurate and specific, so too will our
knowledge of neural network functioning.

Brain Development and Neural Plasticity


Swiftly the brain becomes an enchanted loom, where millions of flashing shuttles weave a dissolving pattern—always a meaningful pattern—
though never an abiding one.
—Sir Charles Sherrington

Experience sculpts the brain through selective excitation of neurons, which shapes functional neural networks.
Paradoxically, the number of neurons decreases with age while the size of the brain increases. The surviving
neurons continue to grow from what look like small sprouts into microscopic oak trees. This process of
growth and connectivity is sometimes referred to as arborization.
In order for a neuron to survive and grow, it must wire with other neurons in increasingly complex
interconnections. Just as we survive and thrive through our relationships with others, neurons survive and
grow as a function of how well connected they are. Through what appears to be a competitive process referred
to as neural Darwinism, cells struggle for connectivity with other cells in the creation of neural networks
(Edelman, 1987). Cells connect and learning occurs through changes of synaptic strength between neurons in
response to stimulation. Repeated firing of two adjacent neurons results in metabolic changes in both cells,
which provides an increased efficiency in their joint activation. In this process, called long-term potentiation
(LTP) or Hebbian learning, excitation between cells is prolonged, allowing them to become synchronized in
their firing patterns and joint effectiveness (Hebb, 1949). LTP is believed to be a fundamental principle of
neuroplastic learning. Underlying LTP is the constant reaching out of small portions of the dendrites in an
attempt to connect with adjacent neurons. When these connections are made, neurons synthesize new protein
to build more permanent bridges between them.
Through LTP, cell assemblies organize into functional neural networks that are stimulated through trial-
and-error learning. This is only one small piece of a vastly complex set of interactions involving the
connection, timing, and organization of firing within and between billions of interconnected neurons in the
CNS (Malenka & Siegelbaum, 2001). Early in development, there is an initial overproduction of neurons that
gradually decreases through the process of apoptosis. Synapses that are formed may be subsequently
eliminated if they become inactivated or inefficient (Purves & Lichtman, 1980). In fact, elimination of
synaptic connections in the cortex continues to shape neural circuitry throughout life (Cozolino, 2008;
Huttenlocher, 1994).
In contrast to the brain stem and limbic system, the cortex is immature at birth and continues to develop
throughout adulthood. Because of this developmental timing, brain stem reflexes organize much of the
infant’s early behaviors, and the behavior of a newborn is dominated by subcortical activity. The neonate will
orient to the mother’s smell, seek the nipple, gaze into her eyes, and grasp her hair. A good example of a brain
stem reflex is the Moro reflex, by which the infant reaches out with open hands and legs extended, creating a
position conducive to grasping and holding (Eliot, 1999). The child’s eyes reflexively orient to the mother’s
eyes and face, and a baby’s first smiles are controlled by brain stem reflexes to attract caretakers. In fact,
children born with a genetic malformation that results in having only a brain stem still smile (Herschkowitz,
Kegan, & Zilles, 1997). These reflexes enhance physical survival and jump-start the attachment process by

44
connecting parent and child, enhancing their bond.
As anyone who has been pregnant can tell you, babies begin to engage in spontaneous activity of the arms
and legs well before birth. While the baby is practicing using its arms and legs, parents-to-be grow
increasingly excited as these signs of activity grow in frequency and strength. After birth, newborns continue
to move all parts of their bodies, which allows them to discover their hands and feet as they pass in front of
their faces. Although these movements may look random, they are the brain’s best guess at which movements
will eventually be needed. These reflexive movements jump-start the organization of motor networks to build
the skills the child will need later on (Katz & Shatz, 1996).
Through months and years of trial-and-error learning, these best guesses become shaped into purposeful
and intentional behaviors that are reflected in the organization of underlying neural networks (Shatz, 1990).
As sensory systems develop, they provide increasingly precise input to guide neural network formation for
more complex patterns of behavior. As positive and negative values are connected with certain perceptions and
movements—such as the appearance of the mother and reaching out to her—emotional networks will
integrate with sensory and motor systems. In the development of these and other systems, we find the
sequential activation of reflexive and spontaneous processes priming neural development, which comes to be
shaped by ongoing experience.

Cortical Inhibition and Conscious Control


He who conquers others is strong; he who conquers himself is mighty.
—Lao-tzu

The gradual attenuation of neonatal reflexes and spontaneous behavior corresponds with rising levels of
cortical activity and intentional behavior. As the cortex develops, vast numbers of top-down neural networks
connect it with subcortical areas. These top-down networks provide the pathways for inhibiting reflexes and
bringing the body and emotions under increasing cortical control. An example of this is the development of
the fine motor movements between the thumb and forefinger that are required to hold a spoon. Primitive
grasping reflexes allow only for the spoon to be held in a tight fist, rendering it useless as a tool. The
developing cortex enables the grasping reflex to be inhibited, while cortical networks dedicated to finger
sensitivity and hand-eye coordination mature. Thus, a vital aspect of the development of the cortex is
inhibitory—first of reflexes, later of spontaneous movements, and even later of emotions and inappropriate
social behavior.
Only through repeated trial-and-error learning are early clumsy movements slowly shaped into functional
skills. Children (and their brains) intuitively know this and will resist being held back or helped too much.
When we attempt to help, a child’s impatient protest of “Let me do it!” reflects instinctual wisdom of the
importance of trial-and-error learning in the growth of neural networks. This makes for many years of messes
and boo-boos. Another good example of the process of brain maturation is our ability to swim. The newborn’s
brain stem reflex to hold its breath and paddle when dropped into water is lost (inhibited by higher brain
circuitry) just weeks after birth. The skills involved with swimming need to be relearned as cortically organized
skills in the years to come; motor networks need to be taught body movements, as breathing becomes timed
and synchronized with each stroke.
Cortical inhibition and descending control are also central to emotional regulation. The rapidly changing
and overwhelming emotions displayed by very young children reflect this lack of control. As the middle
portions of the frontal cortex expand and extend their fibers down into the limbic system and brain stem,
children gradually gain increasing capacity to regulate their emotions and find ways to self-soothe. When
these systems are damaged or developmentally delayed, we witness symptoms related to deficits in attention,
emotional regulation, and impulse control.
We see the changes in motor control and posture as a child moves from being able to sit upright without
help at about six months, to crawling at about nine months, and then to walking without help by about one
year. At two years, a child will walk up and down stairs; by three years she can peddle a tricycle. As these skills
are shaped, so too are the brain systems dedicated to balance, motor control, visual-spatial coordination,
learning, and motivation that control them. The growth, development, and integration of neural networks
continue to be sculpted by environmental demands. In turn, neuronal sculpting is reflected in increasingly
complex patterns of behavior and inner experience.

45
Sensitive Periods
The principal activities of brains are making changes in themselves.
—Marvin L. Minsky

The brain continues to grow as long as we continue to learn, essentially until the day we die. Early brain
development is highlighted by periods of exuberant neural growth and connectivity called sensitive periods
triggered by the interaction of genes and experience. These sensitive periods are times of rapid learning,
during which thousands of synaptic connections are made each second (Greenough, 1987; ten Cate, 1989).
The timing of sensitive periods varies across individuals, which is why different abilities appear at different
ages.
The most widely recognized sensitive period is the development of language. At 24 months, an average
child understands and uses about 50 words, which increases to 1,000 words by 36 months (Dunbar, 1996).
The extent of neural growth and learning during sensitive periods results in early experience having a
disproportionate impact on our brains, minds, and experiences. As we learn of the brain’s ability to create new
neurons and retain plasticity throughout life, the importance of sensitive periods takes on new meaning. The
question for therapists is: How amenable are these established structures to modification? This is a topic we
will come back to again and again in later chapters.
The growth of neurons and the development of increasingly complex neural networks require large
amounts of energy. Patterns of increasing glucose metabolism during the first year of life proceed in
phylogenic order, meaning that the development of more primitive brain structures precedes those which
evolve later (Chugani, 1998; Chugani & Phelps, 1991). Early sensitive periods account for the higher level of
metabolism in the brains of infants compared to adults. Ever notice how warm a baby’s head is? It has been
estimated that in rats’ brains, 250,000 synaptic connections are formed every second during the first month
after birth (Schuz, 1978). Imagine what the number must be for humans.
Networks dedicated to individual senses develop before the association areas that connect them to one
another (Chugani, Phelps, & Mazziotta, 1987). The growth and coordination of the different senses parallel
what we also witness in such behavioral changes as hand-eye coordination and the ability to inhibit incorrect
movements (Bell & Fox, 1992; Fischer, 1987). As the cerebral cortex matures, a child at eight months is able
to distinguish faces and compare them to his or her memory of other faces. It is around this period that
stranger anxiety and separation anxiety develop. As the brain matures, we witness increasing cortical activation
and the establishment of more efficient neural circuitry firing in increasingly synchronous patterns.
Although both the left and right cerebral hemispheres are developing at very high rates during the early
years of life, the right hemisphere appears to have a relatively higher rate of activity and growth during the
earliest years (Chiron et al., 1997). During this time, vital learning in the areas of attachment, emotional
regulation, and self-esteem is organized in neural networks biased toward the right hemisphere. At around 18
months, this pattern of asymmetrical growth shifts to the left hemisphere.

Summary

The maturation and sculpting of so much of the cortex after birth allows for highly specific environmental
adaptations. The caretaker relationship is the primary means by which physical and cultural environments are
translated to infants. It is within the context of these close relationships that networks dedicated to feelings of
safety and danger, attachment, and the core sense of self are shaped. The first few years of life appear to be a
particularly sensitive period for the formation of these networks. Because there is so much neural growth and
organization during sensitive periods, early interpersonal experiences may be far more influential than those
occurring later. The fact that they are preconscious and nonverbal makes them difficult to discover and more
resistant to change. Because these neural networks are sculpted during early interactions, we emerge into self-
awareness preprogrammed by unconsciously organized hidden layers of neural processing. The structure of
these neural networks organizes core structures of our experience of self.

46
Chapter 5
Multiple Memory Systems
in Psychotherapy
To “do memory” is essentially to engage in a cultural practice.
—Kenneth Gergen

The process of psychotherapy is totally dependent upon memory. From what we know of clients’ past and
current lives to their ability to bring the lessons of therapy into practice, everything depends on their ability to
learn and remember (Kandel, Dudai, & Mayford, 2014). Yet, despite its central role in our work, the majority
of psychotherapists receive little or no training in the hows and whys of memory. In this chapter, we explore
various aspects of memory and their role in both mental illnesses and psychotherapy.
Psychotherapists have traditionally divided memory into the broad categories of conscious, preconscious,
and unconscious. Conscious memory is expressed in recollections of the past, the content of previous therapy
sessions, and reports of current day-to-day life. The preconscious contains memories that are not the focus of
current attention that can easily be brought into conscious awareness with a minimum of difficulty.
Unconscious memory unavailable to conscious consideration can manifest in behaviors, attitudes, and feelings
as well as in more complex forms such as defenses, self-esteem, and transference. Much of the training of
psychodynamic therapists is the identification and deciphering of unconscious memory into a form that is
accessible to the patient.
Freud believed that a fundamental goal of therapy is to make the unconscious conscious. From the
perspective of rebuilding the brain, this goal can be described as increasing the interconnection and
integration of neural networks dedicated to unconscious and conscious memory. This process makes
understanding the evolution, development, and functioning of the various systems of memory crucial to
conceptualizing and treating psychological distress and mental illnesses. It also aids in explaining to clients
some of the paradoxes and confusion they experience based on the variety of ways that their brains process
information.

Resistance to Therapy or Memory Deficit?


Our sense of worth, of well-being, even our sanity depends upon our remembering. But, alas, our sense of worth, our well-being, our sanity also
depend upon our forgetting.
—Joyce Appleby

For almost a year, I treated a woman named Sophia who had experienced repeated traumas and chronic stress
dating back to early childhood. During our time together, she discussed family conflict, early sexual abuse, and
current relationship problems. One of Sophia’s long-standing complaints was severe memory difficulties,
especially when it came to remembering names, dates, and appointments. In high school her teachers told her
she was stupid because she was unable to recall what was said in class from one day to the next. Sophia was so
embarrassed by her inability to remember names that she avoided parties and all but essential work gatherings.
On the other hand, her memory for emotionally laden experiences was like a steel trap, continually evoking
fear and sadness. Sophia was convinced that the part of her brain responsible for remembering shame was very
different from the one that recalled names.
Sophia had gone to many therapists throughout her adult life, repeatedly missed appointments, and had
been told she was resistant to treatment. Sophia found this very frustrating but had no explanation of her own.
Based on her history, these therapists assumed that her problems with memory were caused by denial,
avoidance, or repression, and they encouraged her to face her fears. While each therapist offered his or her
own interpretation of her defensiveness to treatment, none rang true, and she usually terminated therapy after

47
just a few sessions. Certain that it was her fault, Sophia’s treatment failures led her to feel increasingly
hopeless about ever finding the help she needed. The annoyance and criticism she received from therapists
also increased her feelings of shame. Although she feared our work together would meet the same fate, she
was willing to give therapy one more try.
After learning her history, I shared a bit of neuroscience to help her better conceptualize her issues with
memory. My mini lecture focused on the destructive role of early and prolonged stress on the development
and the well-being of the hippocampus and associated neural networks responsible for explicit memory. I
suggested that we begin our work by studying memory together and exploring pragmatic ways to improve it.
Along the way we experimented with the use of memory aids from the field of cognitive rehabilitation.
Timers, watches with alarms, and personal digital assistants all proved useful. The development of smart
phones now allows us to carry all of these functions in one device, a real boon to many patients.
For the first two months, Sophia and I scheduled telephone contact every other day for a few minutes.
During these contacts, we exercised her memory, checked on the various strategies we had set up during our
previous session, and reinforced her successes. Initially, Sophia needed help learning to remember to use her
strategies that helped her remember. Utilizing her memory aids and checking them on a regular basis
gradually became automatic even if, in the moment, she would forget why she was checking her book or
calling me.
After six weeks, Sophia was consistently able to remember appointments. This success stimulated
confidence in herself and in therapy. She began to see that her memory problems in no way meant she was
stupid or harbored deep psychological problems. On the contrary, her self-respect increased as our discussions
helped her to realize how much she had accomplished in her life despite her traumatic history and struggles
with memory. Once memory-related issues were no longer an impediment to maintaining consistent contact,
we shifted the focus of treatment to the impact of her life experiences on her relationships and career. The
initial focus of therapy, using a formulation from neuroscience and cognitive rehabilitation, turned out to be a
necessary first step in a sustained and successful therapeutic relationship. From this point, the therapy was
characterized by a more traditional psychodynamic approach with regular memory checkups and adjustments
to her strategies.
Many psychological disorders manifest a variety of memory deficits. Any disorder that results in
substantial arousal and triggers the secretion of the stress hormone cortisol can damage neural networks of
explicit memory. In fact, most psychiatric disorders involve high rates of cortisol and smaller hippocampi,
both of which are correlated with memory disturbances. In addition to problems with remembering, some
illnesses serve to distort both learning and memory. Depression, for example, results in a negative bias in the
recollection and interpretation of past, present, and future events (Beck, 1976). It also leads us to selectively
scan the environment, reinforcing our negative perceptions. Depression convincingly demonstrates the
influence of emotional states in the organization of conscious memory, sometimes called state-dependent
memory. Clients report that if they wake up depressed, everything looks worse than it did the day before, even
though they know, intellectually, that nothing has changed.
The rapid and unconscious networks of emotion shape our understanding of the world microseconds
before we become aware of our perception. Through similar mechanisms, our past experiences create our
expectations for the future. Implicit, unconscious memories that were created in dysfunctional situations years
before can repeatedly lead us to re-create unsuccessful but familiar patterns of thought, emotion, and behavior.
Thus, our perception of the world is a creation based on past experience.

Multiple Memory Systems


Memory . . . is the diary that we all carry about with us.
—Oscar Wilde

Research and clinical experience support the existence of multiple memory systems, each with its own
domains of learning, neural architecture, and developmental timetable (Tulving, 1985). Learning within all
systems of memory is dependent on the process of long-term potentiation in the Hebbian synapses we have
already discussed, as well as dendritic remodeling and changes in the relationships between neurons (Hebb,
1949; Kandel, 1998). The two broadest categories of memory are explicit and implicit. The concepts of
explicit and implicit memory, although similar in some ways to Freud’s concept of the conscious and
unconscious, do not directly overlap.
Explicit memory describes conscious learning and memory, including semantic, sensory, and motor forms.

48
These memory systems allow us to recite the alphabet, recognize the smell of coconut, or play tennis. Some of
these memory abilities remain just beneath the level of consciousness until we turn our attention to them.
Implicit memory is reflected in unconscious patterns of learning stored in hidden layers of neural processing,
largely inaccessible to conscious awareness. This category extends from repressed trauma to riding a bicycle, to
getting an uneasy feeling when we smell a food that once made us sick. Explicit memory is the tip of our
experiential iceberg; implicit memory is the vast structure below the surface (Kandel et al., 2014).
Many of our daily experiences make it clear that we have multiple systems of explicit and implicit memory.
For example, moving your fingers over the keypad of an imaginary phone sometimes helps you recall a phone
number. This process demonstrates that implicit systems of motor and visual memory can aid in the explicit
recall of numbers. Another example is a phenomenon common among older adults, in which they have
difficulty learning new information but easily recall stories from their youth. This may be because the
networks involved in the storage of long-term explicit memory are distributed throughout the cortex and are
more resistant to the effects of aging than those systems responsible for short- and medium-term memory
(Schacter, 1996).
Thinking back to the triune brain, each tier is involved with different aspects of memory functioning. The
reptilian brain contains instinctual memories, the lessons of past generations (genetic memory) that control
reflexes, and inner bodily functions. The paleomammalian brain (limbic system) contributes to emotional
memory and conditioned learning—a mixture of primitive impulses and survival programs sculpted by
experiences. These two systems are nonverbal and contain aspects of the Freudian unconscious. The
neomammalian brain, although largely unconscious in its processing, contains networks responsible for
explicit verbal memory biased toward the left hemisphere.
Because of the order in which they develop, implicit and explicit memory (detailed in Table 5.1) are
referred to as early and late memory. Systems of implicit memory are active even before birth, as demonstrated
in the newborn’s instincts to orient to the sound of the mother’s voice (de Casper & Fifer, 1980). During the
first months of life, basic sensory memories combine with bodily and emotional associations (Stern, 1985).
These networks allow for the sight of one’s father to be paired with raised arms, a smile, and a good feeling.
Somatic, sensory, motor, and emotional experiences help sculpt neural networks during the first few years into
a sense of a physical self.

TABLE 5.1
Multiple Memory Systems
A number of the basic distinctions between implicit and explicit systems of memory

Implicit Explicit
Early developing Late developing
Highly functional at birth Matures later with hippocampus and cortex
Subcortical/amygdala bias Cortical/hippocampal bias
Nondeclarative Declarative
Emotional Organized by language
Visceral/sensory-motor Visual images
Context free Organized within episodes and narratives
Procedural learning Conscious organization of experience
Behavior patterns and manual Construction of narrative self

The development of conscious memory parallels the maturation of the hippocampus and higher cortical
structures over the first years of life (Fuster, 1996; Jacobs, van Praag, & Gage, 2000; LeDoux, 1996;
McCarthy, 1995). Childhood amnesia or the absence of explicit memory from early life likely results from this
maturational delay and other developmental changes in how our brains process information (Bauer, 2015). In
the absence of explicit memory, however, we learn how to walk and talk, whether the world is safe or
dangerous, and how to attach to others. These vital early lessons, stored in networks throughout our brain,
lack source attribution; that is, we do not remember how we learned them. Although many of us think that we
have explicit memories from the first years of life, these are most likely constructed later based on hearing
stories from others and incorporating them as personal memories of our early life.
Explicit memory can be sensory and linguistic, as we associate and remember sights, sounds, and smells
with words and organize them in conscious memory. For most of us, words and visual images are the keys to
conscious memory. Different types of semantic memory include episodic, narrative, and autobiographical,
which can all be organized sequentially. Autobiographical memory maintains the perspective of the narrator at

49
the center of the story. Stories about the self combine episodic, semantic, and emotional memory with the
self-awareness needed for maximal neural network integration (Cabeza & St. Jacques, 2007). This form of
memory is especially important for the formation and maintenance of emotional regulation, self-identity, and
the transmission of culture.
Overall, the development of the different systems of memory reflects the early primacy of implicit memory
for learning in sensory, motor, and emotional networks. These early-forming neural networks depend on the
more primitive brain structures such as the amygdala, thalamus, and middle portions of the frontal cortex
(Figure 5.1). As the cortex and the hippocampus continue to develop over the first few years of life, there is a
gradual maturation of the networks of explicit memory. These systems provide for conscious, contextualized
learning and memory that becomes more consistent and stable over time. The various systems of memory are
distributed throughout the brain and where a particular memory is stored depends on the type of memory and
how it is encoded (McCarthy, 1995; Alberini & Ledoux, 2013).

FIGURE 5.1
The Amygdala and Hippocampus
The right half of the brain viewed from the left side. The hippocampus and the amygdala are located on the inferior and medial aspects of the temporal
lobes.

A good example of the distribution of memory comes from an experiment measuring cerebral blood flow
while subjects were asked to name pictures of either animals or hand tools (Martin, Wiggs, Ungerleider, &
Haxby, 1996). Naming both animals and tools resulted in increased activity in the temporal lobes and Broca’s
area. This makes sense, because the temporal lobes are known to be important for the organization of
memory, whereas Broca’s area organizes verbal expression. More specifically, naming tools activated areas in

50
the left motor cortex involved in the hand movements that would be used to control them (Martin et al.,
1996). This suggests that part of our tool memory is stored in neural networks that utilize them. While there
is overlap of activation during picture naming, the nature of the visual image triggers brain areas relevant to
what is depicted. Thus, memory is a form of internal enactment of whatever is being recalled.
The portion of the visual system activated by pictures of animals is an area involved with very early stages
of visual processing. This may be a reflection of how evolution has shaped the primitive areas of our visual
brains to recognize and react quickly to threats from possible predators (animals chosen for this study
happened to be a bear and an ape, both evolutionarily relevant based on their potential danger to us). Research
has consistently demonstrated that the occipital lobe becomes activated when something is seen and, later,
imagined. In the case of the imagined memory, the prefrontal area also becomes activated, reflecting its role in
processing the instructions, staying on task, and accessing imagination. How neural networks in the prefrontal
cortex know how to do this is as yet unknown (Ungerleider, 1995).
Although these studies focus primarily on cortical activity, psychotherapy often involves the retrieval of
subcortical emotional memories. Emotional memories rely on subcortical structures such as the amygdala and
hippocampus, which are central to upcoming discussions of psychopathology and the impact of childhood
experiences, stress, and trauma on adult functioning.

Amygdaloid Memory Networks


Nothing fixes a thing so intensely in memory as the wish to forget it.
—Michel de Montaigne

The amygdala, the central hub of fear processing, is located within the limbic system and beneath the
temporal lobes on each side of the brain. It is fully developed by the eighth month of gestation, so that even
before birth, we are capable of experiencing intense physiological states of fear. During the first few years of
life we are dependent on caretakers for external modulation of the amygdala until we are able to regulate it
ourselves. In some ways the amygdala is our first cortex, playing a significant role in the networks involved in
emotional learning (Brodal, 1992). Portions of the amygdala (the basolateral areas) have evolved in tandem
with the expansion of the cerebral cortex in humans along with our abilities to assess and appraise the
environment (Stephan & Andy, 1977; Roozendaal & McGaugh, 2011).
The amygdala’s neural connectivity supports its participation in the integration of the different senses
within humans, with a special emphasis on vision (van Hoesen, 1981). It functions as an organ of appraisal for
danger, safety, and familiarity in approach-avoidance situations (Berntson et al., 2007; Elliott et al., 2008;
Sarter & Markowitsch, 1985). In association with medial areas of the frontal cortex, it connects emotional
value to the object of the senses based on both instincts and learning history, and translates these appraisals
into bodily states (Davis, 1992; LeDoux, 1986). It is a central neural player in associating conscious and
unconscious indications of danger with preparation for a survival response (Ohman et al., 2007). Most
important for psychotherapy is that it plays a behind-the-scenes role in creating emotional bias in conscious
processing by spinning our experience in ways that make us, for example, see the glass as half empty or half
full (Kukolja et al., 2008).
Two circuits of sensory input reach the amygdala in the adult brain. The first comes directly from the
thalamus and the other loops through the cortex and hippocampus before reaching the amygdala (LeDoux,
1994). The first system allows rapid responses during survival decisions based on a minimum of information.
The slower second system adds cortical processing (context and inhibition) to appraise ongoing perceptions
and behaviors. The amygdala’s direct neural connectivity with the hypothalamus, limbic-motor circuits, and
many brain stem nuclei allows it to trigger a rapid survival response. The emotional power of phobias and
flashbacks is greatly enhanced by the activation of intense somatic arousal provided by this direct connectivity.
Thus, the amygdala is one of the key components of affective memory, not just in infancy but throughout
life (Chavez, McGaugh, & Weinberger, 2009; Ross, Homan, & Buck, 1994). In a fully developed brain, the
amygdala enhances hippocampal processing of emotional memory by stimulating the release of
norepinephrine and glucocorticoids (McGaugh, 2004; McGaugh et al., 1993). Through these chemical
messages, the hippocampus is alerted to the importance of remembering what is being experienced—a key
component of new learning. The activation of the sympathetic nervous system enhances the chemical
environment within and between neurons, stimulating long-term potentiation and neural plasticity. We
discuss this topic in greater detail in later chapters when we address the impact of stress and trauma on the
brain.

51
The Amygdala and Unusual Experiences
The “uncanny” elements we know from experience arise either when repressed childhood complexes are revived by some impression, or when
primitive beliefs that have been “surmounted” appear to be once again confirmed.
—Sigmund Freud

Given the amygdala’s early development and its unique role in learning and memory, abnormalities of
amygdala functioning may be involved in some unusual human experiences (Brázdil et al., 2012). Electrical
stimulation of the amygdala has been shown to result in a wide variety of bodily sensations, feelings of anxiety,
déjà vu, and memory-like hallucinations (Chapman, Walter, Markham, Rand, & Crandall, 1967; Halgren,
Walter, Cherlow, & Crandall, 1978; Penfield & Perot, 1963; Weingarten, Cherlow, & Holmgren, 1977).
Because of its low seizure threshold, subtle seizure activity may trigger the amygdala to activate normally
inhibited sensory and emotional memories that then break through into conscious awareness (Sarter &
Markowitsch, 1985). These primitive memories may also become triggered by sensory cues of past fears and
account for posttraumatic intrusions (van der Kolk & Greenberg, 1987). Individuals under stress may be
particularly vulnerable to the intrusion of powerful but conscious memories, even from very early childhood
(Cozolino, 1997).
Primary process thinking and dreamlike experiences are more likely to merge with conscious awareness in
situations of decreasing contextual cues, such as near-sleep states or conditions of sensory deprivation
(Schacter, 1976). Decreasing contextual cues lessen the ability of the corticohippocampal systems to utilize
past learning to make sense of present experience and inhibit amygdala input to conscious awareness. This
may account for the success of projective testing in tapping into unconscious processing. In attempting to
make sense of ambiguous situations, subcortical circuits are more likely to guide conscious awareness.
Individuals with temporal lobe epilepsy (TLE) often experience extreme religiosity, suggesting that
stimulation of the amygdala can infuse everyday experience with a sense of deep significance. In other words,
its ability to inform the rest of the brain that we are experiencing something highly significant can be applied
in an inappropriate manner leading to odd and delusional thinking. The central nucleus of the amygdala also
has a high density of opioid receptors, which are biochemical mechanisms of bonding and attachment behavior
that are also implicated in alterations of consciousness (Goodman, Snyder, Kuhar, & Young, 1980; Herman
& Panksepp, 1978; Kalin, Shelton, & Lynn, 1995; Kalin, Shelton, & Snowdon, 1993). This suggests that
unregulated activation of the amygdala may be a neurobiological trigger for the religious preoccupations
occurring in some individuals with TLE. The fact that hypergraphia (writing a lot) can also be a symptom of
TLE has led many to speculate that some religious texts have been driven by unusual amygdala activation
stimulated by seizure activity.

Hippocampal Memory Networks


A memory is what is left when something happens and does not completely unhappen.
—Edward de Bono

The hippocampi, shaped like seahorses on either side of the human brain, are essential structures for the
encoding and storage of explicit memory and learning (Zola-Morgan & Squire, 1990) and play a central role
in the organization of spatial and temporal information (Edelman, 1989; Kalisch et al., 2006; O’Keefe &
Nadel, 1978; Selden, Everitt, Jarrard, & Robbins, 1991; Sherry, Jacobs, & Gaulin, 1992). The hippocampus
also participates in our ability to compare different memories and make inferences in new situations from
previous learning (Eichenbaum, 1992). If damaged, it can prevent new learning from occurring, condemning
the victim to forgetting everything a few seconds after it is experienced (Squire, 1987).
The hippocampus is noted for its late maturation, with the myelination of cortical-hippocampal circuits
continuing into early adulthood (Benes, 1989; Geuze, Vermetten, & Bramner, 2005). The late development
of the hippocampus and its connectivity with the cortex reflects both its delayed functional availability and
prolonged sensitivity to developmental disruption and traumatic insult. It remains particularly vulnerable to
hypoxia (lack of oxygen) throughout life. Mountain climbers and deep sea divers who experience regular
periods of decreased oxygen availability have been shown to have hippocampal damage and short-term
memory deficits. Gradual atrophy of the hippocampus appears to be a natural component of aging, along with
a corresponding decrease in explicit memory abilities (Gartside, Leitch, McQuade, & Swarbrick, 2003;
Golomb et al., 1993).
Research suggests that sustained stress results in excessive exposure of the hippocampus to glucocorticoids

52
(cortisol), released in response to acute stress (Sapolsky, 1987). Prolonged high levels of glucocorticoids can
result in dendritic degeneration, cell death, increased vulnerability to future neurological insult, and inhibited
hippocampal functioning (Kim & Diamond, 2002; Watanabe, Gould, & McEwen, 1992). Patients suffering
from posttraumatic stress disorder (PTSD) secondary to childhood trauma or combat exposure, prolonged
depression, TLE (de Lanerolle, Kim, Robbins, & Spencer, 1989), and schizophrenia (Falkai & Bogerts, 1986;
Nelson, Saykin, Flashman, & Riordan, 1998) have also been shown to have hippocampal cell loss. Decreases
in hippocampal volume have been shown to correlate with deficits of encoding short-term into long-term
memory and an increased vulnerability to psychological trauma (Bremner, Scott, et al., 1993; Gilbertson et al.,
2002). Given that chronic stress correlates with decreased hippocampal volume and that so many patients in
psychotherapy have experienced chronic stress, it is logical to assume that many patients (like Sophia) have
difficulty in those functions which depend upon the hippocampus.

Amygdaloid-Hippocampal Interaction
The struggle of man against power is the struggle of memory against forgetting.
—Milan Kundera

The relationship between the amygdala and hippocampus is extremely important to human experience and
contributes significantly to top-down and left-right integration. The participation of the amygdala is biased
toward both right and down systems, whereas the hippocampus plays a large role in left and top processing.
Put another way, the amygdala has a central role in the emotional and somatic organization of experience,
whereas the hippocampus is vital for conscious, logical, and cooperative social functioning (Tsoory et al.,
2008). Their relationship will impact affect regulation, reality testing, resting states of arousal and anxiety, and
our ability to learn emotional and more neutral information. The level and quality of the functional
connectivity of the amygdala and hippocampus will be impacted by temperament, life stress, and epigenetic
factors (Canli et al., 2006).
Douglas and Pribram (1966) suggested that the amygdala and hippocampus play opposite roles in an
attention-directing process. By accentuating small differences among inputs, the amygdala heightens
awareness of specific aspects of the environment (attention), whereas the hippocampus inhibits responses,
attention, and stimulus input (habituation) (Douglas, 1967; Kimble, 1968; Marr, 1971). The amygdala is
involved with generalization, while the hippocampus is involved with discrimination (Sherry & Schacter,
1987). In other words, the amygdala will make us jump at the sight of a spider, while the hippocampus will
help us to remember that this particular spider is not poisonous, so we shouldn’t worry. Their proper balance
will also allow us to stay close to others even when they cause us to become upset.
We can immediately see the relevance of these two systems to psychotherapy. The amygdaloid memory
system, organizing early shame experience, makes the patient with borderline personality disorder react to the
perception of abandonment when in reality, little or none exists. Therapy with this patient would utilize the
hippocampal-cortical systems to test the reality of these amygdala-triggered cues for abandonment in order to
inhibit inappropriate reactions. This reality testing helps us to distinguish real abandonment from innocent
triggers such as someone showing up a few minutes late for an appointment and inhibit inappropriate
emotional reactions. Remember, for a young primate, abandonment means death. The catastrophic reaction
of borderline patients to perceived abandonment is a result of the fact that they experience it as life
threatening.
Flashbacks, intrusive memories from traumatic experiences, likely reside in amygdala-driven memory
networks. PTSD victims describe flashbacks as powerful and multisensory, often triggered by stress, and
experienced as if they were occurring in the present (Gloor, 1978; LeDoux, Romanski, & Xagoraris, 1989;
van der Kolk & Greenberg, 1987; Brewin, 2015). These flashbacks are also characterized by being stereotyped
and repetitive (van der Kolk, Blitz, Burr, Sherry, & Hartmann, 1984), suggesting that they are not subject to
the assimilating and contextualizing properties of the cortex and hippocampus. A model of dual memory
processing, paralleling the amygdala/hippocampus distinction made here, has been previously proposed as
underlying mechanisms in both PTSD (Brewin, Dalgleish, & Joseph, 1996) and the reemergence of past fears
and phobias (Jacobs & Nadel, 1985).
Given the reciprocal nature of amygdaloid and hippocampal circuits, impairment of the hippocampus
should lead to an increased influence of the amygdala in directing memory, emotion, and behavior. This
imbalance toward the amygdala would also disrupt affect regulation. Depressed patients are overwhelmed by
their negative feelings and are unable to engage in adequate reality testing. Indeed, Sheline and her colleagues

53
noted both decreased hippocampal and amygdaloid volume in depressed patients (Sheline, Wang, Gado,
Csernansky, & Vannier, 1996; Sheline, Gado, & Price, 1998). Dysregulation of hippocampal-amygdaloid
circuits are likely involved in depressive symptomatology and disturbed reality testing (Pittenger & Duman,
2008). Research with rats has found that increased levels of serotonin lead to enhanced neurogenesis in the
hippocampus (Jacobs et al., 2000). This suggests that Prozac and Paxil may be effective in treating depression
because they boost hippocampal volume and its ability to moderate amygdala activation.

The Intrusion of Early Implicit Memory Into Adult Consciousness


All our knowledge has its origins in our perceptions.
—Leonardo da Vinci

Early memories stored in circuits of the amygdala and right hemisphere can intrude into adult consciousness
in a variety of ways. They become especially relevant to psychotherapy when they are the result of trauma and
impact our ability to love and work. Children who suffer early abuse may enter their school-age years agitated,
aggressive, and destructive. They may engage in fights, property damage, setting fires, or hurting animals,
resulting in criticism, punishment, and social exclusion. Although these behaviors are expressions of their
memories of abuse, others react with criticism and retaliation. This feedback, in combination with the
emotional damages from their abuse, evolves into an ever-deepening negative self-image.
In the absence of an explicit memory of their early trauma, these children experience their behavior not as
a reaction to a negative past event, but as an affirmation of their inner feelings of essential badness. Because
these experiences date back to the formation of preverbal sensory, motor, and affective memory systems,
victims often report feeling “evil to the core.” This is common in children who grow up in cults or with highly
authoritarian or abusive parents. The kids of soldiers, police officers, and ministers appear to be at particular
risk for the internalization of a negative self-image. Children of parents with obsessive-compulsive disorder
may also find they hold an extremely negative view of themselves. While people with this disorder need order,
cleanliness, and control, a newborn brings just the opposite into their lives. The child’s early implicit
memories are likely to be centered around being a source of annoyance, anxiety, and disgust to the parents.
The formation of attachment schema (a key form of implicit memory) guides and shapes relationships
throughout life. Because so many clients come to therapy with relationship difficulties, this implicit memory
system may be one of the most important to explore in psychotherapy. These networks of social memory give
rise to the phenomenon of transference, a process that brings these early unconscious memories into the
consulting room as they are played out between client and therapist. Enactments in psychotherapy, involving
the interplay between unconscious elements within the patient and the therapist, also activate these implicit
memories.
We have all experienced having our buttons pushed by someone; many of these buttons are the emotional
traces of personal experiences, stored in implicit systems of memory. Overreacting to something implies that
the difference between an appropriate reaction and how we actually react is attributable to a sensitivity that
comes from our learning history. The most common distortions based on the input of early memory are in the
direction of shame, a primary socializing affect starting at about 12 months (Schore, 1994). Individuals who
are “shame based” (Bradshaw, 1990) can find criticism and rejection in every interaction, resulting in a life of
anxiety, a struggle for perfection, and depression.
Silence is an ambiguous stimulus that activates systems of implicit memory. Silence may be golden but, in
therapy, it evokes a variety of implicit memories. The reaction of clients to the silence teaches us something of
their emotional history. During periods of silence, many clients assume that the therapist is thinking critical
thoughts. They imagine the therapist thinks they are boring, stupid, a waste of time, or bad clients. These
feelings usually mirror those based in problematic relationships with one or both parents. Furthermore, these
feelings are deep seated and tenacious, often taking many years to make conscious, examine, and modify. On
the other hand, some clients find silence to be a form of acceptance and a relief from the pressures of being
articulate and communicative. These stark differences in each client’s reactions to a similar situation are
convincing evidence of the workings of implicit memory and their effects on conscious experience.
A similar phenomenon occurs in individuals who become uncomfortable when they try to relax without
any distractions. The emotions, images, and thoughts that emerge in conditions of low stimulation (or the
absence of distraction) may hold clues to the workings of our brains and the aftereffects of early learning.
Defenses to escape negative feelings come to require constant action and distraction to keep us from becoming
frightened or overwhelmed.

54
The Malleability of Memory
The only paradise is paradise lost.
—Marcel Proust

The false memory debate of years past highlighted the limits of our knowledge of episodic memory. Highly
publicized cases of clinicians’ contribution to the construction of false memories have resulted in increased
training focused on memory processes. Most therapists are now aware of the vulnerability of conscious
memory to suggestion, distortion, and fabrication from both client and therapist (Loftus, 1988; Paz-Alonso &
Goodman, 2008). Those with otherwise excellent autobiographical memories are equally vulnerable to
creating false memories (Patihis et al., 2013). In fact, false memories have also been created in mice by
manipulating neurons in their hippocampi (Ramirez et al., 2013).
Research has demonstrated that memory can be implanted in experimental situations, and the subject soon
becomes certain that the false memories have actually occurred (Ceci & Bruch, 1993; Loftus, Milo, &
Paddock, 1995). A therapist’s belief that her client has been abused may influence that patient to
unconsciously fabricate a memory that they both then come to believe is true. Memory is also distorted by a
drive to achieve explanatory coherence in what we are trying to remember (Chrobak & Zaragoza, 2013). This
process is a clear demonstration of both the malleability of memory and the power of coconstructed narrative
and retrospective creativity in shaping our memory of experiences (Alberini, 2005; Anderson, Wais, &
Gabrieli, 2006; Dudai, 2006; Nielson, Yee, & Erickson, 2005).
Given that memory is encoded among neurons and within neural networks, the malleability of memory is
an observable manifestation of the plasticity of these neural systems. This malleability is certainly a stumbling
block to our justice system and must lead us to rethink our reliance on eyewitness testimony (Peterson, 2012).
The hundreds of convictions that have been overthrown by DNA evidence all attest to the inadequacy of our
present methods as well as the invisible prejudices in their perspective. On the other hand, this malleability
provides an avenue to the alteration of destructive memories in psychotherapy. Revisiting and evaluating
childhood experiences from an adult perspective often leads to rewriting history in a creative and positive way.
The introduction of new information or scenarios to past experiences can alter the nature of memories and
modify affective reactions to current situations.

The Magic Tricycle


The greatest weapon against stress is our ability to choose one thought over another.
—William James

Sheldon was a man in his late 60s who came to therapy for help with his many anxieties and fears. As a child,
his parents had hidden him from the Nazis in a storage room behind the home of a family friend. One day,
after finding out that she and Sheldon’s father would soon be taken to the concentration camps, Sheldon’s
mother told him to be a good boy, said goodbye, and left. While the family friends were kind to him, he spent
his days alone with a few toys, his small tricycle, and some scraps of food. Describing these days, Sheldon
recalled alternating states of terror and boredom, during which he would either sit and rock or ride his tricycle
around in slow tight circles. The slightest noise would startle him, and he feared that each passing siren might
be the police coming for him. Each day, exhausted by fear, he would eventually fall asleep.
The intervening decades had not diminished the impact of his experiences during the war; 60 years later,
he still found himself reflexively rocking or walking in small slow circles when he became frightened. His life
felt like one long fear-filled day. In repeatedly recalling these experiences in treatment, he sometimes
mentioned how he wished he could have left the house where he was hidden and traveled down the narrow
streets to his grandmother’s house. Sheldon remembered long afternoons he spent there before the war,
listening to stories of her childhood on her father’s farm. His grandmother and his parents perished in the
war, and he never saw them again.
One day, I asked him for permission to change his memories just a bit. After a few quizzical looks he
agreed to close his eyes and tell me the entire story again, during which I would interrupt him and make some
suggestions. As he came to the part of the story where he rode around in circles, I asked him, “What would
you do if this was a magic tricycle and it could take you through walls without getting hurt?” I felt Sheldon
had sufficient ego strength to allow him to simultaneously engage in the role-play while staying fully in touch
with present reality.

55
After some hesitation, Sheldon said, “I would ride right through the house and out onto the sidewalk.”
“Fine,” I said. “Let’s go!” Sheldon had been primed for our imaginary therapy play because he had spent
many enjoyable hours of storytelling, cuddling, and laughing with his grandchildren. I felt that an imaginative
task like this not only was accessible to him but would also serve the purpose of bridging the positive affect
from his grandchildren to his lonely and frightened experiences as a child. Imagining that he was making up
the story for his grandchildren might also help him cope with the embarrassment of doing this with another
adult.
After some mild hesitation, he pedaled through the house. As he got close to the door, however, he said,
“They’ll see me and kill me.”
“What if the magic tricycle has the power to make you invisible?” I asked.
“I think that’ll do,” said Sheldon, and he pedaled through the front of the house and out onto the sidewalk.
Once he got out of the house, he knew what to do. He described the street to me as he pedaled toward his
grandmother’s house. The storekeepers, the neighbors, the park, his rabbi, even some of his young friends
were all alive in his memories. Sure enough, when he finally got to his grandmother’s house she was home
and, as always, happy to see him. He told his grandmother about his invisible tricycle and how scared he was
in his hiding place. He went on to tell her of the end of the war, his travels, and raising his family. Finally,
almost like a prayer, Sheldon told her how, many years from now, she would have the most beautiful great-
great-grandchildren living in freedom, redeeming her suffering.
Over the next few months, whenever Sheldon experienced his childhood fears and anxieties, we would
revisit his story and modify different details. These changes seemed to grow more detailed and more vivid in
his mind. His imagination gave him the power to master many of his past fears. Because memory is modified
each time it is remembered, Sheldon’s brain was able to gradually contaminate his painful childhood with his
present safety and joy. He even began to tell his grandchildren stories about a little boy with a magic tricycle
who accomplished great things with his courage and wit. Sheldon was a very special man who was able to take
advantage of the malleability of memory to make his inner world a safer place. Nothing had changed about his
childhood except that now, when he remembered his hiding place, he also remembered his magic tricycle.
An important part of restructuring memory is something Freud called Nachtraglichkeit, which means the
ability to reconceptualize a memory based on evolving maturity. This process requires being able to hold the
memory in mind without being emotionally overwhelmed and simultaneously bringing it into the present,
picturing it as it would look from the perspective of who we are and what we know today. Both Freud’s idea
and Sheldon’s experiences highlight the fact that memory is an evolving process that is also subject to positive
influence.
The construction and reconstruction of autobiographical narratives require that the semantic processing of
the left hemisphere integrate with the emotional networks in the right. Storytelling also invokes participation
of the body as we gesture and act out the events we are describing. As such, narratives are a valuable tool in
the organization and integration of neural networks prone to dissociation. Because we can write and rewrite
our own stories, new ones hold the potential for novel ways of experiencing. In editing our narratives, we
change the organization and nature of our memories and, hence, reorganize our brains. This is a central
endeavor in many forms of psychotherapy.

Summary

As a boy in the early 1960s, I remember being fascinated by news stories of Japanese soldiers attacking tourists
who landed on tiny islands in the South Pacific. I learned that during World War II, the Japanese navy left
soldiers on many islands throughout the Pacific but never retrieved them at the end of the war in 1945.
Decades later, pleasure crafts would innocently land on these islands only to be attacked by these soldiers who
thought the war was still being fought. They had dutifully kept guns oiled and remained vigilant for decades
in anticipation of an American attack. I was awed by their loyalty and saddened by the thought of the years
they had spent caught in a war that no longer existed while their families and friends went on without them.
Like these soldiers, early amygdala-based memory systems retain struggles, stress, and trauma from a time
before conscious memory. We may grow and move on to new lives, yet our implicit memory systems retain
old fears. While remaining vigilant for signs of attack for early attachment pain, approaching intimacy can set
off all of the danger signals. Therapists are trained to be amygdala whisperers who land on these beaches,
attempting to convince the loyal soldiers within implicit systems of memory that the war is over.

56
57
Chapter 6
Laterality: One Brain or Two?
Though the brain is enclosed in a single skull, it is actually made of two separate lumps . . . which are
designed to disagree with each other.
—Jonah Lehrer

We now switch our focus from the multiple and diverse systems of memory to another realm of neural
complexity—hemispheric laterality. As you know, the human cerebral cortex is divided into right and left
hemispheres, each controlling the opposite side of the body. The term laterality refers to the specialization of
certain tasks to one side of the brain or the other. Laterality is also reflected in how the hemispheres differ
from one another in their organization, processing strategies, and neural connectivity. Keep in mind that
laterality varies among individuals, left- and right-handed people, males and females, and young and old.
Although most neural processing requires the contribution of both hemispheres (Calvo & Bettran, 2014;
Shobe, 2014), there are situations in which the hemispheres not only think differently but also compete with
one another. This may be why we can be “of two minds.” John Hughlings Jackson, the eminent 19th-century
neurologist, believed that the left side of the brain was, for most people, the “leading” side. This seemed
logical given Broca’s finding that the left hemisphere was responsible for our ability to use semantic language.
Jackson later suggested that the right hemisphere was the leading side of the brain in visual-spatial abilities.
Over the years, it has become clear that dividing the brain into two discrete halves is not the best approach.
Given the fact that most neural systems integrate circuitry from the left and right sides of the brain, research
attempting to localize functions in one hemisphere or the other often results in “untidy” findings (Christman,
1994). When we speak of functions of the right or left brain, we are more accurately referring to functions
that are either represented more fully or performed more efficiently in one hemisphere than the other. Over
the past 40 years, much has been written about the artistic right brain and the logical left. Although this view
may be appealing to the imagination, it is far too simplistic. Assigning specific functions to particular areas of
the brain needs to be done with both caution and the recognition that our knowledge is still evolving.

Evolution and Development


A scientific truth does not triumph by convincing its opponents and making them see the light, but rather because its opponents eventually die and a
new generation grows up that is familiar with it.
—Max Planck

Lateral specialization is an evolutionary choice that does not exist in all animals. Many birds and fish, for
example, have identical hemispheres. These animals are able to sleep one hemisphere at a time, allowing them
to keep swimming or flying to avoid predators, continue feeding, or rest during long migrations. Although
redundant hemispheres are a backup system in case of injury, hemispheric specialization allows for greater
neural space and complexity for different tasks. Through human evolution, the right and left cerebral
hemispheres have become increasingly dissimilar (Geschwind & Galaburda, 1985). Lateral dominance
appears to have been delegated depending on the functional domain in question (Cutting, 1992; Goldberg &
Costa, 1981; Semmes, 1968). For example, areas of the left and right cortices have become specialized in the
organization of the conscious linguistic self in the left and the physical emotional self in the right.
During the first two years of life, the right hemisphere has a growth spurt that parallels the rapid
development of sensorimotor, emotional, and relational capabilities (Casey, Galvan, & Hare, 2005; Chiron et
al., 1997; Thatcher, Walker, & Giudice, 1987). The child learns hand-eye coordination, crawling, walking,
and how to attach to caretakers. An organized sense of the body in space and the embodied self form in
subcortical and cortical networks involving the thalamus, cerebellum, and parietal cortex. At the same time,
middle portions of the prefrontal cortex are maturing and integrating with subcortical structures to establish

58
the basic structures of attachment and emotional regulation. During this period, the development of the left
hemisphere is slowed a bit and reserved for later-developing functions (Gould, 1977).
In the middle of the second year, a growth spurt occurs in the left hemisphere and an explosion in
language and locomotion launches children into the broader physical and social worlds. In the frontal lobes,
there is a shift of development to the dorsolateral areas, linking back to other cortical regions, that sculpts the
language network (Tucker, 1992) while connecting the movements of hands and eyes to visual stimuli and
words. The corpus callosum begins to develop at the end of the first year, achieves considerable connectivity
by age four, and continues to mature past the age of 10. Because of this slow maturation, the two hemispheres
at first function relatively autonomously, gradually gaining interconnection and coordination throughout
childhood (Galin, Johnstone, Nakell, & Herron, 1979).
A great deal of what is known about the functions of the different hemispheres has been the result of the
split-brain research of Sperry and his colleagues (Sperry, Gazzaniga, & Bogen, 1969). Split-brain patients
have traditionally been individuals suffering from medication-resistant epilepsy, who have their corpus
callosum surgically severed to limit the spread of their seizures. Presenting information separately to each of
their hemispheres has revealed divisions of awareness and specialization across a range of cognitive and
emotional tasks (LeDoux, Wilson, & Gazzaniga, 1977; Ross et al., 1994; Sperry, 1968).

Lateral Asymmetry
All organs of an animal form a single system . . . and no modification can appear in one part without bringing about corresponding modifications
in all the rest.
—George Cuvier

The earliest form of language was most likely hand gestures, which may explain why handedness and language
functions are so closely linked in the brain (Corbalus, 2003; Hopkins & Cantero, 2003). Most of us are right-
handed (controlled by the left brain) and have semantic language lateralized in the left hemisphere. Neural
networks for both spoken and sign language are located in the left hemisphere for most adults, and damage to
these areas usually results in language disturbances such as aphasia (Corina, Vaid, & Bellugi, 1992). In left-
handed or ambidextrous individuals, lateralization of language is somewhat less clear. As the semantic
functions of the cortex expanded during evolution and language became more descriptive and useful, words
gradually replaced gestures in importance. Our present use of hand gestures to augment spoken language
likely betrays this evolutionary path. Our tendency to use hand gestures even when talking on the telephone
demonstrates that they also aid us in organizing and supporting our thinking.
The left hemisphere appears to be significantly more involved in conscious awareness, coping, and problem
solving than the right. This is most likely a function of its language skills and prosocial orientation. The left
hemisphere functions best within the middle range of affect and is biased toward positive emotions, anger
expressed toward others, and approach behaviors (Grimshaw & Carmel, 2014; Silberman & Weingartner,
1986). Strong personal affect, especially anxiety and terror, result in high levels of right-hemisphere activation
and appears to inhibit the left hemisphere and language—hence, the experience of stage fright and speechless
terror.
It has been suggested that Wernicke’s area in the left temporal lobe, known to be centrally involved in
language comprehension, acts as a probability calculator for other forms of behavior as well as language
(Bischoff-Grethe, Proper, Mao, Daniels, & Berns, 2000). Given the rapidity with which we process speech,
Wernicke’s area may process what is heard based as much on what it expects to hear as on what is actually
said. This would certainly help to explain why human communication can be so problematic and
misunderstandings so common. Broca’s area may have similar predictive functions, which allow us to speak
faster than we think and even, at times, be surprised by what we hear ourselves saying (Nishitani, Schürmann,
Amunts, & Hari, 2004). In fact, William James, considered the father of American psychology, said that he
needed to hear himself talk to know what was on his mind.
For most individuals, the right hemisphere processes information in a holistic fashion and is densely
connected to the limbic systems and the viscera (Nebes, 1971). The left hemisphere, on the other hand,
processes information in a linear, sequential manner and has less connection with the body. The right
hemisphere is heavily wired to the limbic system and is more directly involved in the regulation of the
endocrine and autonomic nervous systems than the left (Wittling & Pfluger, 1990). It also contains centers
within the parietal lobes that might contain a representation of the entire body.
The right hemisphere is generally responsible for both appraising the safety and danger of others and

59
organizing a sense of the corporeal and emotional self (Devinsky, 2000). Appraisal simply means attaching a
positive or negative association to a stimulus, while emotion is the conscious manifestation of this appraisal
process (Fischer, Shaver, & Carnochan, 1990; Fox, 1991). The vast majority of appraisal occurs at an
unconscious level. This is why the right hemisphere is more often associated with the unconscious mind, that
is, what guides our thoughts and behavior outside of our awareness.
The bias against left-handedness across many cultures may reflect an intuitive understanding of the left
hand’s (right brain’s) relationship to the dark, primitive aspects of our nature. These biases likely date back
into prehistory, when the left hemisphere may have exerted less inhibitory control over the right. Think about
the French word gauche and the Italian sinestre for left and all their tasteless and evil connotations. By offering
the right hand in greeting, early humans may have been more likely to behave in a civilized manner and less
likely to act out selfish or violent impulses. An examination of cave drawings in southern Europe suggests that
the bias toward right-handedness has existed for at least the last 5,000 years (Cashmore et al., 2008; Coren &
Porac, 1977).
Although the left hemisphere generally produces semantic language, it is unclear whether it has any
advantage in language comprehension. The right hemisphere may, in fact, be better at comprehending the
emotional aspects of language such as the tone of voice or the attitude with which words are said (Searleman,
1977). Social-emotional processing, the ability to evaluate emotional facial expressions, and visual-spatial and
musical abilities are primarily right-hemisphere processes (Ahern et al., 1991; De Pisapia et al., 2014).
Damage to the right hemisphere results in an impairment not only of our ability to assess facial gestures, but
also to comprehend other nonverbal aspects of communication such as hand gestures and tone of voice
(Blonder, Bowers, & Heilman, 1991).

Laterality and Emotion


When angry, count to four; when very angry, swear.
—Mark Twain

Evidence suggestive of a relationship between laterality and emotionality was first observed in cases of damage
to the prefrontal cortex. Patients with damage to the left hemisphere appeared to be far more likely to have a
depressive reaction than those with damage to the right (Gainotti, 1972; Goldstein, 1939; Sackheim et al.,
1982). It was later found that the closer these lesions were to the prefrontal regions, the more severe the
symptoms of depression (Robinson et al., 1984). Right-brain-damaged patients were also found to describe
experiences with less emotional intensity than left-brain-damaged patients or normal controls (Borod et al.,
1998).
Imaging studies have shown that people without brain damage who suffer from depression have lower
levels of glucose metabolism and cerebral blood flow in the left prefrontal cortex (Galynker et al., 1998; Kalia,
2005; Mathew et al., 1980). This may be true primarily for right-handed, left-dominant individuals (Costanzo
et al., 2015). In addition, people experiencing mania in the absence of brain damage demonstrate decreased
right prefrontal activity (Al-Mousawi et al., 1996). These studies expand the association between laterality
and emotion to the general population. An examination of Table 6.1 reveals that the left hemisphere is biased
toward positive affect, safety, and positive social approach, as well as anger and aggression directed toward
others. Overall, the left side of the brain appears to be in charge of the successful navigation of the rules of the
social world.
The intimate association between emotion and cognition has been demonstrated in many laterality studies.
For example, sad faces are rated as relatively sadder when presented to the left visual field compared to the
right (Sackheim et al., 1988). Negative stimuli are consciously perceived most often when presented to the
right hemisphere (Smith & Bulman-Fleming, 2004). Research has shown that anesthesia of the left
hemisphere results in greater expressions of negative emotions and less prosocial explanations of experience
(Dimond & Farrington, 1977; Ross et al., 1994). Orienting eye gaze to the left (stimulating the right
hemisphere) results in decreased optimism, while the opposite is true with rightward eye gaze (Drake, 1984;
Thayer & Cohen, 1985). In addition, right-hemisphere-biased neural processing correlates with low self-
esteem (Persinger & Makarec, 1991).

TABLE 6.1
Laterality and Emotion

Increased left hemisphere activation occurs in response to:

60
Happy stimuli1
Positive pictures2
Positive affect in response to positive films3
Approach-related dispositional tendencies4
More positive disposition5
Smiling and facial expressions of enjoyment6
Reported well-being7
Infant smiling in response to mother approach8
Trait anger9

State anger10
State aggression11
Resilience to negative life events12

Increased right hemisphere activation occurs in response to:


Facial expressions of disgust13
Tastes associated with disgust14

Negative pictures15
Avoidance behavior16

Negative affect17
Threat-related vigilance18

Stranger approach19
Maternal separation20

Depression21

Higher levels of left prefrontal activation have been associated with a resilient affective style, faster recovery
following negative events, and lower levels of the stress hormone cortisol (Davidson, 2004; Jackson et al.,
2003; Kalin, Larson, Shelton, & Davidson, 1998). While there appears to be an overall bias of positive
left/negative right, the picture is more complicated. The hemispheres are also lateralized for social/private and
approach/avoidance (left/right) behavior. These patterns of left/right activation suggest that health and
happiness may be associated with general lateral balance as well as the ability to be aggressive and express
anger biased toward the left and grief and shame biased toward the right.

The Integration of the Body in the Right Hemisphere


The body never lies.
—Martha Graham

The parietal lobes, located above our ears toward the top of our heads, are at the crossroads of neural networks
responsible for vision, hearing, and sensation. They serve as a high-level association area for the coordination
and integration of these functions. The anterior (front) portion of the parietal lobes organizes tactile
perception, while the posterior (back) portion integrates sensory and motor experience (Joseph, 1996).
Accordingly, cells in the parietal lobes respond to hand position, eye movement, words, motivational
relevance, body position, and other factors relevant to the integration of experience.
The purpose of the association of all of these high-order processing networks is to provide a coordinated
and integrated awareness of one’s own body and its relation to the external environment (Ropper & Brown,
2005). This makes sense in that the parietal lobes evolved from the hippocampus, which, in lower mammals,
serves as a cognitive map for external space (O’Keefe & Nadel, 1978). Part of the job of the parietal lobes is to
organize an integrative map of our bodies in space, which is available for conscious reflection. Thus, damage
to the parietal lobes, especially on the right side, results in a variety of disruptions in our experience of the self
and the world around us.

61
Although the left hemisphere seems to contain a network to monitor attention on the right side of the
body, the right hemisphere of right-handers has a specialized ability to direct attention bilaterally to both the
right and left sides of “extrapersonal space” (Mesulam, 1981). Hemineglect, or the lack of awareness of the
existence of the left side of the body, can result from lesions to the right parietal lobe. When neglect is severe,
the patient behaves as if the left half of the world has ceased to exist. Patients with hemineglect will dress and
put makeup only on the right side of their bodies while denying ownership of their left arm or leg. Asked to
draw the face of a clock, they may put all 12 numbers on the right side or simply stop at 6 o’clock.
The phenomenon of hemineglect has also been shown to exist in imaginary space. Bisiach and Luzzatti
(1978) examined two patients with right parietal injuries and left-side neglect who were asked to describe the
Piazza del Duomo in Milan. The piazza was very familiar to both patients. But when asked to imagine the
piazza from one end, they could recall and describe the details on their imagined right side and not their left.
Later, they were asked to reimagine the piazza from the other end. Looking back to where they previously
pictured themselves sitting, they were now able to accurately describe what was on the right side but not on
the left. In other words, once they imagined turning around 180 degrees, they now had access to memories
that they were unable to remember just a short while earlier. Further, the information they previously
provided was no longer accessible. This remarkable demonstration suggests neural networks that organize and
attend to the body in space are also utilized in imagination.
In later research, Bisiach and his colleagues (Bisiach, Rusconi, & Vallar, 1991; Cappa, Sterzi, Vallar, &
Bisiach, 1987; Vallar, Sterzi, Bottini, Cappa, & Rusconi, 1990) found that vestibular stimulation via cold
water irrigation of the left ear (the caloric test) in patients with right parietal lobe lesions resulted in temporary
remission of their left hemineglect. Putting cold water into the left inner ear stimulated areas within the right
temporal lobe and caused the patients to orient toward the left (Friberg, Olsen, Roland, Paulsen, & Lassen,
1985). Although the mechanism of action is not certain, one possible explanation could be that activation of
the right temporal lobe resulted in a reintegration of right and left hemispheric attentional processes, bringing
the world temporarily into an organized whole (Rubens, 1985). This theory is supported by the fact that being
shown fearful faces also appears to overcome the attentional neglect of these patients (Tamietto et al., 2007).
The survival value of these faces may surpass a higher threshold established in the hemineglect phenomenon.

The Language Network and the Left-Hemisphere Interpreter


All men are frauds. The only difference between them is that some admit it. I myself deny it.
—H. L. Mencken

The left-hemisphere language network relies on the convergence of auditory, visual, and sensory information
from the temporal, occipital, and parietal lobes. Wernicke’s area in the temporal lobe receives input from the
primary auditory area and organizes it into meaningful bits of information. The convergence zone connects
sounds, sights, and touch, so that cross-modal connections can be made, allowing us to name things we touch
and hear without visual cues. It is also necessary for the development of sign language, where words take the
form of gestures. This sophisticated and highly processed information projects forward to Broca’s area where
expressive speech is organized.
Neural networks linking language areas to the rest of the frontal lobes allow both spoken and internal
language to guide behavior and regulate affect. Although the semantic aspects of language are usually
lateralized to the left hemisphere, the right contributes the emotional and prosodic element of speech. The
integrative properties of language may be unequaled by any other function of the brain. Creating and recalling
a story requires the convergence of multisensory emotional, temporal, and memory capabilities that bridge all
vectors of neural networks. In this way, language integrates, organizes, and regulates the brain, and is
therefore used to great benefit in everyday storytelling as well as in psychotherapy.
Consistent findings across a variety of settings have led to a general acceptance that the verbal neocortex
organizes conscious experience and embodies the social self as arbiter of rules, expectations, and social
presentation (Nasrallah, 1985; Ross et al., 1994). Working with split-brain patients, Gazzaniga, LeDoux, and
Wilson (1977) found that the left hemisphere could create an explanation of experience when right
hemisphere information was unavailable. Gazzaniga (1989) later developed the concept of the left-hemisphere
interpreter that synthesizes available information and generates a coherent narrative for the conscious social
self.
The strategy of filling in gaps in experience and memory by making a guess at an explanation parallels
confabulatory processes seen in patients with psychosis, dementia, and other forms of brain damage.

62
Confabulation appears to be a reflexive function of the left-hemisphere interpreter as it attempts to make sense
of nonsense, organize our experiences, and present the self in the best possible light. This phenomenon is
likely related to Freudian defense mechanisms that distort reality in order to reduce anxiety.
A good example of this kind of confabulatory behavior was demonstrated by S. M., a 77-year-old suffering
from parietal and temporal lobe atrophy in her right hemisphere. One day her son saw her using sign language
in front of the mirror in her bedroom (Feinberg & Shapiro, 1989). When asked what she was doing, the
patient told him that she was communicating with the “other S. M.” She went on to tell him that there was
another S. M. who was identical to her in appearance, age, background, and education who was always in the
mirror. She and the other S. M. had gone to the same school, but did not know each other from that time.
The other S. M. also had a son with the same name who looked just like him.
S. M. and her double were identical in every respect, except that the other S. M. had a tendency to talk too
much and did not communicate as well as she did in sign language. If her son or the examiner appeared
behind her in the mirror, she would correctly label that person’s image in the mirror as a reflection. Thus, the
experience of a double was only applied to her own image. When it was pointed out that this was her own
image in the mirror, she would reply, “Oh sure, that’s what you think” (Feinberg & Shapiro, 1989, p. 41).
While S. M.’s comprehension and identification of herself and the world had been disrupted by her right-
hemisphere lesion, her left-hemisphere interpreter remained intact. It is somewhat comical to think that she
experienced her reflection in the mirror as talking too much and being less skilled than herself in sign
language. Perhaps the left-hemisphere interpreter may explain why we are all above average in our own minds.
This confabulatory and positive self-bias of S. M. versus her reflection is a perfect example of the left-
hemisphere interpreter at work. It also reflects the brain’s basic instinct to engage in explanatory behavior for
things it cannot understand. Some version of the interpreter concept has previously been used to explain the
development of paranormal beliefs (Cozolino, 1997), schizophrenic delusions (Maher, 1974), and religious
beliefs (Gazzaniga, 1995). The concept is especially relevant to psychotherapy because the construction of
reality is at work in the worldviews of patients with character disorders, the defense mechanisms of neurotics,
and the day-to-day reality of healthy individuals. The left-hemisphere interpreter is an internal press agent for
the self, putting a positive spin on what is experienced and how it is presented to others. If the interpreter is
not doing its job adequately, as in the case of left-hemisphere damage or decreased activation of the left
frontal cortex, we can become too realistic, pessimistic, and depressed.

Communication and Coordination Between the Hemispheres


Is the brain, which is notably double in structure, a double organ, “seeming parted, but yet a union in partition”?
—H. Maudsley

As our left and right hemispheres differentiated during evolution, each came to gain dominance for specific
functions after what must have been many failed experiments with transcortical democracy (Levy, Trevarthen,
& Sperry, 1972). Over time, the blending of the emerging strengths of each hemisphere allowed for an ever-
increasing integration of our cognitive and emotional functioning. When we are awake, the right hemisphere
silently provides information to the left, which we experience as intuition, feelings, fantasy, and visual images
(Nasrallah, 1985). The momentary bubbling up of feelings or images, which are then quickly lost, may reflect
one aspect of the intrusion of right-hemisphere processing into the stream of left-hemisphere conscious
awareness (Gainotti, 2011). The filtration of right-hemispheric processes may be necessary to allow us to
remain focused on the tasks in which we are engaged, although it may not necessarily register, understand, or
allow the information into awareness.
What happens when the hemispheres find themselves disconnected from one another? Jason and
Pajurkova (1992) reported the case of a 41-year-old right-handed man who suffered damage to the front
portion of his corpus callosum and the medial portion of his frontal cortex. The most salient aspect of his
behavior after his injury was that the two sides of his body seemed to be in conflict with one another. During
neuropsychological testing, the patient’s right hand would attempt to perform a task but the left would move
in and disrupt what had been accomplished. When he would try to go down a set of stairs, his right foot
would lead but then his left hand would grab the doorjamb and refuse to let him move forward. He found
himself unable to do things that required the cooperation of both hands.
The patient said, “My left foot and my left hand want to do the opposite of what my right one does all the
time” (Jason & Pajurkova, 1992, p. 252). On another occasion he stated, “My left hand doesn’t go where I
want it to” (p. 249). In each situation, the right hand and side (controlled by the left hemisphere) attempted to

63
carry out the conscious will of the patient. But the left side (controlled by the right hemisphere) would have
no part of it. The authors reported that it seemed as if the right hemisphere was acting like a spiteful sibling,
competing for attention and control (Jason & Pajurkova, 1992). Although this conflictual behavior decreased
over time, it was still evident six months after the injury. Similar left-right conflicts, usually resolving in the
first few weeks after surgery, have also been reported in split-brain patients.
It is clear in these cases that the left hemisphere takes the lead in the organization of the conscious self
(ego) while the behavior of the right hemisphere is experienced as a force from outside the self (ego-alien).
The experience and behavior of such patients suggests not only alternate ways of processing information in
each hemisphere, but also two separate wills. The unconscious and oppositional quality of the behavior of this
client’s right hemisphere suggests that the left hand may have been acting out unconscious emotional
reactions organized in the right hemisphere.

Right-Left Integration and Psychopathology


We use our brains too little and when we do, it is only to make excuses for our reflexes and instincts.
—Martin Fischer

I postulated earlier that neural network integration should correlate with mental health, while dissociation or
imbalance among neural networks should correlate with mental illness. If this is true, we can assume that
integration between the right and left hemispheres is one element of optimal brain functioning. It turns out
that anxiety, affective disorders, psychosis, alexithymia, and psychosomatic conditions have all been linked to
deficits in the integration and balance between the left and right brains.

Anxiety and Depression


Anxiety is love’s greatest killer.
—Anaïs Nin

As mentioned earlier, each hemisphere has an emotional bias, and so it appears that the proper balance of
right-left activation allows us to experience a healthy mix of positive and negative emotional experiences, as
well as to regulate and manage anxiety (Silberman & Weingartner, 1986). The left hemisphere has a bias
toward positive affect, prosocial behavior, and assertiveness, all of which help us to connect with others and
find safety in the group, while the right hemisphere’s bias toward suspiciousness and negativity keeps us
vigilant and alert to danger (Harmon-Jones et al., 2010).
Frontal lobe and amygdala activation, when biased toward the right hemisphere, correlates with the signs
and symptoms of anxiety and depression (Nikolaenko, Egorov, & Freiman, 1997; Perlman et al., 2012).
Primates with extreme right frontal activity are more fearful and defensive, and have higher levels of stress
hormones, than do those with activity biased toward the left hemisphere (Kalin et al., 1998). Adults with a
history of childhood trauma demonstrate a significantly greater shift to right-hemisphere processing when
asked to think about unpleasant memories (Schiffer, Teicher, & Papanicolaou, 1995). Activation of many
structures of the right hemisphere is also evident in PTSD (Engdahl et al., 2010; Rauch et al., 1996).
If anxiety and depression are, in part, the result of a bias toward right-hemisphere processing, then any
form of successful treatment will enhance a rebalancing of these systems. Cognitive therapies for both anxiety
and depression utilize rational thought that may work by activating left-hemisphere processes to regain lateral
balance. Symptomatic relief can also be achieved by a downregulation of right-hemisphere processes through
relaxation training.
An unfortunate artifact of the evolution of laterality may be that the right hemisphere is biased toward
negative emotions while also having primary control over emotional self-awareness (Keenan et al., 1999). In
addition, because there is so much early, unconscious right-hemisphere emotional learning, early negative
experiences have a long-lasting yet hidden impact on our self-esteem, attitudes, and personalities. These
aspects of laterality may create a bias toward shame, guilt, and pessimism while possibly explaining the
neurobiological mechanism underlying Nietzsche’s statement, “Man is the only animal who has to be
encouraged to live.” As we explore in a later chapter, this may not be an accident. Evolution may have utilized
what we consciously experience as shame to make us more easily controlled by others for the purpose of group
organization.

64
Alexithymia and Psychosomatic Illness
It is precisely because a child’s feelings are so strong that they cannot be repressed without serious consequences.
—Alice Miller

Alexithymia—the inability to consciously experience and describe feelings—is characterized by deficits in the
awareness and integration of right-hemisphere functions. Such patients are not prone to depression or mania
but instead have a poverty of emotional expression and experience. They are able to recognize that others have
feelings, but report being unable to locate any within themselves.
From a psychodynamic perspective, these patients seem trapped in secondary process thinking,
disconnected from their inner physical and emotional worlds. Patients with alexithymia are described as
having a concrete or stimulus-bound cognitive style, restricted imagination, and a lack of memory for dreams
(Bagby & Taylor, 1997). They have difficulty benefiting from traditional modes of talk therapy because of
their inability to bring emotions into the session or to use imagination or role-playing to expand their thinking
about themselves. Although the neurological correlates of this disorder are still unknown, alexithymia has
been described as a “bidirectional interhemispheric transfer deficit” (Taylor, 2000). The resultant failure of the
integration of affect and cognition leaves the conscious self of the left hemisphere with little input from the
emotional, intuitive, and imaginative right.
Patients with other psychiatric disorders reveal patterns similar to those with alexithymia. Hoppe (1977)
found that patients with psychosomatic disorders have characteristics similar to those with alexithymia such as
impoverished dreams, a paucity of symbolic thinking, and trouble putting feelings into words. Similar
difficulties were also found in Holocaust survivors, split-brain patients, and individuals with traumatic brain
injuries. Hoppe and Bogen (1977) hypothesized that problems during development or underlying genetic
processes could lead hemispheres to organize and function autonomously. The theory of such an
interhemispheric transfer deficit was supported by research with patients suffering from PTSD and
alexithymia who were found to have deficits in transferring sensorimotor information between hemispheres
(Zeitlin, Lane, O’Leary, & Schrift, 1989).

Psychosis
Reality is merely an illusion, albeit a very persistent one.
—Albert Einstein

Whereas normal states of awareness require an integration and balance of right- and left-hemisphere
processing, psychosis may be a result of the intrusion of right-hemisphere functioning into conscious
awareness. Hyperactivation of the right hemisphere or a decrease in the inhibitory capacities of the left may
diminish the ability to filter primary process input from the right hemisphere. This shift in right-left bias may
occur for many reasons, including changes in levels of important neurochemicals such as dopamine,
neuroanatomical abnormalities, or changing activation in subcortical brain areas such as the thalamus.
Schizophrenic patients and their close relatives demonstrate reduced left-hemisphere volumes in the
hippocampus and the amygdala, which has been shown to correlate with thought disorder (Seidman et al.,
1999; Shenton et al., 1992).
Auditory hallucinations, or hearing one or more voices talking, are a core symptom of schizophrenia. In
fact, the term schizophrenia means split mind. These aberrant, intrusive, and ego-dystonic experiences may
reflect right-hemisphere language (related to primary process thinking and/or implicit memories) breaking
into left-hemisphere awareness (Mitchell & Crow, 2005). These voices, often heard as single words with
strong emotional value, are experienced as coming from outside the self. For example, patients report hearing
profanities or critical words (jerk, idiot) as people walk by them on the street. Command hallucinations to
hurt oneself or others or to engage in dangerous behaviors have the same qualities. Schizophrenic patients
appear to openly struggle with shameful aspects of their inner world (likely stored in the right hemisphere)
that the rest of us are better able to inhibit, repress, and deny.
In psychosis, primary process thinking breaks into normal states of awareness to create what are diagnosed
as deficits in reality testing and thought disorders. Patients describe this as a feeling of dreaming while awake
and struggling to make sense of the simultaneous superimposition of primary and secondary process
experiences. This attempt to make sense out of nonsense fires up the left-hemisphere interpreter, leading to
the elaboration of bizarre delusions (Maher, 1974). Although a hemispheric model of psychosis is still

65
speculative, tests of lateral dominance (measured by a listening task) have shown that decreased lateral
dominance in these patients correlates with more severe psychotic symptoms (Wexler & Heninger, 1979).
Inspired by both modern science and ancient texts, the neuropsychologist Julian Jaynes (1976) developed a
theory of the evolution of human consciousness based on the increasing ability of the left hemisphere to
inhibit input from the right. Jaynes argued that prior to 1000 B.C., the two halves of the human brain acted
independently; the right hemisphere unconsciously controlled the body, while the left witnessed and described
the social environment and actions of the body. This model of laterality may have reflected an intermediate
evolutionary stage between having two modes of conscious awareness and our current bias toward right-
hemisphere inhibition.
Jaynes suggested that when our forebears were in situations of extreme stress, such as combat, the right
hemisphere provided auditory commands to the left, which were experienced as coming from outside the self.
This could reflect an internalized auditory memory of tribal leaders and warriors, commands similar to those
reported by modern-day schizophrenics. With the expansion of the corpus callosum and increasing
dominance of the left hemisphere, a more unified sense of self grounded in the left hemisphere has become
dominant and able to inhibit these inner voices. Jaynes felt that psychotic symptoms seen in patients in
modern times may be the result of a breakdown of the left hemisphere’s capacity to inhibit these messages
from the right.

Laterality and Psychotherapy


Happiness is not a matter of intensity but of balance, order, rhythm, and harmony.
—Thomas Merton

The proper balance and integration of the right and left hemispheres does not appear to be a given in the
course of development. I strongly suspect that left-right integration is an experience-dependent process that
relies on adequate assistance with affect regulation through secure attachment. It is also dependent on the
coconstruction of narratives where a model is presented for the recognition and labeling of feelings, as well as
integrating them into experience. Psychotherapy can serve as a means to reintegrate the patient’s disconnected
hemispheres through reality testing, emotional expression, and putting words to feelings in the context of a
caring relationship (Schore, 2011, 2014).
Examples from psychiatry and neurology strongly suggest that psychological health is related to the proper
balance of activation, inhibition, and integration of systems biased toward the left and right hemispheres.
Genetic and neuroanatomical factors can combine with early neglect or trauma to interfere with the
development of optimal neural network integration and regulation. The similarity between hemispheric
specialization and Freud’s notion of the conscious and unconscious mind has not been lost on
psychotherapists. Right-hemisphere functions are similar to Freud’s model of the unconscious in that they
develop first and are emotional, nonverbal, and sensorimotor (Galin, 1974). This nonlinear mode of
processing allows the right hemisphere to contain multiple overlapping realities, similar to Freud’s primary
process thinking most clearly demonstrated in dreams. The linear processing of conscious thought in the left
hemisphere parallels Freud’s concept of secondary process, which is bound by time, place, and social
expectations.
When patients come to therapy, the left-hemisphere interpreter tells its story. But something is usually
wrong: the story does not fully account for what is happening in their lives. The narratives that organize their
identities inadequately account for their experiences, feelings, and behaviors. The right hemisphere speaks via
facial expressions, body language, emotions, and attitudes. Thus, we listen to both stories for the congruence
between the verbal narrative and nonverbal and emotional communication. In this process, we analyze the
integration and coherence of left-right and top-down neural networks. A primary tool across all models of
therapy is editing and expanding the self-narrative of the left hemisphere to include the silent wisdom of the
right.
Hopefully, the therapist will be better integrated than the client in a therapeutic relationship. This will
allow the therapist to react to what is said with emotion, resonate with the client’s emotions, and then share
thoughts about those emotions with the client. Thus, the therapist’s ability to traverse the colossal bridge
between his or her own right and left hemispheres serves as a model and guide for the client.
Another way of describing therapy from the perspective of laterality is that we teach clients a method by
which they can learn to attend to and translate right-hemisphere processing into left-hemisphere language.
We teach them about the limitations and distortions of their own conscious beliefs presented by their left-

66
hemisphere interpreter. Many clients need to be suspicious of the ideas that their left hemispheres offer them.
This is why reality testing is so important for treatment success. It is the therapist’s job to hear what is not
said, resonate with what the client is unable to consciously experience, and communicate it back to him or her
in a way that will allow it to become integrated. This human process serves hemispheric integration.

Summary

The integration of dissociated processing systems is often a central focus of treatment. Gradually, clients come
to learn how the therapist gathers and interprets the information they present (Gedo, 1991). This process
closely parallels what is done during positive interactions with parents during childhood. If the method taught
during childhood is maladaptive, it leaves the child (and later the adult) in a state of limited self-awareness
and neural network dissociation. The learning of these skills in therapy occurs in the context of emotional and
cognitive integration, requiring the participation of both hemispheres, reflective language, feelings, sensations,
and behaviors. In the language of neuroscience, we are integrating dissociated systems of memory and
processing systems by teaching new strategies for integrating rational and emotional information. These
processes aid in the construction of a more inclusive self-narrative, which, in turn, serves as a blueprint for
ongoing neural integration.

67
Part III
Executive Functioning and
Neural Integration

68
Chapter 7
The Executive Brain:
Directed Action and Inhibition
My own brain is to me the most unaccountable of machinery—always buzzing, humming, soaring roaring
diving, and then buried in mud. And why? What’s this passion for?
—Virginia Woolf

After countless adaptational challenges and cycles of natural selection, we find ourselves with staggeringly
intricate and sophisticated brains. Ancient neural networks have been conserved, expanded, and reorganized,
while new ones have emerged to perform increasingly complex functions. In the process, some executive
functions remained within the brain stem and limbic system, while new ones developed in the cortex. Still
others were assumed by languages, relationships, and culture. So we still have primitive reflexes controlled by
the brain stem, and the amygdala of the limbic system continues to take over in crisis situations, while we now
also rely on the expectations of others to help guide our thoughts and behaviors.
Neurology was born in the context of phrenology, the study of bumps on our heads to diagnose
personality. This foundation led scientists to instinctually try to locate a place in the brain for every human
behavior. Ever since Paul Broca designated an area of the left frontal cortex as the “speech center,” the maps
that once located character traits such as wisdom and courage now support functions like vision, memory, and
receptive language. Certainly a step in the right direction, but still too simplistic. It turns out that each area of
the brain is part of a complex system from which it cannot be separated.
The control of the vast majority of our bodily and mental functions is on automatic pilot. Under normal
circumstances, we pay virtually no attention to breathing, walking, talking, and thousands of other complex
processes. We can drive a car safely (and mindlessly) for hours while listening to music or talking on the
phone. All of this automaticity allows us to focus our conscious attention on just a small fraction of what is
happening inside and around of us at any given moment.

The Central Dogma of Executive Functioning


The chief function of the body is to carry the brain around.
—Thomas Edison

There is no doubt that the correct answer for a seven-letter word describing “the region of the brain that
controls executive functioning” would be F-R-O-N-T-A-L. In recent years, however, neuroscience has gone
broader and deeper in search of an answer. For one, we realize that executive functioning consists of many
interwoven skills and abilities involving somatic cognitive, emotional, and social functions (Koziol et al.,
2011). For example, the notion of intelligence has expanded from cognitive functioning (problem-solving
skills and memory abilities) to include interpersonal and emotional abilities such as mind reading, attunement,
empathy, and affect regulation.
The impulse to localize functions to one area of the brain is evolving into more sophisticated strategies
that explore interactive networks distributed throughout the brain and nervous system (Takeuchi et al., 2012).
As we come to understand the complexity of executive functioning, it becomes obvious that something so
multidimensional must certainly require the contributions of many neural networks. While the frontal lobes
are central players, it turns out that executive functioning is a group effort.

In Search of a CEO

69
Effective leadership is putting first things first.
—Stephen Covey

Think for a moment of a large corporation with a CEO at the top of its executive hierarchy. Lower-level
managers who specialize in particular areas of operation are employed by the corporation to control thousands
of diverse functions. Utilizing multiple lower-level executives frees the CEO to monitor market forces, keep
an eye on the competition, and plan for the future. In the same way, the executive areas of the cerebral cortex
are freed from attention to basic bodily functions and well-learned motor behaviors. The executive brain
participates in more basic functions only in situations that are novel or problematic (Bermpohl et al., 2008).
Although the executive areas of the brain are traditionally thought of as being responsible only for our
rational abilities, they actually combine sensory, motor, memory, and emotional information to shape our
ideas, plans, and actions. This broader view of executive functioning has been guided, in part, by an increasing
appreciation of the contribution of emotion and intuition in decision making. Because so much of brain
functioning is hidden from conscious observation, the executive brain is strongly influenced by nonconscious
processes. Psychotherapy utilizes the executive brain to update and reorganize the relationship between the
conscious and unconscious networks in the service of mental and physical health.
The executive brain contains the control mechanisms that enable us to attend to a particular activity, filter
out distractions, make decisions, and act in an organized and purposeful way. If these functions are carried out
successfully, we feel safe enough to turn our attention inward for other aspects of executive functioning such as
contemplation, imagination, and self-awareness. These capabilities, in turn, create the possibility for art,
religion, philosophy, and other uniquely human endeavors. Executive systems also change through the life
span as the challenges of adulthood, parenting, and grandparenting call us to new roles and more complex
challenges. The function and organization of the brain continue to develop throughout life, allowing for
increasing perspective, compassion, and wisdom.

The Frontal Lobes


The highest possible stage in moral culture is when we recognize that we ought to control our thoughts.
—Charles Darwin

The frontal and prefrontal cortices are central to cognitive, behavioral, and emotional executive functioning in
both primates and humans. Their organization and connectivity provide for the integration of input from the
entire brain (Fuster, 1997). Because there are no primary sensory areas in the frontal cortex, they are entirely
dedicated to the association of sensory information that has already been processed in other regions of the
brain (Nauta, 1971). For example, projections from the parietal regions contain integrated visual, motor, and
vestibular information, while those from the temporal lobes transmit sensory information that has already
been synthesized with socioemotional appraisal.
Although the human frontal lobes initially evolved to organize complex motor behavior, the expansion of
the prefrontal lobes added capacities for planning, strategy, and working memory. Networks in both cerebral
hemispheres feed information to the frontal and parietal cortices. While hierarchical networks, which loop
through the cortex, limbic system, and brain stem, provide the frontal cortex with somatic and emotional
information, the convergence of all of these networks within the frontal and prefrontal lobes allows them to
synthesize diverse information to integrate our attention, emotions, and cognition with action (Alexander et
al., 1986; Marien, Custers, Hassin, & Aarts, 2012). Neurons and neural networks within the frontal cortex
organize our conscious sense of time by sustaining a memory for future consequences of behaviors about to be
performed (Dolan, 1999; Fuster, Bonder, & Kroger, 2000; Grinband et al., 2011; Ingvar, 1985; Watanabe,
1996). And, as the executive networks engage in different tasks, they reintegrate into different configurations
to address the specific tasks at hand (Braun et al., 2015).
Because of the evolutionary links between motor behavior and cognition, some theorists consider cognition
to be a derivative of motor behavior (Wilson, 1998). Support for this idea may exist in that much of our
symbolic and abstract thinking is organized by the visceral, sensory, and motor metaphors that permeate our
language (Goldman & de Vignemont, 2009; Johnson, 1987). To comprehend something is expressed as
“grasping” an idea or “understanding” a concept. Broca’s area, in the left frontal cortex that controls expressive
speech, is located adjacent to the area of the motor cortex dedicated to the lips and tongue. This proximity
may reflect the coevolution and interdependence of spoken language and fine motor control.
The prefrontal cortex also participates in constructing ideas about the beliefs, intentions, and perspective of
others in an extremely heterogeneous process referred to as theory of mind (TOM) (Goel, Grafman, Sadato,

70
& Hallett, 1995; Herrick, Brown, & Concepcion, 2014; Schaafsma, Pfaff, Spunt, & Adolphs, 2015; Stuss,
Gallup, & Alexander, 2001). Damage to the prefrontal cortex in early childhood usually results in deficits in
TOM, learning social roles, perspective taking, and empathic abilities (Balconi, Bortolotti, & Gonzaga, 2011;
Dolan, 1999; Koenigs, 2012). Damage to these areas later in life can sometimes result in a similar set of
deficits, referred to as pseudopsychopathy (Meyers, Berman, Scheibel, & Hayman, 1992).

The Cortex and Inhibition


What a man’s mind can create, man’s character can control.
—Thomas A. Edison

When we think of the evolution of the human cerebral cortex, we tend to think of the accomplishments of
music, art, and culture. Although we focus on these visible and impressive products of the human brain, the
hidden role of the cortex in inhibiting itself and other brain structures is just as important. Consider this
example: we are born with a broad array of primitive brain stem reflexes conserved from our primate ancestors.
One of these is the grasping reflex that allows us to pick up infants by putting our index fingers in their palms
and lifting. For the first few months of life, infants can hold their own weight, but later they are no longer
able to hold on.
It is believed that this grasping reflex is a holdover from a time when newborn primates had to hold onto
the mother’s fur to free the mother’s hands to traverse branches, gather food, and defend the child. Although
this behavior is no longer required for survival by humans, it has been retained within our genetic blueprint
through a process called evolutionary conservation (Jeannerod et al., 2005). It may continue to play a role in
enhancing the experience of bonding between newborn and parent as parents become physically and
emotionally “captivated” by their infant’s grasp.
Over the first few months of life, this reflex gradually diminishes as descending fibers from the cerebral
cortex connect with the brain stem regions that control it. But why does the cortex make this inhibitory
process such an early priority? After all, there is so much to learn. The most likely reason is that before the
cortical motor areas can begin to shape the dexterity of the hands and fingers, the hands need to be released
from the control of this primitive brain stem reflex. In other words, before we can move each of our fingers
independently and in complex patterns of coordination with each other, they need to be free from the impulse
to act as a unit with a single purpose.
Now fast forward to later in life, when the same child is 70 or 80 years old. Her children notice that she
seems forgetful and disorganized, and they become concerned that something may be wrong. So they take her
to a neurologist who performs a series of clinical tests. In one of these tests, the doctor asks her to hold her
arms out straight in front of her, hands open and palms facing down. Extending his arms under hers with his
palms up, the doctor slides his fingers under her arms from the elbows up toward her hands.
As the doctor reaches her wrists, he curls his fingers slightly and holds them rigid. As the doctor’s fingers
slide under her palms, he is looking to see if his touch triggers her fingers to curl inward and grasp his own. If
they do, he will try it again after telling her not to grasp his fingers. If it happens again, it is likely that the
touch of his fingers is triggering the same brain stem grasping reflex that she showed early in life. But why is
this clinically significant?
It turns out that early reflexes remain embedded within the brain stem throughout life and are continually
inhibited by descending fibers from the cortex. With diseases like dementia, the neurons in the frontal and
temporal regions of the cortex dedicated to descending inhibition gradually die off, thus releasing the grasp
reflex. The doctor is looking for signs of compromised cortical inhibitory functioning, which are suggestive of
a potential stroke, tumor, or the onset of dementia. Early reflexes that reemerge after damage to the brain in
adulthood are referred to as cortical release signs (Chugani et al., 1987; Walterfang & Velakoulis, 2005).
This inhibitory cortical function is not limited to primitive reflexes. A major neurobiological component of
secure attachment is the building of descending fibers from the prefrontal cortex down to the amygdala and
other regulatory structures. These inhibitory processes allow children to develop emotional control by using
their parents as emotional scaffolding as they learn to regulate their own fear through self-talk, memory of
positive outcomes, and proactive problem solving (Ghashghaei, Hilgetag, & Barbas, 2007). Because empathy
requires conceptual understanding, emotional attunement, and the ability to regulate one’s own affect, damage
to the prefrontal cortex often results in an impairment of empathic abilities (Eslinger, 1998).
Empathic thinking requires both affect regulation and cognitive flexibility in order to pull back from the
environment, put our own needs aside, and think about the feelings of others. People who have committed

71
murder, the ultimate lack of empathy, demonstrate significantly lower glucose metabolism in both dorsal and
orbital portions of the frontal lobes. The results were found in the absence of brain damage or decreased
metabolism in other areas of the brain (Raine et al., 1994). Although antisocial behavior is a complex
phenomenon, deficits in affect regulation, impulse control, and TOM are common in these individuals.
The classic example of damage to the orbitomedial prefrontal cortex (ompfc) is the case of Phineas Gage
(Harlow, 1868; Damasio, 1994). Mr. Gage was a well-respected New Hampshire railroad foreman who was
known for his maturity and “balanced” mind. An explosion on the job sent an inch-and-a-quarter-wide iron
bar up through his cheek and head, obliterating much of his ompfc. Although free of any “neurobehavioral”
deficits from the accident (such as aphasia, paralysis, or sensory loss), his workmates reported that Gage was
“no longer Gage.” After the accident he was unable to control his emotions, sustain goal-oriented behavior, or
adhere to social conventions, causing him to qualify for the diagnosis of pseudopsychopathy mentioned above.
Gage went from being a young man with a promising future to an aimless, unsuccessful, and troubled drifter
who was never able to regain his former level of functioning.

The Prefrontal Cortex


One of the most remarkable aspects of an animal’s behavior is the ability to modify that behavior by learning, an ability that reaches its highest
form in human beings.
—Eric Kandel

The prefrontal cortex is generally divided into two functionally related areas; the first consists of the orbital
and medial regions (ompfc), and the second is made up of the dorsolateral prefrontal cortex (dlpfc). Although
physically contiguous, the orbitomedial and dorsolateral prefrontal areas differ in their connectivity, neural
architecture, biochemistry, and functional responsibilities (Wilson, O’Scalaidhe, & Goldman-Rakic, 1993).
While both areas play a role in action and inhibition, the dlpfc and ompfc specialize in attentional and social-
emotional tasks, respectively.
The ompfc, first to evolve and first to develop during childhood, sits at the apex of the limbic system and is
richly connected with the amygdala and other subcortical regions involved in learning, memory, and emotion
(Barbas, 1995; Murray, O’Dougherty, & Schoenbaum, 2007). These connections, and their bias toward the
right hemisphere, are associated with processing social information and emotional experience (Bertouz et al.,
2012). Like the right and left hemispheres with which they are linked, the ompfc and dlpfc can demonstrate
various degrees of integration and dissociation. Table 7.1 lists some of the key functions of the prefrontal
regions.
The cognitive and emotional intelligences in which the dlpfc and the ompfc specialize have different
developmental sensitive periods and contexts for learning. Orbital and medial prefrontal areas begin to
organize in the context of interpersonal relationships even before birth. During the first 18 months of life, the
ompfc shares a sensitive period of development with the right hemisphere. Dorsolateral areas then have a
growth spurt along with the left hemisphere that correlates with the development of language and the
exploration of our physical and social worlds into the fifth year.
Our prefrontal cortex has two overarching and interwoven areas of function, the regulation of affect and
attachments on the one hand, and the synthesis and coordination of cognitive and motor processes on the
other. Although these two tasks seem quite different, each is dependent upon the other. Abstract thinking
and problem solving are particularly dependent on adequate emotional regulation. In turn, emotional
regulation is enhanced via rational thought and conscious problem solving. The prefrontal cortex also appears
necessary for thinking about thinking (metacognition), which includes observing our thoughts, revisiting
memories, and changing our minds (Jenkins & Mitchell, 2011).

TABLE 7.1
Functions of the Prefrontal Lobes

Orbital and Medial Regions


Sensitivity to future consequences9
Attachment1
Achieving goals10
Social cognition2
Stimulus-independent thought11
Thinking about a similar other3
Inhibitory control in emotional processing12
Self-referential mental activity4
Decisions based on affective information13
Appreciating humor5

72
Encoding new information6
Sensory-visceral-motor linkage7
Estimating reward value and magnitude8

Dorsal and Lateral Regions Learning motor sequences20

Cognitive control14 Decisions based on complex information21

Directing attention15 Thinking about a dissimilar other22

Organizing temporal experience16 The integration of emotion and cognition23

Organizing working memory17


Organizing episodic memory (right)18
Voluntary suppression of sadness19

We can observe an array of prefrontal functions by examining the kinds of problems that emerge when
they are damaged (see Table 7.2). With most traumatic brain injuries, like the one suffered by Luis, whom
you will hear about soon, all of these areas are negatively impacted. On the other hand, more localized lesions
as in the case of Phineas Gage may result in some symptoms but not others. Each psychiatric illness also has a
characteristic profile of cognitive distortions, difficulties with emotional regulation, and deficits of self-
awareness reflective of different patterns of frontal lobe involvement.
Problem solving, which requires emotional regulation, sustained attention, and cognitive flexibility, is a
central executive function that can become impaired with prefrontal damage. Some patients get stuck in a
particular way of thinking (perseveration), while others have difficulty utilizing abstract concepts (concrete
thinking). They may have difficulty remembering the outcome of past behaviors and repeatedly apply the
same unsuccessful solutions to new problems. Patients with frontal deficits often have a difficult time
monitoring social interactions, such as keeping the listener’s perspective in mind and abiding by social rules.

TABLE 7.2
Manifestations of Prefrontal Compromise

Orbital and Medial Regions Dorsal and Lateral Regions

Social and Emotional Disinhibition Loss of Focus and Planning


Tactlessness or silly attitude Forgetfulness
Decreased social concern Distractibility
Sexual exhibitionism and lewd conversation Decreased concern with consequences
Grandiosity Decreased anticipation
Flare with anger and irritability Poor planning ability
Restlessness Deterioration of work quality

Apathy Loss of Abstract Attitude


Decreased attention Concreteness
Loss of initiative Stimulus bound
Lack of spontaneity Loss of aesthetic sense
Indifference Perseveration
Depression Set stuckness

Luis
The very essence of instinct is that it’s followed independently of reason.
—Charles Darwin

Luis was in a serious auto accident a few days after his 20th birthday. He and his parents came in to see me
after his neurologist suggested that they all might benefit from family therapy. At the time of their first
appointment, I opened the door to find eight people packed tightly into my small waiting room. As Luis, his
parents, and five younger siblings filed into my office, I noticed the scars and indentations across Luis’s
forehead and imagined the damage beneath them. I knew from talking with his neurologist that he had
sustained severe injuries to his prefrontal cortex and that he had become impulsive, irritable, and occasionally
violent. Luis now possessed limited inhibitory capacity, reasoning abilities, and recognition of social

73
expectations.
After we all settled in my office, I turned to the father and asked how I could help him help his family. He
became tearful, shook his head slowly from side to side, and rubbed his hands together. “He drives too fast,”
he said quietly. “I don’t!” exclaimed Luis. “Except for that one time!” Everyone in the family looked away and
appeared embarrassed. It was immediately clear that talking back to his father was part of the problem.
Although he had always been somewhat impulsive, his parents claimed that he was far worse than before the
accident. I suspected that no matter how impulsive Luis might have been before the accident, this degree of
disrespectful behavior was new.
As the family discussed their situation, I found out that Luis’s parents had moved to the United States
from Mexico shortly before his birth and had adapted well to their new home. Despite their successful
acculturation, they remained true to traditional Mexican values of loyalty to the family and respect for elders.
In this context, Luis’s reflexive and loud contradiction of his father was a source of shame for everyone except
Luis. His injury had damaged the networks of his brain that allowed him to monitor and control his own
behavior and take into account the expectations of his family. A year after the accident, he returned to his auto
repair job but was unable to focus on his work or get along with coworkers and customers. For Luis, the
descending networks of cortical inhibition had been compromised through the loss of so many prefrontal
neurons.
Luis didn’t remember anything about his accident and, in fact, had no memory for the weeks before the
event. He read the police reports to discover that he had lost control of his car while street racing and crashed
into a pole. His injuries were compounded by the fact that he was not wearing a seat belt and had installed a
steel steering wheel without an airbag. It’s impossible to tell if this was the foolishness of adolescence or
evidence reflecting his lack of judgment prior to the accident.
His mother reported that Luis now spends most of his time at home with her and that his behavior was
erratic and sometimes frightening. At times he would cry for no reason, yell at her and the others, or jump in
her car and race off. A few times, he went into a rage and threw furniture around the house. He had also
made sexual statements and cursed using Jesus’s name during the holidays, upsetting everyone in the family.
Family members were torn between their love for Luis and their disgust with his behavior.
Automobile, industrial, and recreational accidents, as well as community and domestic violence, all
contribute to the increasing number of people who experience traumatic brain injuries. Because the frontal
areas are located directly behind the forehead, they are also most likely to be damaged in fights and accidents.
Although patients with head injuries come from all walks of life, young males are disproportionately
represented. Their youthful impulsivity, risk taking, and lack of judgment make them more vulnerable to
damaging the very regions required for understanding consequences and inhibiting impulses.
The massive reorganization of prefrontal brain areas along with biochemical and hormonal changes during
adolescence likely contributes to these dangerous behaviors (Spear, 2000). Many of these young men may have
already had frontal deficits or slowed frontal development prior to their accidents, amplifying more typical
adolescent risk taking. In this way, frontal injuries often compound preexisting deficits of impulse control and
judgment, complicating treatment and recovery.
Treatment with Luis and his family was multifaceted. I began by educating the entire family about the
brain and Luis’s particular injuries. The specific information was less important than labeling his behaviors as
symptoms of his injury rather than moral failings. In particular, I targeted his cursing and sexual statements,
which were, in their minds, connected to his character and spiritual health. By sharing case studies of others
with them, I was able to show that Luis’s symptoms were part of a pattern of pathological disinhibition related
to his brain damage, not the result of bad character or inadequate parenting.
More specific interventions included enrolling Luis in an occupational therapy program to help him
develop the instrumental and interpersonal skills needed to obtain and maintain employment. As the oldest
son, it was important for him and the rest of the family that he be productive and regain a sense of self-worth.
One of my goals was to reduce his resistance to taking medication that would help him with the anxiety and
depression caused by his changed circumstances. I also worked with Luis and his family to develop skills
related to stress reduction and anger management. We turned these exercises into family role-playing games
that alleviated tension and allowed everyone to participate in helping Luis.
Over time, Luis was able to apply his knowledge of cars to a part-time job in an auto parts store. His
occupational therapist helped him establish routines that allowed him to successfully use the computer.
Antidepressants proved helpful with both his mood and irritability, and the role-playing games became woven
into the family’s everyday interactions. All of these improvements made the occasional outbursts more
tolerable, and they could be seen more easily as a consequence of his illness. Luis was very fortunate to have
the unquestioning love and support of a strong and involved family. The quality of a patient’s support system

74
plays a large role in his or her recovery.

The Orbitomedial Prefrontal Cortex


Opinion is ultimately determined by the feelings, and not by the intellect.
—Herbert Spencer

Tucked under and between the lobes of the frontal cortex and sitting directly above the eyes, the ompfc is
densely connected to the anterior cingulate, amygdala, and other structures of the basal forebrain (Heimer,
Van Hoesen, Trimble, & Zahm, 2008; Zahm, 2006). These networks are of special interest to
psychotherapists because they both generate and regulate emotion, attachment, and preference (Chaudhry et
al., 2009; Kern et al., 2008; Lévesque et al., 2004; Rogers et al., 2004; Wager et al., 2008; Walton,
Bannerman, Alterescu, & Rushworth, 2003).
The anterior cingulate, involved with attention, reward-based learning, and autonomic arousal, first
appeared during evolution in animals demonstrating maternal behavior, nursing, and play (Devinsky, Morrell,
& Vogt, 1995; MacLean, 1985; Nair et al., 2001; Shima & Tanji, 1998). Consequently, damage to either the
ompfc or the anterior cingulate results in deficits of maternal behavior, emotional functioning, and empathy
(Etkin, Egner, & Kalisch, 2011). As described earlier, disorders of emotional control are also seen with
damage to these regions, including inappropriate social behavior, impulsiveness, sexual disinhibition, and
increased motor activity (Lewis et al., 2011; Price, Daffner, Stowe, & Mesulam, 1990). Together, the anterior
cingulate and the anterior insula serve as a salience network to direct our attention to novel and survival
relevant stimuli.
The ompfc is vital for appraisal—interpreting complex social events and linking them with their emotional
value via connections to the amygdala and other subcortical structures (Iidaka et al., 2011). A good example of
this is the ability of the ompfc to modulate the amygdala’s reaction to fearful faces based on the context in
which the faces are presented (Hariri, Bookheimer, & Mazziotta, 2000). So while the amygdala will alert us
to the sight of an angry face, the ompfc will add relevant environmental variables and information based on
past learning.
If the ompfc recognizes the face as that of a feared predator, the fight-or-flight response that has already
been activated by the amygdala will remain in place. If the ompfc evaluates the face as that of a distressed
baby, the ompfc will inhibit the fight-or-flight response so that we may approach the child to find out what is
wrong and if there is something we can do to help. Damage to either the amygdala or ompfc at any time
during life can result in an inability to organize vital social information in a useful manner, resulting in deficits
in communication, attunement, and social status (Adolphs, 2010; Cristinzio et al., 2010).
Research has demonstrated that the ompfc, along with the amygdala, calculates the magnitude of reward
or punishment value of our behavior such as approaching another for help and winning or losing money while
gambling. Much of this analysis occurs out of conscious awareness and is commonly called intuition. Those of
us who are good at reading people might just be aware of having an instinct or intuition about something. In
actuality, basal forebrain and somatosensory areas work together to appraise huge amounts of information that
provide us with this feeling about what to do even if it is sometimes contrary to reason (Damasio, 1994).

The Dorsolateral Prefrontal Cortex


Two things control men’s nature, instinct and experience.
—Blaise Pascal

The dlpfc integrates information from the senses, the body, and memory to organize and guide behavior. It
performs a variety of functions, including directing attention, coordinating working memory, learning motor
sequences, and organizing temporal experience (Fuster, 2004). The dlpfc is the latest-developing region of the
cortex and continues to mature into the third decade of life. This gradual maturation of neural networks is
vital to attention and judgment. It can be tracked by looking at the increasing complexity of school curricula
and later through the slow decline of automobile insurance rates from the teens into the 30s. The role of the
dlpfc in interacting and coping with the environment is highlighted by the reduced spontaneity and flattened
affect seen when they are damaged.
A component of the integration of top-down, cortical, and limbic processing occurs in the communication

75
between the ompfc and the dlpfc. The bias of these regions toward the right and left hemispheres,
respectively, allows them to support the integration of the left and right cerebral cortices. In addition, the
dorsal and lateral areas of the frontal cortex network with the hippocampus while the medial regions become
densely interwoven with the amygdala. Thus, the communication among prefrontal regions provides pathways
of integration for the hippocampal and amygdaloid memory systems described earlier (Gray, Braver, &
Raichle, 2002).
Like a tennis doubles team, the ompfc and the dlpfc depend on one another’s performance for optimal
functioning. If the ompfc is not doing an adequate job regulating amygdala activation, heightened levels of
autonomic arousal will interfere with dlpfc-directed cognitive processes (Dolcos & McCarthy, 2006). This is
why we often have difficulties in comprehension and problem solving when we are frightened or distraught.
On the other hand, if the dlpfc is not properly managing environmental demands, the resultant anxiety will
disrupt emotional regulation. In essence, both inner and outer worlds need to be balanced and integrated for
optimal functioning.

The Salience Network and the Focus of Attention


To pay attention, this is our endless and proper work.
—Mary Oliver

When it comes to choosing and maintaining our focus of attention, the prefrontal cortex is assisted by a
combined effort of the anterior portions of the cingulate and insular cortices. This is referred to as the salience
network. The cingulate cortex serves as a detector of novelty in our environment and becomes active when we
confront something new. The cingulate and insula also work together in (1) the integration of somatic and
cognitive processing, (2) the conscious experience of subjective feelings, (3) the selection of appropriate
behavioral reactions to the external world, and (4) the simulation of the internal states of others (Medford &
Critchley, 2010). Put in another way, the salience network serves to direct the focus of our attention to either
the inner or the outer world, and toward the experience of others or to our selves. See Table 7.3 for a list of
some of the activities regulated by this system.
By combining these functions, the anterior cingulate cortex (ACC) and anterior insula (AI) appear to
segregate information and direct attention toward the most relevant internal and external stimuli to guide our
emotional and interpersonal responses to others (Allman et al., 2006; Menon & Uddin, 2010; Wiech et al.,
2010). Together, they also appear to organize the introspective information available to make judgments based
on subjective preference (Chaudhry et al., 2009). In contrast, negative connectivity between the ACC and
insula has been shown to correlate with autistic traits in adults (Di Martino et al., 2009b).
As our brains mature, integration increases among the circuits of cognitive and emotional intelligence.
Activations in the ACC and the AI are modulated by cognitive instructions, expectations, or the
consequences of outcomes, demonstrating the influence of the prefrontal cortex on these structures (Lamm et
al., 2007; Lutz et al., 2009; Newman-Norlund et al., 2009; Sawamoto et al., 2000). While younger children
show activation mostly in the frontal and parietal lobes during executive functioning, there is an increasing
participation of the ACC and AI during adolescence (Houde et al., 2010). Based on its functions and
developmental trajectory, it is safe to assume that the healthy maturation of the salience network, along with
the regions of the prefrontal cortex we have discussed, all contribute to the development of self-regulation and
attentional capabilities.

TABLE 7.3
Joint Activation of the Anterior Cingulate and Anterior Insula

Bodily Awareness and Control

Awareness of one’s heart beating1 Emotional reactions to asthma symptoms4


Breathlessness or air hunger2 Suppression of natural urges (blinking)5
Performance monitoring3 Disgust in self and others6

Social Emotions

Resentment7 Deception14
Embarrassment8 Guilt15
Humor9 Being happily in love16

76
Sadness10 Grief after a romantic breakup in women17
Perceived unpleasantness11 Perceived unfairness18
Reaction to fearful faces12 Perceiving others’ pain19
Empathy for pain13

Cognitive Processing

Insight into problem solutions20 Sustained task focus across time22


Emotion-based decisions21 Risk-based decision making23

Attention-Deficit/Hyperactivity Disorder
Thinking is the momentary dismissal of irrelevancies.
—Buckminster Fuller

Jimmy, an elfin 8-year-old, was referred to me to assess whether or not he had attention-deficit/hyperactivity
disorder (ADHD). Before meeting him, I read notes from his parents, teachers, and soccer coach that
described his behavior. All agreed he was more distracted and energetic than other children his age. His coach
noted Jimmy’s inability to stay focused on the game; one teacher described him as a bundle of energy; his
father wrote, in big letters, “Exhausting!” Jimmy’s restlessness and impulsivity made it difficult for other kids
to interact with him, and his mother felt he was becoming isolated as his peers sought calmer company.
I walked into the testing room to find Jimmy’s mother slumped in a chair with her face in her hands. She
did not react when I entered the room, and I wondered if she might be crying. I scanned the room, looked
behind the chair and small sofa, but could not see Jimmy anywhere. Before I could speak, Jimmy shouted,
“I’m up here!” Startled, I looked up and saw him perched on top of a six-foot storage cabinet. I saw his mother
momentarily pick up her head, roll her eyes, and lower it back down into her hands. She wasn’t crying, just
overwhelmed. It was clear that making a diagnosis might not be as difficult as getting through the session.
Jimmy did have ADHD, with the same symptoms that his father had when he was a boy. ADHD does
sometimes run in families. Apparently, his father still suffered from many symptoms of distractibility and
restlessness that created difficulties in his work and relationships. After many failed career attempts, he had
found considerable success in real estate. The constant movement and transient relationships utilized his
energy and personality, while his choice of a business partner—who excelled at handling the details of his sales
—protected him from his deficits in attention. Being a stable husband and father, however, proved more
problematic.
The treatment for Jimmy included behavioral therapy to help with his attention and social skills, martial
arts classes, and stimulant medications. These and other interventions were designed to boost frontal
functioning through biochemical and behavioral interventions and give him constructive avenues through
which to channel his considerable energy.
Individuals like Jimmy who suffer from ADHD are characterized by an inability to sustain attention and
inhibit extraneous impulses, thoughts, and behaviors. These individuals can be easily lost in daydreams or can
be in constant motion. They are also in danger of leaping before they look. In fact, Jimmy had been injured a
year earlier when he raced into a neighbor’s backyard and jumped into the pool before noticing it had been
drained for repair.
Since Satterfield and Dawson (1971) first pointed to a dysfunction of frontal-limbic circuitry, ADHD has
been understood to be a disorder of descending inhibition and executive control. The common explanation
from psychiatrists to parents is that their children have a lag in frontal lobe development. The results are
inadequate inhibition of impulses and difficulties with tasks that require sustained attention. Parents are also
told that there is a good chance their child will grow out of it as the frontal lobes mature. In the meantime,
stimulant medications will turbocharge these lagging frontal regions, allowing for more functional behavior.
While this is a good anecdotal explanation, the underlying mechanisms and the etiology of ADHD are likely
much more complicated.
Functional imaging comparing ADHD to non-ADHD subjects reveals a variety of patterns of higher and
lower levels of activation throughout the brain. So like most psychiatric disorders, ADHD is heterogeneous
and emerges from a spectrum of genetic, biological, and interpersonal factors (Sun et al., 2005). It is likely
that the explanations of the causes and treatments of ADHD lie within hierarchical networks between the
attentional and inhibitory circuitry of the frontal and parietal cortex, and subcortical networks in the striatum

77
and cerebellum that trigger and organize motor behavior. Thus, it is unwise to localize these deficits solely in
the frontal lobes as these complex behaviors rely on far-reaching circuitry with similar deficits occurring
regardless of the specific location of the lesions (Rubia et al., 2005; Seidman, Valera, & Makris, 2005;
Willcutt et al., 2005).
Stimulant medications (such as Ritalin) may be working on the frontal lobes, the striatum (Vaidya et al.,
1998), the cerebellum (Anderson et al., 2002), or more systemically by boosting general levels of dopamine
and norepinephrine (Arnsten, 2000; Arnsten & Li, 2005). All we can be sure of is that when these
medications work, they rebalance this hierarchical circuitry in a way that decreases motor agitation while
enhancing attention. Because the brain works in interactive networks, the safest working hypothesis at this
point is that there is a problem in the hierarchical neural networks that both activate and regulate behavior
and attention (Durston et al., 2003; Lee et al., 2005; Rubia et al., 1999).
Think of playing a game of Simon Says. Simon Says tests our abilities to respond to the command while
monitoring and inhibiting our behavior based on whether or not Simon says. The winner will be someone
with well-developed, balanced, and integrated bottom-up networks of motor responses and top-down
networks of inhibitory control. When we hear a command in the absence of the words, “Simon says,” we feel
our body react and the tension of inhibition as we exert control to stop ourselves. The popularity of this game
with small children reflects the development of these systems as well as a way to exercise voluntary control
over impulses. When individuals with ADHD engage in tasks similar to Simon Says, they show a lower level
of activity in the usual cortical areas dedicated to inhibition (Durston et al., 2003; Schulz et al., 2004; Zang et
al., 2005).
Children with ADHD have difficulties in organizing their behavior when they are confronted with
situations that require them to inhibit motor responses and sustain attention to continuous or complex tasks.
Thus, they have difficulties in learning skills or activities that require attending to and recalling verbal
material, complex problem solving, and planning. They require much more motivation to maintain attention,
so they often excel at video games, which capture their attention and for which their ability to rapidly shift
attention serves them well. Although our understanding of the brains of individuals with ADHD is limited, a
variety of findings have emerged from research using various imaging techniques (Bush, Valera, & Seidman,
2005). Table 7.4 lists some of the studies that point to an array of differences between ADHD and non-
ADHD individuals using different methods of measurement.

TABLE 7.4
Attention-Deficit/Hyperactivity Disorder

Functional Magnetic Resonance Imaging (fMRI)

Decreased Activation In
Parietal attentional systems1

Anterior-mid cingulate cortex2


Supplemental motor area3

Right middle prefrontal cortex4

Right inferior frontal cortex, left sensorimotor cortex, and bilateral cerebellum lobes and vermis5

Increased Activation In
Left temporal gyrus6

Basal ganglia, insula, cerebellum7

Right anterior cingulate cortex8

Regional Cerebral Blood Flow (rCBF)

Hypoperfusion or Decreased Activation


White matter regions of the frontal lobes and caudate nuclei9

Hyperperfusion or Increased Activation


Right striatum and somatosensory area10

Brain Morphology

Smaller cerebral and cerebellar volume11

78
Smaller right prefrontal and caudate volume12

Reduction of left cortical convolutional complexity in boys13


Cortical thinning in adults in right parietal, dorsolateral, and anterior cingulate areas—all involved with attentional control14
Loss of cerebellar volume15
Decreased frontal and cerebellar white matter density16

The best guess at this point is that individuals diagnosed with ADHD likely reflect a number of subgroups
with different types of neurodevelopmental profiles impacting the size, shape, and function of their brains
(Vaidya et al., 2005). The usual cortical systems of attentional control and inhibition appear compromised
while other networks attempt to compensate. Subcortical structures involved in motor movements are also
affected in ways that result in greater but less organized impact on experience and behavior.
Lastly, I want to mention a phenomenon I have witnessed repeatedly over the years—children who are
diagnosed with ADHD but who are better described as using a manic defense to cope with overwhelming
anxiety (Cozolino, 2014). An assessment of the psychological state of the household including parental
relationship, parental psychopathology, emotional context of siblings and extended family, external stressors,
and so on can go a long way in making a proper diagnosis. Chronic stress and depression can negatively
impact memory, impulse control, and deficits of attention, which can be misdiagnosed as ADHD (Birnbaum
et al., 1999).

Summary
Executive functioning is a complex evolutionary accomplishment that we are still in the process of
understanding. Many brain regions contribute to our abilities to focus, organize our thoughts, regulate our
emotions, and create our experience of self. Head injury, ADHD, and many psychiatric illnesses provide
selective insight into the results of dysregulation of neural networks central to executive processing.

79
Chapter 8
The Executive Brain:
Navigating Space and Time
The soul never thinks without a mental picture.
—Aristotle

Because executive problems often arise after damage to the frontal areas and because these regions are
especially vulnerable to injury, it has generally been assumed that our internal executive resides right above and
behind our eyes. We can be somewhat confident that consciousness emerges from the coordination and
cooperation of many neural networks throughout the brain and that the frontal lobes are a major contributor.
But as we saw in Chapter 7, the cingulate and insula cortices also contribute to focused attention, a key
element of executive functioning. Here, I suggest that the parietal lobes are another major contributor to
executive functioning.
This thought first occurred to me many years ago when I discovered that studies of primate brain
evolution suggested that the expansion of the parietal, not the frontal, lobes was most characteristic of the
transition to the human brain (von Bonin, 1963). Could the fact that we don’t think of the parietal lobes as a
component of the executive brain reflect a cultural bias of equating individuals with their external behavior
rather than the quality and nature of their inner experiences? If neurology had been born in a Buddhist
country, would our view of executive function be different?

The Parietal Lobes


Logic will get you from A to B. Imagination will take you everywhere.
—Albert Einstein

You may remember that the parietal lobes evolved from the hippocampi which, in lower mammals, organize a
three-dimensional map of the external environment (Joseph, 1996; O’Keefe & Nadel, 1978). This allows
animals to navigate a habitat for finding, storing, and retrieving food, as well as keeping track of the kids. The
hippocampi of mother rats actually increase in size when they have babies in preparation for having more
mouths to feed. The human hippocampi continue to serve this function, as evidenced by the fact that they are
larger in cab drivers (Maguire, Woollett, & Spiers, 2006). The parietal lobes expand these functions by adding
a map of internal space and maps of our bodies, and our bodies in external space (Husain & Nachev, 2007). In
addition to the right hippocampus and inferior parietal lobe, the medial parietal regions on both sides of the
brain contribute to spatial navigation.
We measure sensory-motor development through the unfolding of increasingly precise and useful
behaviors such as opening a box of cookies and pouring a glass of milk. Since our brains evolved to adapt
within a four-dimensional environment (three-dimensional space plus time), executive functioning emerged as
a system of spatial navigation. It is within this four-dimensional context that the sensory, motor, and spatial
components of our attachment systems are formed. Think for a moment about how we describe our inner
emotions—We fall in love, fly into a rage, or have a hard time handling bad news about a friend. Our language
demonstrates the fact that our experience is grounded in moving through space.
The parietal lobes, in connection with the rest of the cortex, allow for the integration of working visual
memory, attentional capacities, and bodily awareness necessary for these imaginal abilities. Athletes take
advantage of these networks when they imagine making a difficult catch or working on their tennis serve. This
suggests that our self-awareness was likely built in a stepwise manner during evolution through a series of
overlapping maps of the physical environment, of the self in the environment, and later of self as an
environment. Thus, the growth of imaginal abilities allowed us to create an increasingly sophisticated inner

80
topography for self-reflection and expanded awareness.
The lower parts of the parietal lobes develop through the first decade of life in parallel with our increasing
abilities related to reading, calculations, working memory, and three-dimensional manipulation (Joseph, 1996;
Klingberg, Forssberg, & Westerberg, 2002; Luna, 2004). Mirror neurons in these inferior parietal regions
respond to the hand position, eye movement, words, motivational relevance, and the body position of others
as we use these individuals to model how to navigate four-dimensional space.
Left parietal damage disrupts mathematical abilities while damage to the right results in disturbances of
body image and the neglect of the left side of the body. Despite these debilitating symptoms, patients are
either oblivious to or deny the significance of their deficits, suggesting that the parietal lobes contribute to the
organization of self-awareness. Damage to the parietal lobes disrupts the experience of location, self-
organization, and identity—in other words, where, what, and who we are (see Table 8.1).
The posterior parietal regions weave together sensory information with motor actions to organize goal-
directed action plans, another executive function (Andersen, Snyder, Bradley, & Zing, 1997; Colby &
Golberg, 1999; Medendorp, Goltz, Crawford, & Vilis, 2005). In combination with episodic and working
memory, these regions provide an inner work space for decision making and deciding whether or not to
perform an action. Utilizing these abilities, frontal-parietal networks support the integration of perception and
action over time (Quintana & Fuster, 1999; Shomstein, 2012).

TABLE 8.1
Manifestations of Parietal Compromise

Left Parietal Compromise Results in


Gerstmann syndrome, which includes the following symptoms:
Right-left confusion
Digital agnosia (inability to name the fingers on both hands)
Agraphia (inability to write)
Acalculia (inability to calculate)1
The symptoms of Gerstmann syndrome are linked through a unitary deficit in spatial orientation of body—sides, fingers, and numbers2

Right Parietal Compromise Results in Deficits of

Mental imagery and movement representations3 Detecting apparent motion9


Visual-spatial awareness4 The analysis of sound movement10
Visual-spatial problem solving5 Spatial-temporal abnormalities11
Temporal awareness and temporal order6 Contralateral neglect of the body and external space12
Spatial perception7 Denial of hemiparalysis and neglect13
Somatosensory experience8

The parietal lobes have been shown to participate in our conscious awareness of visual experience,
voluntary actions, and a sense of agency during our actions (Chaminade & Decety, 2002; Decety et al., 2002;
Rees, Kreiman, & Koch, 2002; Sirigu et al., 2003). The multimodal representation of four-dimensional space
in the posterior parietal areas integrates our goal-directed behavior and attention with higher cognitive
functions (Andersen et al., 1997; Bonda, Patrides, Ostry, & Evans, 1996; Corbetta & Shulman, 2002;
Culham & Kanwisher, 2001). If we act, we act to meet a goal. Goal-directed action reflects our movement
through the four dimensions of space and time.
Along with the frontal, cingulate, and insula cortices, areas of the parietal lobes become activated by
novelty and appear to be involved in coding intentions and calculating the probability of success (Platt &
Glimcher, 1999; Snyder, Batista, & Andersen, 1997; Walsh, Ashbridge, & Cowey, 1998). These findings
point to the fact that the parietal lobes are involved in the deployment of attention, understanding the
environment, and the experience of self (Brozzoli, Gentile, & Ehrsson, 2012) (see Table 8.2).
The medial parietal area can be conceptualized as the central structure for self-representation, self-
monitoring, and a state of resting consciousness (Lou et al., 2004). Damage at the junction of the parietal and
temporal lobes can result in out-of-body experiences and a variety of other disturbances of identity and self
(Blanke & Arzy, 2005). There is also evidence to suggest that the parietal lobes participate in the creation of
internal representations of the actions of others within us (Cabeza, 2012; Shmuelof & Zohary, 2006). In other
words, we internalize others by creating four-dimensional representations of them in our imaginations,
allowing us to both learn from others and carry them within us. These inner objects, as described in
psychoanalysis, likely serve as a cornerstone of learning emotional regulation, and the construction and

81
maintenance of our sense of self and others (Macrae et al., 2004; Tanji & Hoshi, 2001).

TABLE 8.2
Functions of the Parietal Lobes

Hemisphere Function
Right Analysis of sound movement1
General comparison of amounts2
Attention3
Self-face recognition4

Left Verbal manipulation of numbers5


Mathematics6
Multiplication7
Motor attention8

Bilateral Findings
Visual-spatial work space9 Controlling attention to salient event and maintaining attention
across time17

Visual-spatial problem solving10 Preparation for pointing to an object18


Visual motion11 Grasping19

Construction of a sensory-motor representation of the internal world in Movement of three-dimensional objects20


relation to the body12
A sense of numerosity, defined as nonsymbolic approximations of
Internal representation of the state of the body13 quantities (l)21
Verbal working memory14
Retrieval from episodic memory15 Processing of abstract knowledge22

Sequence and ordering of information in working memory16


Perspective taking (r)23
Processing of social information (r)24
Taking a third-person perspective (r)25

(l) left hemisphere; (r) right hemisphere

Parietal activation across a wide variety of cognitive tasks strongly suggests that the coordination of sensory
and motor processing lies at the core of our abstract thinking (Culham & Kanwisher, 2001; Jonides et al.,
1998). It is likely that evolution has conserved and built upon these primitive visual-spatial networks to serve
as an infrastructure for language and higher cognitive processes (Klingberg et al., 2002; Piazza et al., 2004;
Simon et al., 2002). This may be why occupational therapy with children who struggle with motor
coordination often results in cognitive, linguistic, and social benefits.
Johnson (1987) asserts that the experience of our bodies provides the internal basis for our sense of
numbers, quantity, and space. The brain’s ability to take our physical experience and use it metaphorically for
abstract thinking serves as the basis of imagination. For example, going down a slide may serve as a sensory-
motor metaphor for falling in love. The child’s experience of emerging from under the covers into the light of
day provides a metaphor for religious enlightenment later in life. The balance provided by the vestibular
system may be the internal working model for emotional stability and a more balanced life (Frick, 1982).
Physical metaphors provide a contextual grounding in time and space that grounds our experience. Many
believe that cognition is embodied and that bodily experiences serve as the infrastructure for higher cognitive
processes (Koziol, Budding, & Chidekel, 2011).
Albert Einstein, who did poorly in math during his formal education, went on to answer some of physics’
most difficult questions. The relationships he intuited between time, matter, and energy brought us closer to
understanding the workings of the universe. As you can imagine, many neuroscientists were interested in
having a look at Einstein’s brain to see if and how it differed from yours and mine. In comparison to 91 other
brains, Einstein’s was different only in the size of the inferior parietal lobe (Witelson, Kigar, & Harvey, 1999).
A subsequent examination of the same region revealed lower ratios of neurons to glial cells when compared to
other areas of Einstein’s brain as well as to the brains of other people (Diamond et al., 1966; Diamond,

82
Scheibel, Murphy, & Harvey, 1985). It is highly likely that this larger parietal region and enhanced neural-
glial relationship led to Einstein’s superior visual-spatial and abstract abilities (Nedergaard et al., 2003;
Oberheim et al., 2006; Taber & Hurley, 2008).
These neuroanatomical findings are especially interesting in light of Einstein’s reported use of mental
imagery to solve complex mathematical problems. He described translating numerical equations into images
that he would manipulate in his imagination, solve visually, and then translate back into equations. It is within
one of his major findings, E = mc2, that he discovered the inextricability of space, matter, and time. This
ability to conceptualize and manipulate three-dimensional objects imaginally appears to separate us from other
primates and may be a uniquely human evolutionary accomplishment (Orban et al., 2006; Vanduffel et al.,
2002). Based on his description of his problem-solving strategies and the findings concerning his brain, it
appears possible that Einstein’s unusual parietal lobes may have been central to his genius.
Einstein was not without his deficits, and his difficulty in navigating the simple demands of day-to-day life
was notorious, making him the archetypical absent-minded professor. But the external world is more of a
focus of the frontal cortex. Interestingly, one study has shown that the volumes of the frontal and parietal
lobes demonstrate a significant negative correlation (Allen, Damasio, & Grabowski, 2002). In other words,
we may be biased in the direction of one or the other. Being absent-minded may have been the price Einstein
paid for an overdeveloped parietal lobe. Research also suggests that inner imaginal space enhances the
possibility for creative problem solving, empathy, and compassion. Perhaps this is one of the reasons that
Einstein turned his attention to world peace and other humanitarian concerns later in life.

A Frontal-Parietal Executive Network


I never came to any of my discoveries through the process of rational thinking.
—Albert Einstein

Explorations of the functional neuroanatomy of human intelligence suggest that it is distributed far beyond
the frontal lobes (Nachev, Mah, & Husain, 2009; van den Heuvel, Mandl, Kahn, & Pol, 2009). The complex
functional connectivity of proximate and distal brain regions have been linked and choreographed over eons of
natural selection. We are only beginning to appreciate how each area of the brain contributes to executive
functioning (Bartolomeo, 2006).
While most neuropsychologists still consider the frontal lobes as the seat of our general intelligence, new
findings suggest that the domains of intelligence measured by standardized testing are mediated via frontal-
parietal networks and the neural fibers that connect them. Put in a slightly different way, measures of
intelligence are correlated with how efficiently our brains integrate and synthesize information via frontal-
parietal cooperation. Fascinated by this problem, Jung and Haier (2007) reviewed studies examining the
relationship between focal brain damage and performance on the Wechsler Adult Intelligence Scale (WAIS).
Analysis was performed on the four broad factors that emerge from the WAIS: verbal comprehension,
perceptual organization, working memory, and processing speed.
Statistically significant lesion–performance deficit relationships were found for three of the four factors:
verbal comprehension (left inferior frontal cortex), working memory (left frontal and parietal cortex), and
perceptual organization (right parietal cortex). Based on this research, Jung and Haier came to the conclusion
that what we consider to be intelligence includes the frontal and parietal lobes. Their work led them to
propose the parieto-frontal intelligence theory of a network that includes the dorsolateral prefrontal cortex,
the inferior and superior parietal lobes, the anterior cingulate, some regions within the temporal and occipital
lobes, and their connecting white matter tracts (Colom et al., 2009; Costa & Averbeck, 2013; Jung & Haier
2007; Langer et al., 2012; Smith et al., 2011).
This network may also contain an orienting system with a right frontal ventral network that interrupts and
resets ongoing activity and a dorsal frontal network for matching stimuli and responses (Corbetta, Patel, &
Shulman, 2008; Hu et al., 2013). The frontal and parietal lobes work together with the salience network to
analyze the context and location of specific aspects of the environment and interrupt ongoing behavior in
order to direct attention to new targets (Bush, 2012; Corbetta & Shulman, 2002; Peers et al., 2005).
Around the same time, Anderson and his colleagues (2008) focused on essentially the same set of
structures they believed to be responsible for the control of rational thought. They named the lateral inferior
prefrontal region (information retrieval), posterior parietal region (representation in imagination), anterior
cingulate (information update related to goals), and the caudate nucleus (for the execution of behaviors) as key
processing hubs for intelligence and executive functioning. These two models are backed up by other research,

83
which points to the parietal-frontal circuit being involved in executive semantic processing as well as deciding
and planning goal-directed actions (Andersen & Cui, 2009; Buneo & Andersen, 2006; Crowe et al., 2013;
Goodwin, Blackman, Sakellaridi, & Chafee, 2012; Whitney, Kirk, O’Sullivan, Ralph, & Jefferies, 2012). A
similar architecture for attending, looking, and pointing has also been discovered in macaque monkeys
(Astafiev et al., 2003).
The volume of frontal and parietal gray matter, and the white matter tracts that connect them,
demonstrate the highest correlation with intelligence (Glascher, Hampton, & O’Doherty, 2009; Gray et al.,
2002; Haier et al., 2004; Langeslag et al., 2012; Lee, Josephs, Dolan, & Critchley, 2006). Like most other
central processing networks, their white matter connectivity continues to mature into adulthood (Klingberg et
al., 2002; Klingberg, 2006; Olesen, Nagy, Westerberg, & Klingberg, 2003). This distributed region of the left
hemisphere is responsible for skills measured by the WAIS that include verbal knowledge, verbal reasoning,
working memory, cognitive flexibility, and executive control (Barbey et al., 2012).
Frontal-parietal circuits are also involved in the sustained focus and updating of information in working
memory (Edin et al., 2007; Salazar, Dotson, Bressler, & Gray, 2012; Sauseng et al., 2005). Frontal-parietal
networks work together to analyze the context and location of specific variables, interrupt ongoing behavior,
and direct attention to new targets (Coull, Cotti, & Vidal, 2014; Corbetta & Shulman, 2002; Genovesio,
Wise, & Passingham, 2014; Peers et al., 2005). Neural fibers connecting the middle portions of the frontal
and parietal lobes appear to serve a general integrative function of linking right and left hemispheres, limbic
and cortical structures, and anterior and posterior regions of the cortex, possibly giving rise to a global work
space or central representation that allows for conscious working memory and self-reflection (Baars, 2002;
Cornette et al., 2001; Fedorenko, Duncan, & Kanwisher, 2013; Lou et al., 2004; Taylor, 2001). This
neurological reality reflects that our experience of space and time are inextricably interwoven. Because of this,
the frontal-parietal network may be a necessary prerequisite for the construction of the experience of self
(Caspers et al., 2012; Lou, Nowak, & Kajer, 2005).
A properly functioning frontal-parietal network allows for the successful negotiation of our moment-to-
moment survival and the ability to turn our attention to inner experience. A compromised or poorly developed
prefrontal cortex can ensnare us in “a noisy and temporally constrained state, locking the patient into the
immediate space and time with little ability to escape” (Knight & Grabowecky, 1995, p. 1368). Without the
ability to reflect on and sometimes cancel reflexive motor and emotional responses, there is little freedom
(Schall, 2001). A similar phenomenon can occur in states of chronic anxiety when victims become “stuck” to
the environment or “stimulus bound” and are unable to override reflexive fear reactions (Brown et al., 1994).

The Role of Mirror Neurons in Executive Functions


A human being only becomes human at all by imitating other human beings.
—Theodor Adorno

In the 1990s, researchers discovered that neurons in the frontal lobes of primates fire when another primate
(or the experimenter) is observed engaging in a specific behavior such as grasping an object with a hand. It
turned out that the same neurons fire when the subject engages in the identical task. Because these neurons
fire when the subject is both observing and performing a particular action, they were dubbed mirror neurons (di
Pelligrino et al., 1992; Gallese, 2001; Gallese et al., 1996; Gallese & Goldman, 1998; Jeannerod et al., 1995;
Rizzolatti et al., 1999). Some mirror neurons were found to be so specific that they only fired when a
particular object was grasped in a certain way by specific fingers, such as picking up a banana with the right
hand at a certain angle or peeling it with the thumb and forefinger (Rizzolatti & Arbib, 1998).
These neurons are all the more interesting because they are sensitive to both sensory information and goal-
directed behaviors. They don’t fire in response to the hand or the banana or even the two presented together,
but only when the hand is acting on the banana in a specific way for a specific purpose, a neural reflection in
the mirror of space-time we have been exploring. Mirror neurons found thus far in monkeys focus on actions
of the mouth and hands, suggesting that their original role may have involved getting food to the mouth and
eating (see Table 8.3). Through the course of evolution, they have expanded to include gestures, oral
communication, and language. This is demonstrated when we yawn after seeing someone else yawn or when
our lips move in response to watching someone eat.
Mirror neurons have since been found in the parietal cortices as well (Kilner et al., 2004, 2009). Mirror
neurons lie at the crossroads of inner and outer experience, where multiple networks of visual, motor, and
emotional processing converge within our executive systems (Iacoboni et al., 2001). Because of their privileged

84
position, mirror neurons are able to bridge observation and action and tap into maps of other people and the
space around us, as well as our own musculature, survival needs, and strategies for attaining goals.

TABLE 8.3
Brain Regions With Mirror Neurons

Frontal Lobe, Specifically the Inferior Frontal Gyrus


Dorsal Motor actions of the hand1
Ventral Ingestive (85%) Chewing/sucking
Communicative (15%) Lip movements/licking
Posterior Oro-facial and oro-laryngeal movements

Functions
Imitation,2 gestures3
Integration of somatic, sensory, motor systems with planning and motivation
Sequencing of motor goals and anticipation of the consequences of actions4

Action understanding and action monitoring


Semantic representation of action
Expressive language

Parietal Lobe, Specifically the Inferior Parietal Lobe

Functions
Synthesis of visual-spatial organization and somatosensory information
Pragmatic analysis of objects5
Linking objects, actions, and goals6

Global representation of action patterns of motor behavior and limb position7

Over the last 20 years, mirror systems have helped us to understand how our brains link together in the
synchronization of such group behaviors as hunting, dancing, and emotional attunement (Jeannerod, 2001).
In addition, they may also likely be involved in the learning of manual skills, the evolution of gestural
communication, spoken language, group cohesion, and empathy.

Navigation and Attachment


There is more wisdom in your body than in your deepest philosophies.
—Frederick Nietzsche

Although we usually focus on the emotional and cognitive aspects of attachment schemas, attachment is the
regulation of positive and negative bodily states through proximity to caretakers. Our brains accomplish
attachment schemas by linking good feelings via oxytocin, dopamine, and serotonin with proximity to good
attachment figures and associating bad feelings with separation from them (Feldman, 2012). We measure
attachment by watching how a child responds to the return of the mother after being left with a stranger.
Does the child approach, avoid, ignore, or act in dysfunctional ways in reaction to the mother’s return? At
least two levels of analysis are taking place in attachment: the first is whether to engage, and the second is how
to engage.
This implies that attachment schemas are stored in sensory and motor systems and become manifest in our
musculature, posture, gait, and interpersonal stance, not just cognitive-emotional constructs. We generally
think of attachment schemas in terms of mental states (thoughts, emotions, and beliefs). However, given that
they exist in primitive animals and infants, it might be more helpful (and perhaps more accurate) to think of
attachment patterns as procedural memories related to approach-avoidance learning activated in close
relationships. Another way to describe this is “motor cognition,” a way to understand both the evolution and
development of how we came to learn from and about others through observation and internal simulation
(Gallese, Rochat, Cossu, & Sinigaglia, 2009).

85
Affordance
The abstract analysis of the world by mathematics and physics rests on the concepts of space and time.
—James J. Gibson

Given that movements are goal oriented, it makes sense that neural networks controlling movement and those
organizing goal-directed behavior would be interwoven in our brains (Rizzolatti & Sinigaglia, 2008). As we
navigate our worlds, our brains automatically generate myriad options, paths, and potential strategies based on
an interaction between our current environment and our learning history. In other words, when we encounter
something or someone in our environment, our brains activate preexisting implicit memories, which provide
us with options for engagement. These systems create affordance, our ability to engage meaningfully with the
objects and people around us. Affordance is neither objective nor subjective, but is activated within the
interface of the person and the social and physical environments. So, if our ability to navigate toward an
attachment figure leads us to approach, how do we engage with that figure?
While our systems of navigation get us where we want to be, neural systems allow us to utilize affordance,
be it a rock formation, a tool, or another person. Affordance is the flip side of ergonomics, which involves
making usable tools such as work spaces, airplane cockpits, or scissors. The focus of ergonomics is to adapt
products to fit us. Affordance works in the opposite direction; it’s our adaptation to the world. Affordance
reflects our ability to respond with sensory, cognitive, motor, and emotional responses in useful ways. When
we know how to use a tool or interact with someone, the tool or the person becomes incorporated as an
extension of our body (Calvo-Merino et al., 2005; Cattaneo et al., 2009). Affordance determines our ability to
grasp and utilize what is in front of us.
Affordance strategies are unconscious, multimodal memories that are automatically activated by situations
that we encounter in our day-to-day lives. A straightforward example might occur when we are sitting in a
café talking with a friend and the waiter sets before us a cup of coffee, sugar, cream, and a spoon. Our implicit
procedural memories allow us to use our hands automatically to engage these objects such that we are able to
mix our personal blend of ingredients while not missing a word of the conversation.
In Moore’s (1986) study of how children interact with natural environments, he discovered that children
tend to focus on specific features with which to interact. Some examples were smooth surfaces for running,
things for climbing (trees), places for hiding (bushes), slopes for sliding down, obstacles for jumping over, and
objects for throwing. The ability and style with which children engaged with these features of the English
countryside can be described as affordances. It is not coincidental that the salient environmental features
related to their play are historically relevant for survival (children in Africa may use trees more as a source of
shade). These were the same features that children needed to learn during childhood in order to hunt for food
and escape from danger later. The same skills, once needed for basic survival, are now applied to sports and
games requiring throwing, jumping, and kicking.
Because the notion of affordance has emerged from the fields of ecology and environmental studies, it is
usually conceptualized in terms of how we learn to interact with our physical environment and not other
people (Heft, 1989; Kytta, 2002). We like to think of the connection of two people as the linking of minds
and hearts, as opposed to utilizing them. While more sophisticated types of relationships occur later in life,
early attachments are shaped largely by physical proximity, interactions, and children using their parents for
survival. As adults we engage with our friend utilizing language, facial expressions, eye contact, and a wide
range of communication channels to be able to “afford” our relationship.
Can we expand the idea of affordance to attachment relationships? Imagine sitting with a friend and his
four-year-old son, paying special attention to their many interactions. When we first enter the house, the
child holds his father’s leg, leans into it, and rolls around the back to watch us from a safe vantage point. A
little later he presents his finger, which he has pinched in the door of a toy car, for his father to kiss. At
another point, he sneaks up from behind with a pillow and hits his father on the head as an invitation to some
rough and tumble play. Still later, he finishes his juice box and hands it to his father with the words “all done.”
These interactions demonstrate the boy’s ability to successfully use his father for safety, solace, stimulation,
and service. These affordances are driven by mutual instinct and reinforced via positive feelings related to
bonding and attachment.
Because parent-child interactions eventually shape our affordance patterns with others, a boy like the one
described above may enter school with the expectation of building the same kind of positive connection with
his teacher. He may assume that the teacher will be a source of safety, build on the affordances he has learned,
and build new affordance patterns more appropriate to the classroom. This may include a calm and relaxed
body, leaning toward the teacher when something interests them, curiosity about what is being discussed, and

86
optimism about being able to master the material. This pattern of classroom affordance allows them to sit
quietly, focus on the lessons, and learn the material being presented, maximizing neural plasticity and learning
in the student and enthusiasm in the teacher. None of this is a revelation until you consider that many
children, deprived of available and attuned caretakers, never learn how to “afford” positive relationships until
they enter the classroom.

Building an Internal World


The wise man knows how to run his life so that contemplation is possible.
—Gabriel Marcel

One of evolution’s greatest accomplishments is our ability to create an internal world that allows us to engage
in private thought, self-reflection, and imagination. Our inner worlds can then serve as the grounds for
mentalization, creativity, and consolidating our sense of self (Winnicott, 1965a). Individuals with severe brain
injuries in their frontal and parietal regions can be constantly distracted by sensory and emotional experience,
finding it impossible to maintain focus and utilize their imaginations. These individuals can become trapped
within the present moment, unable to disengage from the constant stream of sensations, emotions, and the
demands of their outer worlds. Although they retain consciousness, many victims of brain injury experience
difficulties in attention, concentration, affect regulation, and motivation.
Being an introvert, I am well acquainted with my inner world. As a child I had an imaginary retreat. I
would close my eyes and picture the back of my grandmother’s closet, always piled high with shoe boxes.
Behind these boxes was a hidden door just large enough for me to squeeze through. Once through the door,
there was a flight of stairs leading up to a medieval laboratory, the kind a scientist or sorcerer might have at
the top of a castle. This was a safe place for me—quiet and private—where I could imagine other worlds,
reflect on life, and fantasize about the future. This is where I would take my science fiction books only to
discover many others before me had their own version of my laboratory.
Winnicott (1965b) suggested that one’s sense of self consolidates during the periods of quiescence when
we feel safe and calm in the presence of our parents. Good-enough parenting scaffolds a child, allowing him
or her to go inside and rest in imagination and the experience of self (Stern, 1985). It is rare to find a child
who is able to be still, centered, and safe in the presence of chaotic adults. Positive early caretaking builds and
shapes the cortex and its relationships with the limbic system, which supports emotional regulation,
imagination, and coping skills. To this we now must add the development of neural networks that allow us to
construct internal space.
One study has shown that when experienced meditators engage in meditation, the frontal lobes become
less active while the parietal lobes become more active, reflecting perhaps a shift from outer to inner attention
and from right- to left-hemisphere activation (Davidson, Kabat-Zinn, et al., 2003; Newberg et al., 2001).
Interestingly, inferior regions of the right parietal lobe become activated when we witness others being still.
This may explain how meditating on inanimate objects or statues of a tranquil Buddha may help us feel
centered within ourselves (Federspiel et al., 2005). We can also internalize calm parents and therapists as a
model for affect regulation and self-reflection. Perhaps there are mirror neurons for imitating a state of
centered calm.

Summary

Working together, the frontal and parietal lobes and the mirror neurons contained within them make it
possible for us to direct our behavior in ways that support our survival. Executive functioning requires the
creation of a four-dimensional model of the world. It also necessitates a set of skills that allows us to navigate
the world and utilize the things within in that we need to survive. With the help of the anterior cingulate and
insula, we are drawn to and away from things that could help or hurt us. This government of networks also
creates the possibility of both inner space and the safety to withdraw from the outside world to a primitive,
inner space. As we grow, we need to construct and learn what is within this inner world; we first create our
experience of others and later in life come to find ourselves.

87
88
Chapter 9
The Executive Brain:
Discovering Others
and Finding the Self
People are accustomed to look at the heavens and to wonder what happens there. It would be better if they
would look within themselves.
—Kotzker Rebbe

Creating an inner world requires more than a four-dimensional inner space and disengaging from the external
world. It also requires that we are able to sustain focus on our bodies and in our imaginations. Until very
recently, this ability, so prized in the Eastern world, has been mostly ignored in the West. In the search for
brain-behavior relationships, research subjects have always been tested in the act of performing a task. It rarely
occurred to researchers to see what our brains are up to when we are not engaging with the external world.
In 1929, Hans Berger (the inventor of the electroencephalograph) noticed that the brain remained active,
even when his subjects were at rest. This finding was ignored for 70 years until researchers observed that a
consistent set of brain regions showed a predictable decrease in activity during tasks but became active in the
absence of attention to these same tasks (Raichle et al., 2001; Shulman et al., 1997a, 1997b). These areas
included the midline regions of the posterior cingulate cortex, the medial prefrontal cortex, and the precuneus
region of the parietal cortex. The name most often used for this new system is the default-mode network
(DMN) (Northoff & Bermpohl 2004; Northoff et al., 2006). Subsequent research has established the DMN
as a consistent and coherent functional network similar to those dedicated to visual and motor processing
(Beckmann et al., 2005; Beason-Held et al., 2009; Greicius et al., 2009; Smith et al., 2009; Weissman et al.,
2010), supporting Berger’s original observation that our brains are always at work.

The Default Mode Network


The hardest victory is over the self.
—Aristotle

The addition of the DMN to our conception of executive function gives us a potential neural home for
internal thought and private reflection, abilities often overlooked as aspects of executive functioning. The
DMN appears to assist in consolidating the past and preparing us for the future (Buckner & Vincent, 2007).
The increased activation and connectivity between the posterior and anterior regions of the cingulate cortex
during resting states are also thought to support the integration of cognition and emotion (Greicius et al.,
2003). It may serve as a psychological baseline for subjective experience as well as a platform for social
cognition (Mars et al., 2012; Schilbach et al., 2008). Further, the resting state most likely provides us with the
ability to sustain focus on relationships, take the perspective of others, and imagine future outcomes of
interpersonal behaviors.
Each region of the DMN is made up of subregions that perform multiple processes by cooperating with
other regions of the brain (Leech et al., 2011; Salomon et al., 2013). The dorsal and ventral areas of the ompfc
show decreased activation during goal-directed activities but become active (along with the anterior cingulate)
during self-generated thought, self-reference, the experience of emotions, and attributing mental states to
others (Frith & Frith, 1999, 2010; Gusnard et al., 2001; Gusnard & Raichle, 2001; Lane et al., 1997; Reiman
et al., 1997). The medial areas of the prefrontal cortex have been established as central to the social and
emotional experience of self and demonstrate a higher metabolic rate in the absence of external demands than
most other brain regions (Beer, 2007). These patterns of activation parallel our subjective experiences when

89
our attention to others’ minds as well as our own are heightened, and we decrease our focus on external tasks.
See Table 9.1 for a list of DMN regions and their hypothesized contributions to human experience.

Table 9.1
DMN Regions and Their Contributions
Medial prefrontal cortex Self-relevant mentation, theory of mind1
Medial temporal lobe Autobiographical memory
Hippocampus Associations with past experiences2
Posterior cingulate cortex Sensory integration
Parietal cortex Self-awareness, self-other comparisons
Precuneus region Memory source attribution, internal mentation,
Visual-spatial organization and coordination3

Together, the regions of the DMN appear to synthesize a sensory experience of the body and internal
world that allows us to have a conscious experience of ourselves in imaginal space within the flow of time.
This function, so vital to the experience of self and the development of culture, suggests that the adaptive
functions of the DMN have exerted a strong selection bias on both neural and sociocultural evolution
(Morcom & Fletcher, 2007).

Functions of the DMN


A friend is, as it were, a second self.
—Cicero

The DMN is inhibited during the encoding of episodic memory and active during retrieval and recall (Chai et
al., 2014). Given its neural connectivity to the temporal lobes and hippocampus, the DMN is likely involved
in the retrieval of episodic memory (Maddock et al., 2001; Fujii et al., 2002; Cabeza et al., 2002). On the
other hand, if the DMN is active during memory encoding, it may trigger old memories that disrupt the
establishment of new memories. This may be a DMN mechanism that is disrupted in PTSD, which interferes
with the updating and reorganization of traumatic memories.
The DMN and the frontal-parietal networks that attend to external executive functioning are generally
anticorrelational, meaning that when one is active, the other is inhibited (Anticevic et al., 2012; Elton & Gao,
2014). When we are in a resting state, DMN activation will continue, and the external executive will stand
down until they are triggered to take over (Chen et al., 2013). The signal for this changing of the guard will
likely come from the salience network when it detects a novel stimulus or a possible threat (Fransson, 2005;
Sridharan, Levitin, & Menon, 2008). Research suggests that the anticorrelation isn’t always complete, as there
appear to be external tasks that require some DMN participation (Greicius & Menon, 2004; Laird et al.,
2009; Singh & Fawcett, 2008).
When we engage in outwardly focused activity, the amount of inhibition of the DMN varies with task
demand and our degree of engagement with the task (Greicius & Menon, 2004; McKiernan et al., 2003;
Pfefferbaum et al., 2011; Uddin et al., 2009). Tasks like psychotherapy may require a more subtle balance of
internal and external focus in order to monitor, facilitate, or influence cognition (Hampson et al., 2006). One
study shows cooperation between the frontal-parietal network and the DMN when maintaining a focused,
internal train of thought (Smallwood et al., 2011). This point of interaction between internal and external
processing may be an avenue for the influence of autobiographical memory on current experience—a central
neural mechanism underlying projective processes. DMN participation during outside interactions may also
be responsible for a client’s ability to simultaneously be engaged in a transference interaction and have insight
into the fact that they are actively projecting onto the therapist (see Table 9.2).
Stimulus-independent thought and detachment from the external environment is likely to be a prerequisite
to imagination. The medial prefrontal regions are capable of inhibiting the amygdala along with task-oriented
networks to allow the posterior cingulate cortex and parietal lobes to provide us with a platform for internal
mentation and a three-dimensional work space. This allows us to manipulate images, play out interpersonal
scenarios, and process emotions without external interference. It also offers us the possibility of experiencing
time travel through the juxtaposition in working memory of the past, present, and future. What we call mind
wandering and daydreaming may be both the soil of creativity and the background processing that helps us

90
function in the world (Baird et al., 2012).

TABLE 9.2
Hypothesized Functions of The Default Mode Network

Self-Awareness
Conscious awareness1 Alterations in states of awareness6
Self-relatedness2 Self-awareness7
Self-reflection3 Autobiographical memory8
Self-referencing4 Sense of self9
Self-other distinctions5

Social Awareness
Processing social relationships10 Recognizing the faces of others12
Theory of mind processing11 Processing moral dilemmas13

Perception and Cognition


Visual fixation14 Resting with eyes closed20

Passive perception15 Spontaneous semantics21


Mnemonic processes16 Stimulus-independent thought22

Implicit processing17 Mental time travel23


Memory scene construction18 Navigation24

Semantic memory construction19 Prospection25

Given its slow development and the range of individual differences in the functions attributed to the
DMN, it is extremely likely that its growth and development are shaped by experiences (Pluta et al., 2014;
Sambataro et al., 2010; Supekar et al., 2010). Early research does suggest a range of differences in its neural
structures and functional connectivity across individuals and during development from infancy to later
adulthood.
The nascent DMN has been detected as early as two weeks after birth, but only forms into a coherent
network around age seven (Fair et al., 2008; Gao et al., 2009). This is around the time that children develop a
sense of separateness from those around them. The DMN remains immature through childhood, but begins
to become functionally coherent during early adulthood. At the same time the medial and dorsal regions of
the frontal lobes also shift from correlational to anticorrelational with one another as we learn to separate
internal and external experience (Anderson et al., 2011; Chai et al., 2014; Thomason et al., 2008). Women
with early life trauma have been shown to exhibit decreased functional connectivity within the DMN (Bluhm
et al., 2009). In fact, DMN connectivity in adults with childhood maltreatment parallels patterns observed in
healthy seven- to nine-year-olds, suggesting that trauma may delay or impair DMN development.

The DMN and the Self


True nobility lies in being superior to your former self.
—Ernest Hemingway

In discussing the DMN, it appears that we have been describing a key aspect of the neural substrate of the
self. An area that becomes active during tasks of autobiographical memory, self-reflection, and memory of the
self suggests that the DMN is key to the experience of self in general. In addition, the DMN appears to be
just as involved with key aspects of other-consciousness such as face recognition, theory of mind, and
processing moral dilemmas. Given the overlap, it is fair to ask the question: what came first, self- or other-
consciousness?
As adults, we generally think of ourselves as being self-aware first and attending to others second.
However, many other primates demonstrate DMN functioning and good theory of mind skills in the absence
of what we consider self-awareness (Mantini et al., 2011). We also see that a child’s developing sense of self

91
emerges from a symbiotic connection to mother and family. Thus, the truth may be the opposite of common
sense; other-awareness may have evolved first because of its primitive survival value, later becoming the
infrastructure of the experience of self. This would certainly parallel psychodynamic ideas such as internal
objects, transference, and projective identification.
Just as regions of the frontal and parietal lobes join together to construct our experience of space-time, the
ompfc and medial precuneus regions of the parietal lobes may construct a unitary experience of self-other. In
other words, there may be no self in the brain fully separate from our experience of others. This is especially
evident in individuals who exhibit poor interpersonal boundaries and are labeled as borderline and dependent
personalities. All of this suggests that the experience of self is an emergent function from a complex
combination of everything from a sense of bodily continuity through time to our sophisticated coconstructed
narratives that organize everything from our social identities and defense mechanisms to our spiritual
aspirations.

The DMN and Psychopathology


I know how men in exile feed on dreams.
—Aeschylus

Many clients we see in therapy do not appear to have an internal world to which they can retreat for solace
and self-reflection. The functions of the DMN give us a way to think about what may and may not be going
on in these clients’ brains. Disruptions in DMN functioning have been found in individuals with autism,
dementia, depression, schizophrenia, stroke, and chronic pain (Balthazar et al., 2014; Lynch et al., 2013; Park
et al., 2014; Whitfield-Gabrieli et al., 2011; Zhu et al., 2012). Each of these disorders include disturbances
related to the sense of self, reality testing, and autobiographical memory. The vulnerability of these key
functions of self that span so many mental disorders reflects both the importance of the DMN and its
sensitivity to disruption.

Schizophrenia
Given the disturbances in the coherence of self in schizophrenia, it would be surprising if the DMN were not
significantly impaired in its victims (Bluhm et al., 2007; Garrity 2007; Jang et al., 2011b). Failure to activate
and deactivate the DMN was found in schizophrenic subjects during memory tasks (Pomarol-Clotet et al.,
2008). This could reflect their difficulties with differentiating of concious awareness with unconscious
processes and the contamination of conscious awareness with dreamlike states. An interesting hypothesis is
that the frontal/dopaminergic deficits in schizophrenia impair its anticorrelational relationship with the
DMN, leading to the symptoms of hallucinations, thought withdrawal, and thought insertion (Manoliu et al.,
2014).

Depression
Functional abnormalities of the DMN have been found to be positively correlated with symptoms of
hopelessness, negative self-emotions, sleep disturbances, and negative self-focus (Grimm et al., 2008, 2009,
2011; Gujar et al., 2010; Guo et al., 2014). Some have hypothesized that a failure of DMN inhibition during
external tasks may result in depressed clients overpersonalizing their experiences (Li et al., 2014; Sheline et al.,
2009), which may explain the guilt and poor self-esteem associated with depression. Some go as far as to
postulate that DMN dysfunction may be a biomarker for both depression and dysthymia. This theory is
supported by the fact that DMN functioning has been shown to normalize with successful treatment with
selective serotonin reuptake inhibitors (Posner et al., 2013; Silbersweig, 2013).

Anxiety Disorders and Posttraumatic Stress Disorder


Overall, the sustained hyperarousal and vigilance that is a core factor of anxiety disorders disrupts the ability to
activate the DMN, exiling victims from their internal worlds, a coherent sense of self, and the ability to
connect with others. In general, all anxiety disorders reflect deficits in fear inhibition and hypervigilance to
danger cues. This state of brain and mind make it difficult to impossible for the DMN to be free to function

92
(Gentili et al., 2009). For example, clients with obsessive-compulsive disorder demonstrate less DMN
inhibition during task focus, which may be a biological signature of their inability to inhibit the thoughts and
motor behavior triggered by their anxiety (Stern et al., 2012).
PTSD is characterized by a hyperactive amygdala, hypoactive medial prefrontal cortex, and enhanced
amygdala connectivity with the insula, all of which inhibit the DMN (Palaniyappan et al., 2012; Patel,
Spreng, Shin, & Girard, et al., 2012; Rabinak et al., 2011; Seeley et al., 2007; Sridharan et al., 2008). In
addition, resting-state connectivity between the DMN and the right amygdala correlates with symptoms of
PTSD and predicts the occurrence and severity of future symptoms (Lanius et al., 2009). PTSD also results in
deficits in self-referential processing, autobiographical memory, and theory of mind, all of which depend on
proper coherence within the DMN. This is in stark contrast to those suffering with depression, who find
themselves unable to inhibit the DMN in order to turn off attention to the self.
At rest, subjects with PTSD demonstrated reduced coupling within DMN and greater coherence between
the DMN and the salience network (Sripada et al., 2012). Those with severe and chronic PTSD showed
deficits similar to those of schizophrenic patients when attempting to switch between DMN (Daniels et al.,
2010). It appears very likely that early trauma interferes with the development, coherence, and functioning of
the DMN (Daniels et al., 2011). Enhanced metabolic activity in the anterior and medial cingulate cortices in
both patients and their nontraumatized twins also suggest a possible genetic vulnerability to PTSD (Shin et
al., 2009).
Now that well over 100 studies have explored the DMN, we can begin to use these data to develop a more
sophisticated model of executive functioning. It is important to keep in mind that the DMN remains
controversial. Some believe that what is attributed to the DMN is background activity that is unrelated to any
kind of thought (Qin & Northoff, 2011). Even if the ideas we now associate with the DMN are found to be
false, the functions attributed to the DMN will eventually find a home within the brain (Northoff et al.,
2006).

A Case of Posttraumatic Stress


What we think we become.
—The Buddha

This is the case of a young prince who was groomed by his father to someday become a powerful king. The
prince was provided with everything he needed but was never allowed to stray beyond the palace’s tranquil
garden walls. Whatever his heart desired was immediately brought to him, and bad news was never allowed to
reach his ears. One day while his father was away, he commanded a driver to take him on an excursion to view
his future kingdom. On three successive days, these trips exposed him to aging, illness, and death. When he
saw a corpse by the side of the road, he asked the driver what it was. “It’s a dead body,” said the driver. “Alas,
death comes to us all.”
So sheltered was the prince that this news struck him like a bolt of lightning. This sudden and unexpected
exposure to the transitory nature of life plunged him into crisis. All that he had assumed to be permanent was
fragile and fleeting. As with so many before and since, the prince’s trauma led him to question reality and the
meaning of life. He began to wonder what in this world has lasting relevance. You may have already
recognized this as the story of Siddhartha, the Buddha. In the face of his crisis of meaning, Siddhartha
abdicated his throne, left his family, and went in search of truth and enlightenment. After decades as an
ascetic, following many teachers and applying many techniques, Siddhartha came to a new understanding of
the human condition and a strategy for attaining liberation.
In the 2,500 years since his passing, Buddhists have been engaged in active exploration of the workings of
the human mind. Simultaneously philosophical and psychological, Buddhism addresses many of the same
issues confronted by therapists. Through self-reflection, introspection, and meditation, Buddhists have
developed detailed interpretations of the human experience and sophisticated strategies to relieve what they
call suffering.
Buddhism and psychotherapy have both emerged as a consequence of the expanding complexity of our
minds and the suffering it can cause us. Our brains have evolved in such a way that our attention can become
completely absorbed by negative events from the past and anxiety about the future. Functions that once
supported our survival now impair the quality of our lives. I have come to believe that integrating Buddhist
psychology into our work with some of our clients can provide an additional and supportive conceptual
framework for treatment.

93
The Irritation of Life
I have taught one thing and one thing only, dukkha and the cessation of dukkha.
—The Buddha

The Sanskrit word dukkha is most often interpreted as sorrow, misery, or suffering. Sukha, the opposite of
dukkha, means happiness, comfort, or ease. This ancient word comes from the nomadic Aryans who brought
Sanskrit to India before recorded history. Kha meant “sky, space, or the hole at the center of the hub of a
wheel” and dukkha was used to describe “an out-of-round or poorly formed axle hole” that led to an irritating
ride. It is this irritation with the state of one’s existence that may be the best translation of dukkha.
This irritation comes from a combination of the physical pain that exists in nature and the suffering
created by our minds. While the painful aspects of life that traumatized our young prince are unavoidable, we
can gain freedom from the suffering that our minds create. Buddhists divide suffering into three categories:

1. The pain of impermanence—anxiety about future loss and fear of change, or dissolution; “Whatever
arises will pass away.”
2. The pain of attachment to ego—guilt, shame, failure, vulnerability, losing to others, and disappointing
others.
3. Existential pain—emptiness, meaninglessness, angst, dissatisfaction, discontent, unsatisfactoriness,
demoralization, apathy, and despair.

Because our minds struggle to predict and control the world around us, we are in a constant state of
anxiety because the nature of the world is change. The voices in our heads constantly remind us of negative
things from the past and dangers that we may encounter in the future. A natural consequence of a cortex
shaped to predict and control is to worry about the future to the point where the present is full of misery.
Chronic PTSD is an extreme example of life’s irritations amplified into a potentially disabling mental illness.

Case Conceptualization and Treatment


Pain is inevitable, suffering is optional.
—The Buddha

The core of Buddhist psychology emphasizes the importance of developing insight into the nature of dukkha,
the conditions that cause it, and how to alleviate it. A Buddhist psychologist would begin conceptualizing a
clinical case by determining the nature of the client’s suffering. Treatment would begin by introducing the
Four Noble Truths: (1) worldly existence causes suffering because (2) we are attached to ideas, beliefs, and
feelings that are illusory. The treatment plan would be to (3) sever our attachments to the illusions (4) by
following the Eightfold Path.
The Eightfold Path is the specific intervention that includes a set of interdependent practices for the body,
mind, brain, and social relationships. The Eightfold Path focuses on our views, intentions, speech, actions,
livelihood, effort, mindfulness, and concentration. Each is directed toward the recognition and cessation of
dukkha. By behaving decently, cultivating discipline, and practicing mindfulness and meditation, one can
come to see reality more clearly, allowing one to be liberated from suffering. Buddhist psychology, as you may
have already noticed, is quite compatible with cognitive-behavioral therapy. The primary differences in my
mind are the depth of analysis and intensity of treatment. While Western therapists focus on symptoms,
Buddhist therapists focus on life as a whole.
Some have said that our minds are prisons and meditation is the tunnel out, a perspective similar to Plato’s
allegory of the cave. From a Buddhist perspective, our knowledge of ourselves and the world is grasped only
through meditation. The Buddha didn’t teach compassion, but discovered it to be a natural outgrowth of
insights revealed during meditation. As we come to comprehend more clearly the fact that everyone is
suffering and we are all in this together, competitiveness and fear are replaced with empathy and compassion.
When you feel the pain in others in the process of meditation, you naturally want to relieve it. When you
attain true enlightenment, suffering still exists, but its foreboding and shameful aspects are taken away.

94
Beware of the Illusions of Consciousness
We don’t see things as they are; we see things as we are.
—Anaïs Nin

While every psychiatric disorder results in exaggerated emotions and distorted thinking, you don’t need a
diagnosis to distort reality. In fact, all of our minds distort reality to enhance survival and decrease anxiety.
Much of what we in the West have learned about these distortions has come from social psychology, clinical
psychology, and neuroscience. The distortions of the psychodynamic unconscious—reflected in defense
mechanisms such as reaction formation, denial, humor, and intellectualization—are thought to keep thoughts
and feelings out of conscious awareness in order to help us regulate and avoid negative emotions. Defense
mechanisms enhance survival by reducing shame, minimizing anxiety, and decreasing awareness of depressing
and demoralizing realities. Some defenses also support social cooperation and lead us to either overlook or put
a positive spin on the bad behavior of family and friends.
Let’s take a look at some of the illusions of consciousness. The first is that our conscious awareness comes
together at some specific location within our heads and is presented to us on a screen. This Cartesian theater
—an homage to Descartes’s articulation of mind-body dualism—creates the subjective illusion of self as a
nonphysical spirit inhabiting the body as opposed to being one with it (Dennett, 1991).
A second illusion is that our experience occurs in the present moment and that conscious thought and
decision making precede feelings and actions. In fact, our brains react to internal and external stimuli in as
little as 50 milliseconds, yet it takes more than 500 milliseconds for conscious awareness to occur. During this
half-second, hidden layers of neural processing shape and organize these stimuli, trigger motor responses, and
select an appropriate presentation for conscious awareness (Gibson, 1966; Panksepp, 1998; Libet, 1983).
By definition, hidden layers of neural processing cannot be directly observed. Like black holes, we are
made aware of their existence by their effects upon the visible world. These unconscious processes will
highlight some aspects of experience while diminishing others, orient us to certain aspects of the environment,
and completely block awareness of others. Our hidden layers translate past experience into an anticipated
future, often converting past trauma into a self-fulfilling prophecy of future suffering (Brothers, 1997; Freyd,
1987; Ingvar, 1985). This carryover of past learning into the present where it may be irrelevant or destructive
is certainly one of the contemporary human brain’s major design flaws and the focus of countless therapy
sessions.
The projection onto the screen of our Cartesian theater is actually generated within the hidden layers of
our neural architecture prior to conscious awareness. This leads us to assume that the world of our experience
and the objective world are one and the same. We also tend to believe that we have all the necessary
information we need to make choices. In truth, we often have little or no access to the information or logic
upon which we base our decisions. In addition, we possess a powerful reflex to confabulate in the absence of
knowledge (Bechara, Damasio, Tranel, & Damasio, 1997; Lewicki, Hill, & Czyzewska, 1992).
A third illusion, which relies on the first two, is that our thoughts and behaviors are under our conscious
control (Bargh & Chartrand, 1999; Langer, 1978). This leads us to consistently overestimate our control over
an outcome, while underestimating the role of chance, unconscious influences, and outside forces (Taylor &
Brown, 1988). Although we may feel as if we are at the wheel of our lives, we are often in the passenger seat.
The illusions of the Cartesian theater, living in the present moment, and being in total control of our
actions have all been exposed through research, clinical experience, and common sense, but they have been
invisibly woven into the fabric of our perception, memory, and character (Levy, 1997; Reich, 1945). If the
Buddha were alive today, I suspect he would take great delight in the array of ways in which his insights have
been confirmed.
Social psychologists have also identified a number of consistent errors in human judgment that can be
especially damaging to relationships between individuals, groups, and nations. Our tendency to explain the
behavior of others based on aspects of their character, while explaining our own behaviors as a result of
external factors, is referred to as the fundamental attribution error (Heider, 1958). In other words, others flunk
tests because they are not smart enough or are too lazy to study; we fail because the test wasn’t fair or because
the professor wasn’t very good. An extension of this attributional bias leads to a phenomenon called blaming
the victim, where individuals victimized by crime or poverty are believed to have done something to create
their misfortune (Ryan, 1971).
While individual perspectives are limited and incomplete, this does not stop us from assuming that we
possess the true view of the world. This egocentric bias leads us to reflexively believe that anyone who sees the
world differently from ourselves is misguided or dull-witted. Another bias is called belief perseverance or

95
confirmation bias—the tendency to only attend to facts supportive of existing beliefs (Janoff-Bulman, 1992;
Lord, Ross, & Lepper, 1979). This tendency is likely driven by the tenacity of fear memories stored within the
amygdala and our desire to avoid the anxiety of the unknown. This may explain why prejudice continues to
persist in the face of evidence that it is wrong.
Another reason that self-deception may have been selected during evolution is because it aids in the
deception of others. The more we believe our own deceptions, the less likely we are to give away our real
thoughts and intentions via nonverbal signals. In fact, it requires considerably more brain power to lie than to
tell the truth, and even more to convince others that we are being honest with them when we are lying (Ganis
et al., 2003). Actions and beliefs that are the opposite of our true desires can be quite effective in deceiving
others. It has also been noted that “people are remarkably reluctant to consider impure motives in a loud
moralist” (Nesse & Lloyd, 1992, p. 611), despite the well-publicized downfall of one moral crusader after
another. In fact, the best con artists are often so convincing that their victims refuse to accept that they have
been cheated at all.
Cognitive-behavioral therapy attempts to shine the light of conscious awareness on belief perseverance and
attribution biases and to undermine the conservative nature of our unconscious processing. Psychodynamic
therapists engage in a deep exploration of unconscious defenses and primitive emotional states. By
encouraging clients to be open to new ideas and to take responsibility for positive change, we challenge them
to reorganize the neural networks of their hidden layers.

Body, Brain, and Experience


Men are disturbed not by things, but by the view which they take of them.
—Epictetus

Sandy came to therapy in her mid-40s with the usual concerns about relationships, family, and career.
Although her mood was generally upbeat and positive, she occasionally came to sessions feeling irritable,
deflated, and hopeless. When I mentioned her fluctuating moods, she was distressed to discover that her
emotional lability was noticeable to others. As Sandy focused on her moods, she said that they seemed to
come out of nowhere and disappear just as mysteriously. She realized that these mood states had been part of
her life as far back as elementary school. When she was down, she felt like a fraud, and planned to quit her job
and leave her husband. “When I feel this way,” she said, “I just lose the will to live.”
We monitored and discussed her experiences through a number of mood cycles and engaged in
considerable speculation about their origin. Her father was prone to moodiness, and she had a maternal aunt
who had a “nervous breakdown” decades earlier, possible signs of a genetic inheritance or modeling of
depression that she witnessed as a child. Sandy struggled to find thoughts, feelings, or events in her life that
would precipitate her mood swings and discovered that they did not coincide with anything related to her
work, family, menstrual cycle, exercise, or diet. All serious medical conditions were ruled out, her only
physical complaint being her allergies and frequent sinus infections. On the outside chance there was some
relationship between her use of antihistamines and her mood changes, we included all of her medication use
on her mood chart.
Although we did not find any connection between mood and medication, it did turn out that she
consistently lost her will to live a day or two before suffering a sinus infection. Her mood would then improve
shortly after the onset of her respiratory symptoms and headaches. Once we made this connection, we waited
for the next dip in mood to see if it would again be followed by a sinus infection. Sure enough, another swing
came on schedule. Although we still did not know what affected her mood, the timing did suggest that it was
related to her cycle of allergies and sinus infections.
We decided to anticipate her next dip in mood with a new plan. We agreed that she would stop evaluating
her life on days when she lost her will to live. She was not allowed to think about leaving her husband or her
job or to assess her worth as a person. Instead, the mood dip would be a cue for her to go to the health food
store, buy vitamin C and zinc tablets, and rearrange her schedule to reduce stress. Sandy had to remain
mindful of the possibility that what she experienced as negative emotions was really a result of biological
changes related to a physical illness and not a collapse of character or impending catastrophe. We worked on
developing a safe internal place for her to retreat to at these times where she could soothe and comfort herself
and focus on healing.
Over time, the association between sinus infections and mood changes held up. For some unknown
reason, Sandy’s biochemistry reacted to a sinus infection with a sharp drop in mood that led her to reinterpret

96
the value of all aspects of her existence. By being mindful of this process and using her executive functioning
and self-awareness, she was able to engage in different behaviors and create a better outcome. We had
converted what usually led to an existential crisis into a trigger for enhanced self-awareness and self-care.
At the heart of psychotherapy are two interwoven processes. The first is the way in which our brains and
minds construct reality, while the second is our ability to modify these constructions to support mental health
and well-being. People come to therapy because one or more aspects of their lives are not how they would like
them to be. Most often our clients know what they should be doing differently but cannot bring themselves to
make changes. They come in with a feeling that something within them is holding them back. The answers to
their questions can usually be found in the architecture of the hidden layers of neural processing that construct
our reality, guide our experience, and shape our identity. The take-home message from psychoanalysis,
Buddhism, and neuroscience is to be a skeptical consumer of the offerings of your mind.

Summary

As meaning makers, we constantly seek answers to questions personal and universal, simple and complex.
Because uncertainty makes us anxious, we often accept incomplete and even bad answers rather than embrace
our ignorance. When we focus on a problem in greater depth, we begin to discover the unconscious and
invisible biases hidden within our commonsense observations. This is how the earth shifted from flat to round
and how the mind shifted from the heart to the brain. Later, and with further observation, we discovered the
earth is not quite round and the heart does influence our thoughts and emotions. So goes scientific progress.
The consistency of cognitive and emotional biases across individuals and cultures reflects our shared neural
evolution. Some of these biases are the result of natural limits to our perspective and judgment, while others
may have evolved to help us cope with living in an uncertain and dangerous world. Although many of our
perceptual biases appear to serve us, they can also lead to the kinds of problems that often become the focus of
psychotherapy.

97
Chapter 10
From Neural Networks to Narratives:
The Quest for Integration
There is no greater agony than bearing an untold story inside you.
—Maya Angelou

Biology expresses a fundamental strategy of the physical universe; it connects things. Neurons, neural
networks, individuals, tribes, and countries reflect nested organizations within more complex organizations.
As we zoom in to look at groups of neurons and zoom out to look at groups of people, the same principles of
connectivity and homeostatic balance hold true. As we learn about the necessary synergistic connectivity of
neural networks, we are also coming to understand the relationship between neural network imbalances and
mental distress. From extreme PTSD to everyday neurosis, we all exhibit a pattern of integration and
dissociation reflective of our adaptational history and the current functioning of our brains. At the level of the
experience of self, networks dedicated to sensation, perception, and emotion seamlessly integrate in the
emergence of conscious experience (Damasio, 2010; Pessoa, 2008; Fox et al., 2005). Let’s take a look at the
impact of a somewhat simple breakdown of neural network integration on the experience of self.
A few years ago, a young man in his late teens came in for a therapy session. The previous September,
Craig had left home to attend his first year of college, but by mid-December, something had gone haywire.
His parents were called by the dean and told that Craig had not been going to classes for weeks. They were
also informed by the resident advisor that five days earlier, Craig had locked himself in his room, thrown all of
his and his roommate’s possessions out of the window, and was listening to the same song 24 hours a day. His
parents raced to campus to find him in the middle of an acute psychotic episode.
Craig had been released from the hospital where I worked just a few weeks earlier, and it was good to see
him once again independent and active. As he walked across my office, I could see his movements were
slowed by the medications that were keeping his hallucinations at bay. I had seen Craig in individual and
group therapy for approximately one month. His symptoms had slowly cleared and he had been released to his
parents’ care a week earlier. This was his first session since being discharged. After he settled in, I asked him
how things had been going since he left the hospital. Slowly, and in a soft voice, he told me that life was
pretty good and that he enjoyed playing his guitar and working on some new songs. He wasn’t feeling
paranoid or hearing voices like he had been weeks ago; his sleep and appetite were okay, and he felt like he
was ready to return to school. “There’s only one problem, Doc. I don’t feel comfortable at home because my
parents and brother have been replaced by doubles.”
“Doubles?” I asked him. “What do you mean, doubles?”
Craig started by saying that he had gotten a strange feeling about his parents and brother when they came
to visit him in the hospital, but he figured he was feeling off because of the medication. Once he got home, he
discovered the reason for his strange feelings. “After a while I realized that they’ve been replaced by doubles!”
I gave him my best quizzical therapist expression and asked what made him think they were doubles. Craig
described how they were excellent copies and well prepared to trick him. He asked them scores of questions
he thought only his parents and brother could answer and, sure enough, they got them right. “Whoever is
doing this to me is good!” he said with nervous admiration. When I asked him again how he could be so sure
they were replacements, he replied with annoyance, “Don’t you think I would know my own parents?”
This syndrome of impostors, Capgras syndrome, can occur alone or with schizophrenia, temporal lobe
epilepsy, dementia, and head injury (Serieux & Capgras, 1909). Although the neurobiology of Capgras
syndrome is not certain, it is likely the result of a disconnection between networks of perception, emotion, and
conscious analysis (Alexander, Stuss, & Benson, 1979; Merrin & Silberfarb, 1979). An EEG study found
“abundant and severe EEG abnormalities” in the area of the temporal lobes, leading the authors to suggest
that the delusion of impostors is caused by a “dysrhythmia” of brainwaves in networks responsible for
matching faces with emotional familiarity (Christodoulou & Malliara-Loulakaki, 1981).

98
Capgras syndrome does not affect networks responsible for recognizing familiar faces. Craig could see that
his imposters were physically identical to his parents and brother. But Craig’s experience was that they no
longer felt like his parents—the emotional glow of recognition of loved and familiar people was missing
(Hirsten & Ramachandran, 1997). We can hypothesize that a disconnection or lack of coherence occurs
between the circuits of the temporal lobes responsible for face recognition and the ompfc-amygdala axis,
which would provide the emotional reaction of seeing a loved one. With this connection somehow disrupted,
Craig’s intact left hemisphere constructed the delusion of imposters, an explanation that is logical if you accept
the experiential premise. The people in Craig’s home looked and acted like his family, but without the usual
input from the emotional circuitry responsible for the feeling of familiarity, his left-hemisphere interpreter
concluded that they must be imposters.
Capgras syndrome may be the opposite of a déjà vu experience, where something new is paired with a
feeling of familiarity. Déjà vu is likely a random firing of familiarity circuits in an unfamiliar setting. The fact
that strong déjà vu experiences are often reported by patients with temporal lobe epilepsy suggests that their
out-of-control electrical firing is activating the amygdala, which is deep within the temporal lobes. The
frequency of déjà vu experiences has been found to correlate with decreased hippocampal gray matter (Brázdil
et al., 2012; Takeda et al., 2011). Craig was experiencing what amounts to the opposite of this experience. He
expected to have the feeling of recognition, but didn’t. This is probably what he was referring to when he said,
“Don’t you think I would know my own parents?”
The delusion of impostors generated by the left-hemisphere interpreter parallels attributions triggered by
déjà vu experiences, such as past lives, clairvoyance, and other paranormal processes. This very normal impulse
to make sense of nonsense is also seen in those with schizophrenia, who attempt to create logical explanations
for bizarre sensory experiences (Maher, 1974). When faced with the experience of thoughts being inserted in-
to their heads, patients ask themselves, “Who would have the technology to do such a thing?” When I worked
in Boston, patients pointed the finger at MIT, while the people I treated in Los Angeles suspected Cal Tech.
Delusional beliefs can become quite central to a client’s life and difficult to dislodge. For example, when three
patients, each of whom believed he was Christ, were housed together at a hospital in Ypsilanti, Michigan,
each maintained his own belief while coming to the conclusion that the other two were delusional (Rokeach,
1964).

Pathways of Integration
All organs of an animal form a single system . . . and no modification can appear in one part without bringing about corresponding modifications
in all the rest.
—George Cuvier

The long and circuitous path of brain evolution has not provided us with a brain that is simple in design. We
have already seen how the brain consists of different memory systems, two hemispheres with different
processing capabilities, and multiple executive systems controlling different skills and abilities. We have also
explored how psychological suffering arises when these systems get out of sync.
Although we are just beginning to understand the functions and the complexities of our neural pathways,
some consistent findings are beginning to emerge. As discussed in Chapter 2, two primary pathways of
integration to consider are top-down and left-right. It is also important to keep in mind that they are not
independent of one another. Left cortical areas (top and left) have developed certain special connections, as
have subcortical and right-hemisphere regions (right and down). Three more specific relationships within the
frontal lobes (the ompfc and the dlpfc), between the hippocampus and amygdala, and between the executive
systems of inner and outer experiences (DMN and frontal-parietal executive circuitry) all exist in a dynamic
relation to each other. These systems also have particular associations with both top-down and left-right
integration.
Let’s review the general map of the brain’s pathways of integration. In Table 10.1, notice the alignment of
these four pathways. Top-down, left hemisphere, dlpfc, the hippocampus, and the task-positive network are
aligned because they tend to be connected more heavily with one another than with those in the column on
the right. They also tend to be biased toward conscious, rational, and language-based functions. Bottom-up
processing, the right hemisphere, the ompfc, the amygdala, and the DMN appear to have denser connectivity
among themselves and are more likely to be involved with unconscious, somatic, and emotional functions. For
example, Capgras syndrome may reflect a disconnection of bottom-up emotional processing involving the
amygdala and cortical analysis involved in face recognition.

99
TABLE 10.1
Pathways of Integration
Top-Down (Cortical) Bottom-Up (Subcortical)
Left hemisphere Right hemisphere
dlpfc ompfc
Hippocampus Amygdala
Task-positive networks DMN

There is presently a great deal of research focused on understanding these networks and creating more
precise anatomical and functional distinctions between them. Separating similar and different roles of each
brain region in each hemisphere is also under exploration, as is the mapping of patterns of activation for
different symptoms and diagnostic groups (Dougherty et al., 2004). As with all of this research, we have to
keep in mind that age, gender, and life experiences all play a role in how these networks organize and function
in each individual. For our present purposes, I have chosen to focus on these general categories because of
their applicability to psychotherapy and mental health.

Top-Down–Bottom-Up
The complexity of the nervous system is so great, its various association systems and cell masses so numerous, complex and challenging, that
understanding will forever lie beyond our most committed efforts.
—Santiago Ramón y Cajal

Although there are many vertical circuits that cut across the horizontal strata of the brain, important top-
down networks for psychotherapists are those connecting the ompfc and amygdala. The ompfc and the
amygdala are connected by dense bidirectional networks that feed physiological and emotional information
upward to the cortex while allowing the ompfc to modulate the output of the amygdala to the autonomic
nervous system (Ghashghaei & Barbas, 2002; Ghashghaei et al., 2007; Hariri et al., 2000, 2003). Think of the
amygdala as a primitive structure designed to link immediate threat with a rapid survival response. Think of
the ompfc as having the ability to gather and update information and use it to predict potential outcomes and
shape behavior (Dolan, 2007; Rosenkranz, Moore, & Grace, 2003). Perhaps a good analogy is a squad of
soldiers trained to fight and survive (amygdala and autonomic nervous system) and a general who continues to
keep an eye on the entire battlefield, update his strategy, and adjust long-range goals (ompfc).
In the normally functioning brain, ompfc-amygdala activation reflects a dynamic moment-to-moment
balance of focused attention and emotional arousal (Simpson et al., 2001). When faced with psychosocial
stressors, we see elevated cortisol levels along with increased activation in the amygdala and lower levels of
activation in the ompfc (Kern et al., 2008). Higher levels of ompfc activity are believed to reflect an inhibition
of affective processes and an enhanced focus on the outside world, while a decrease suggests a shifting to
DMN activation and attention to internal processes. While negative affect decreases along with amygdala
activation, activation in the ompfc increases (Urry et al., 2006). It is now believed that each of us has a unique
homeostatic balance of this circuitry that shapes our emotional regulation and affective style (Davidson, 2002).
Let’s think about what happens in the human brain during public speaking. For most individuals, getting
up to speak results in increased cortical activation. This makes sense because we need our cortex to cope with
the cognitive demands of delivering a talk. But when individuals who have social phobia or a fear of public
speaking go up to the podium, they experience a decrease in cortical activity and an increase in amygdala firing
along with bodily symptoms of anxiety and panic (Tillfors et al., 2001). This may also account for stage fright,
where people either forget their lines or find it impossible to speak when faced with an audience. High levels
of cortisol, dopamine, and bottom-up cortical inhibition from the amygdala can take the prefrontal cortex
“off-line” (Arnsten & Goldman-Rakic, 1998; Bishop, Duncan, & Lawrence, 2004). This “amygdala hijack,”
as it is called in the self-help literature, is the takeover of cortical executive functioning by the amygdala and
other subcortical systems (Goleman, 2006).
The balance and integration of the ompfc and amygdala are influenced by attachment schema, past
trauma, current stress, and a variety of neurochemicals (Hariri, Drabant, & Weinberger, 2006; Heinz et al.,
2005). When people suffer from symptoms of depression or anxiety, there is a general decrease in cortical
activation and an increase in the salience system of the cingulate and insula corticies (Kennedy et al., 2007;
Mayberg et al., 1999). This balance reverses as mood lightens with or without treatment (Kennedy et al.,
2001). It has also been found that pretreatment metabolism in these and other regions predicts response to

100
antidepressant medication (Davidson et al., 2003; Pizzagalli et al., 2001; Saxena et al., 2003; Whalen et al.,
2008; Wu et al., 1999).
Within this broad top-down system, numerous subsystems are likely involved in emotional regulation.
Studies have demonstrated a variety of activation patterns in top-down networks during tasks of affect
regulation and the voluntary suppression of emotions (Anderson & Green, 2001; Beauregard, Lévesque, &
Bourgouin, 2001; Phan et al., 2005). For example, the coordination of activity between the amygdala and the
anterior cingulate has been shown to be correlated with trait anxiety and a susceptibility to depression
(Pezawas et al., 2005). The ability to suppress cigarette craving correlates with increased activation in the
cingulate cortex and an inhibition of sensory and motor regions as subjects respond to smoking-related
stimulus cues (Brody et al., 2007).
The anterior cingulate, amygdala, and insula are modulated by the processing of internal somatic
experience, which may be the same circuitry activated during therapy as we integrate conscious awareness with
somatic, emotional, and memory processing (Critchley et al., 2002). Simultaneous top-down and left-right
inhibition is likely responsible for what Freud called repression. As prefrontal and anterior cingulate regions
are inhibiting conscious recall of explicit memories, left frontal networks can be simultaneously inhibiting
negative somatic and emotional memories stored in right-biased systems (Anderson & Green, 2001). The
result would be a lack of conscious recall of a threatening experience and a dissociation of experience from
conscious awareness.

Left Hemisphere–Right Hemisphere


The interpretive mechanism of the left hemisphere is . . . constantly looking for order and reason, even when there is none—which leads it
continually to make mistakes.
—Michael Gazzaniga

As we saw in Chapter 3 left-right integration is required for proper language functioning, bodily awareness,
emotional regulation, and many other essential human processes. As we will soon discuss, storytelling and
narrative structure as universal aspects of human culture may have emerged, in part, to assist in the integration
and coordination of the two very different brains.
A greater left-hemisphere advantage in verbal processing has been shown to be a predictor of a more
favorable outcome in CBT (Bruder et al., 1997). This suggests that those individuals with more left-
lateralized language abilities may have stronger inhibitory capacities over emotional experience stored in the
right. It has been shown that good readers have less interhemispheric connectivity and are better at processing
rapidly changing sensory input (Dougherty et al., 2007). For some tasks, less integration and cooperation are
an advantage, especially when speed or focus of attention are factors. However, having the input of both
hemispheres may be quite adaptive when we are solving complex social and emotional problems (Cozolino,
2008).
A form of treatment used to readjust right-left balance is transcranial magnetic stimulation (TMS). TMS
is a noninvasive, painless technique for the stimulation and inhibition of neural firing. A coil of wires is placed
on the scalp that generates a magnetic field strong enough to penetrate the skull. This magnetic field is
transformed into current flow in the brain that temporarily excites or inhibits select areas, and it can be
applied either as a single pulse or repetitively (rTMS). Depending on its frequency, the pulse either increases
or decreases cortical excitability—fast rTMS increases activation while slow rTMS decreases it (Daskalakis,
Christensen, Fitzgerald, & Chen, 2002).
Left-biased prefrontal activation downregulates negative affect in nondepressed individuals while
depressed individuals show bilateral frontal activation (Johnstone et al., 2007). In several studies, patients with
treatment-resistant depression experienced symptomatic improvement after a series of fast rTMS treatments
applied to the left prefrontal cortex (Pascual-Leone et al., 1996; George et al., 1997; Figiel et al., 1998;
Teneback et al., 1999; Triggs et al., 1999). These repeated magnetic pulses to the left hemisphere may have
increased activity and shifted the balance of mood in a more positive direction. Slow rTMS applied to the
right prefrontal cortex resulted in similar improvements in depressive symptoms (Klein et al., 1999; Menkes et
al., 1999). Slower-frequency rTMS to the right prefrontal cortex was thought to inhibit right frontal
functioning and have less adverse side effects (Schutter, 2009). However, conflicting results have also been
found, so our understanding of these processes is still developing (Holthoff et al., 2004).
Studies of rTMS and depression lead us to the conclusion that the technique’s ability to both stimulate the
left hemisphere and inhibit the right hemisphere may prove equally useful in depressed patients. Current

101
views take the position that restoring the balance between left and right prefrontal cortex activity is more
important in treating depression than establishing clear increases in left-sided activity. If rTMS can have a
positive effect on depressive symptoms, might it work in the reverse manner for mania? Studies in this area are
less extensive, but findings do suggest some effectiveness of rTMS in the treatment of mania when it is
applied at high frequency to the right prefrontal cortex (Belmaker & Grisaru, 1999; Grisaru et al., 1998;
Michael & Erfurth, 2002; Saba et al., 2004).

Dorsal Lateral pfc–Orbital Medial pfc


Modern Psychology takes completely for granted that behavior and neural function are perfectly correlated, that one is completely caused by the
other. . . . It is quite conceivable that some day this assumption will have to be rejected.
—Donald Hebb

As a whole, the prefrontal cortex sculpts experience and behavior through a complex array of inhibitory and
excitatory activities (Knight, Staines, Swick, & Chao, 1999). You will recall that the prefrontal cortex is
divided into four regions and that the dorsal and lateral regions tend to engage in coordinated activity, as do
the orbital and medial areas. Because of these connections, they are often referred to as the dlpfc and ompfc.
The location of prefrontal activation varies depending on the emotional salience of the task: the more
emotional the task, the more ompfc activation—the more cognitively demanding a task, the more the dlpfc
takes center stage (Goel & Dolan, 2003; Northoff et al., 2004; Schaefer et al., 2002). As the cognitive
demands of a task increase, there is a decrease in activation not only in the ompfc, but also in the amygdala
and anterior cingulate (Pochon et al., 2002; Rushworth & Behrens, 2008). This is likely the reason why
engaging in cognitive tasks, like word or math problems, often reduces anxiety.
The dlpfc exerts control over neural processing based on higher-order rules (environmental context,
prediction, etc.), while the ompfc does the same from the perspective of lower-order rules (impulse, drives,
emotions, etc.). From this we get the sense that top-down and bottom-up processing is interwoven with the
balance of activation between the dlpfc and ompfc, respectively. Interestingly, when people make decisions
congruent with implicit racial and gender biases, the ompfc and amygdala become more active, while the dlpfc
shows more activity when we express beliefs that are incongruent with prejudice (Knutson, Mah, Manly, &
Grafman, 2007). This reflects what we already know—more primitive impulses drive prejudice while more
information and expanded perspectives allow us to go beyond our reflexive limitations.
Experiencing the world from a first-person perspective and tasks of self-regulation activate ompfc regions
while situation-focused regulation activates dlpfc systems (Ochsner et al., 2004). The ompfc becomes involved
in diverse tasks that require differing kinds and degrees of self-referential knowledge (Ochsner et al., 2005).
Within the ompfc, the decoding of the mental states of others based on observable cues such as facial
expressions may rely on the right ompfc while reasoning about their mental states may be lateralized to the left
ompfc (Sabbagh, 2004).
When we consider the types of issues brought into psychotherapy, it is likely that we are working to build,
integrate, and balance the ompfc and dlpfc. Consider what we do when we assist clients in shifting from their
own perspective to looking at a situation from another point of view or from a more objective perspective. We
are calling upon the ompfc and dlpfc in different ways as we attempt to guide them to a more holistic
perspective of any particular situation. This process most likely enhances the growth of ompfc and dlpfc
systems, while building new brain networks to bridge and integrate the two. Optimal functioning necessitates
coordination, flexibility, and complementarity between these modes of functioning.
In situations of stress and trauma, the ompfc and dlpfc are capable of mutual inhibition or dissociation
(Roberts & Wallis, 2000). An inability of the ompfc to modulate stress will result in a decrease of activation in
the dlpfc during a cognitive memory task, causing a performance deficit (Dolcos & McCarthy, 2006; Drevets
& Raichle, 1998). When the ompfc and dlpfc are in proper balance, they create the possibility of true
cognitive-emotional integration (Gray et al., 2002). Building strong connections between ompfc-dlpfc circuits
creates resilience to stress, reduces the need to resort to dissociation, and results in greater affect tolerance and
ego strength.

Hippocampus-Amygdala
Emotions have taught mankind to reason.

102
—Marquis De Vauvenargues

The hippocampus and amygdala play a central role in learning and memory. The amygdala (in connection
with the ompfc) organizes emotional experience and (in moderate states of arousal) signals the hippocampus
about what is important to learn. On the other hand, the hippocampus (along with the dlpfc) participates in
the cognitive evaluation of situations that will inform the amygdala when to ramp up or back down on its
emotional reaction. In other words, if I can see that the dog is wagging his tail, perhaps I don’t need to be as
afraid of being bitten. Since the activation of emotion and the cognitive analysis of experience are both
necessary for normal functioning, the proper regulatory balance of the hippocampus and amygdala is vital.
The hippocampus is necessary for forming new explicit memories while the amygdala organizes highly
stressful and traumatic learning. At low levels of arousal, amygdala activation supports hippocampal learning
by boosting the biochemical aspects of neural plasticity. At higher levels of arousal, the amygdala stimulates
hypothalamic-pituitary-adrenal (HPA) activation, which interrupts hippocampal learning while supporting
fear-based amygdala learning (Kim, Koo, Lee, & Han, 2005; Kim, Lee, Han, & Packard, 2001). In essence,
during states of high arousal, hippocampal and amygdala networks can become dissociated, resulting in a
disconnection between visceral-emotional (amygdala) and declarative-conscious (hippocampal) processing
(Williams et al., 2001). Thus, optimal learning requires a low to moderate level of arousal, allowing for a
balance of amygdala and hippocampal participation.
Many clients, perhaps even the majority, do not come for treatment of a major psychiatric illness. Most
clients who are somewhat less ill have so far not been included in extensive (and expensive) outcome research
that includes brain imaging studies. Many people seek psychotherapy simply because, as they often say
themselves, life has somehow gotten out of balance. This may mean that their fears and worries have taken
control of their lives and limited their ability to function or find happiness and satisfaction. Others find
themselves devoid of emotion and without empathy for others, leading them to seek therapy to save their
marriages or their relationships with their children. Many have the sense that they are not living up to their
potential, or they get in their own way when it comes to worldly success and emotional satisfaction.
These clients are often referred to as the “worried well,” implying that they should somehow get over
themselves and get on with life. My sense is that this group of patients, in which I would include myself, also
suffer various versions of a homeostatic imbalance. They demonstrate an exaggerated reliance on intellectual
defenses, a tendency toward overemotionality, or negative attachment experience, which can become self-
perpetuating patterns leading to social isolation and living below their potential. All of these suboptimal
lifestyles are likely reflected in biased patterns of neural activation. While psychotherapy is a relatively recent
and culture-specific development in human history, talking to one another, seeking out advice, and
exchanging stories is central to human experience. I suggest that the evolution of the brain and the
development of narratives have gone hand in hand.

Task-Positive Networks: Default Mode Network


Happiness is not a matter of intensity, but of balance, order, rhythm and harmony.
—Thomas Merton

In Chapter 9 we learned that brain activation toggles back and forth across broad neural networks as it shifts
between external and internal processes (Jack et al., 2013). The task-positive networks (TPNs), which include
the dlpfc, parietal structures, and sensory and motor networks, organize our navigation of and interactions
with the external world. The default-mode network (DMN), which includes the ompfc, medial temporal
lobes, and the precuneus region of the parietal lobes, takes over during tasks that involve self (e.g.,
autobiographical memory, self-reflection, and a sense of self) and others (e.g., face recognition, mind reading,
moral reasoning). These networks are generally anticorrelational, meaning that when one becomes active the
other is inhibited. However, research has also shown that some tasks, which are less taxing or require an alloy
of external and internal focus, may require cooperation between the two.
The balance, integration, and cooperation of the TPNs and the DMN appear to be disrupted in a variety
of psychiatric diagnoses. These disruptions are likely involved in the disturbances of self seen in schizophrenia
that involve disorganized identities and disturbances of thinking and a breakdown of boundaries between self
and other. Symptoms of depression appear to correlate with an overemphasis on the self and an
overpersonalization of the meaning of events. Most significantly, anxiety disorders result in symptoms of
hyperarousal, hypervigilance, and exaggerated startle triggered by a hyperactive amygdala. This naturally leads

103
to turning on and keeping on TPNs, keeping the DMN inhibited. Chronic anxiety disorders will therefore
keep victims in exile from their inner worlds and make interpersonal connection and empathic processes a
challenge.
In thinking about the importance of the balance among the TPNs and the DMN, we have to keep in
mind that adequate integration and balance between these networks is not just an issue for those of us with
full-blown psychiatric disorders. Most everyone struggles with the stressors, anxieties, and fears of everyday
life that keep us from staying centered and remaining mindful. Our culture is much more focused on us as
human doings instead of human beings and there are no obvious external rewards for being of sound mind.
Because the DMN is an experience-dependent structure, we need to invest in building and strengthening our
internal worlds as a place of refuge, rejuvenation, problem solving, and strength. Children need to learn these
lessons as early as possible and witness parents taking their own internal worlds seriously.

From Neural Networks to Narratives


[Words] leave finger marks behind on the brain, which in the twinkling of an eye become the footprints of history.
—Franz Kafka

The evolution of the human brain is inextricably interwoven with the expansion of culture, the emergence of
language, and construction of narratives (Herrmann et al., 2007). Thus, it is no coincidence that human
beings possess limitless episodic memory and have the impulse to tell stories. Through countless generations,
humans have gathered to listen to stories of the hunt, the exploits of their ancestors, and morality tales of
good and evil. Stories connect us to others, prop up our often fragile identities, and keep our brains regulated.
It has long been known that these stories support the transmission of culture in addition to promoting
psychological and emotional stability. Therefore, I believe that both the urge to tell a tale and our vulnerability
to being captivated by stories are deeply woven into the structures of our brains.
Narratives perform an array of important functions including:

• Grounding our experience in a linear sequential framework


• Remembering sequences of events and steps in problem solving
• Serving as blueprints for emotion, behavior, and identity
• Keeping goals in mind and establishing sequences of goal attainment
• Providing for affect regulation when under stress
• Allowing a context for movement to self-definition

For most of human history, oral communication and verbal memory were the medium and repository of
our accumulated knowledge. The ongoing value of stories to each of us is highlighted in today’s world by the
energy we invest in television, movies, magazines, and everyday gossip and the drive of older generations to
repeatedly tell the same stories is matched by the desire of young children to hear them again and again. This
interlocking conduit of culture across generations carries memories, ideas, and ideals through time. The
importance of narratives in human evolution is further underscored by the fact that our ability to remember
and recall stories is essentially limitless. In fact, the astonishing abilities of memory experts rely on placing
discrete pieces of information into narratives that expand the capacity of working memory to the limits of
their imagination.
Although stories may appear imprecise and unscientific (Oatley, 1992), they serve as powerful tools for
high-level neural network integration (Rossi, 1993). The combination of a linear story line and visual imagery
woven together with verbal and nonverbal expressions of emotion activates and utilizes dedicated circuitry of
both left and right hemispheres, cortical and subcortical networks, the various regions of the frontal lobes, and
the hippocampus and the amygdala. The cooperative and interactive activation involved in stories may be
precisely what is required for sculpting and maintaining neural network integration while allowing us to
combine our sensations, feelings, and behaviors with conscious awareness. Further, stories connect individuals
with families, tribes, and nations and into a group mind that links each individual brain. It is likely that our
brains have been able to become as complex as they are precisely because of the power of narratives and the
group to support neural integration.
Much of neural integration takes place in the association areas of the frontal, temporal, and parietal lobes,
which serve to coordinate, regulate, and direct multiple neural circuits. They are our conscious switchboard
operators, able to use language and stories to link the functioning of systems throughout the brain and body.

104
An inclusive narrative structure provides the executive brain with the best template and strategy for the
oversight and coordination of the functions of mind. A story well told, containing conflicts and resolutions,
gestures and expressions, and thoughts flavored with emotion, connects people and integrates neural
networks.

A Story Well Told


Man’s mind, once stretched by a new idea, never regains its original dimensions.
—Oliver Wendell Holmes

Have you have ever watched the faces of small children as they listen to a gifted storyteller? You can see the
unfolding drama reflected in their eyes, on their faces, and throughout their bodies. Listeners will experience a
range of drastically shifting emotions, be absorbed in every detail, and even shout out warnings to characters
in danger. Narratives allow us to live vicariously, to take alternate points of view, and to learn about living. We
can escape our bodies in imagination to other possible selves and worlds yet to be created. Through stories we
have the opportunity to ponder ourselves in an objective way across an infinite number of contexts. In life and
in therapy, we can use stories to imagine our problems happening to someone else or view ourselves at a
distance. We can experiment with new emotions, actions, and language to edit the scripts of our lives
(Etchison & Kleist, 2000). Our ability to edit narratives summons us to try on new ways of being. (Recall the
case of Sheldon and his magic tricycle.)
Every screenwriting student learns the formula for constructing a successful narrative structure. Every story
needs a hero, a protagonist with whom we can identify. The protagonist is facing an external challenge and
possesses an inner wound that causes him persistent pain. At first the hero either avoids the challenge or fails,
leading him to question his ability to succeed or even his desire to change. The challenge confronting the hero
is at first resisted, then rejected, and eventually accepted. During the journey, the hero leaves behind old
definitions of self and travels into uncharted territory. Some inner transformation takes place that allows the
hero to face his demons, succeed in his worldly challenge, and solidify his identity.
This is essentially the universal myth of the hero, describing the transition from adolescence to adulthood
(Campbell, 1949). Redemption—a word commonly used for this transition—can happen at any age. The
adolescent struggling to attain adult status, Dickens’ emotionally shut-down Scrooge encased in the pain of
past trauma, or a client trying to make sense of early deprivation all have a wound that needs healing. My
explanation would be that what we share in common—brain, culture, language, and the fight for growth and
survival—are the underlying motives of the heroic narrative. Our shared struggle for survival and meaning is
far deeper and more powerful than our superficial differences.

Narratives and Emotional Regulation


Good psychiatry is a blend of science and story.
—Jeremy Holmes

As the language areas of the left hemisphere enter their sensitive period during the middle of the second year
of life, grammatical language in the left integrates with the interpersonal and prosodic elements of
communication already developed in the right. As the cortical language centers mature, words are joined
together to make sentences and can be used to express increasingly complex ideas flavored with emotion. As
the executive networks continue to expand and connect with more neural networks, memory improves, a sense
of time slowly emerges, and autobiographical memory begins to connect the self with places and events,
within and across time. The emerging narratives begin to organize the nascent sense of self and become the
bedrock of our sense of self in interpersonal and physical space.
As our experience of self and the stories we tell about ourselves become interwoven, they become the
center of gravity of our identities (Dennett, 1991). As children we are told by others, and gradually begin to
tell others, who we are, what is important to us, and what we are capable of. These self-stories are shaped by
culture and coconstructed with parents and peers. And although it does sometimes seem that children are
little scientists discovering the world, what we often miss is that they are primarily engaged in discovering
what the rest of us already know, especially about who they are (Newman, 1982). This serves the continuity of
culture from one generation to the next as we reflexively strive to re-create ourselves.

105
The role of language and narratives in neural integration, memory formation, and self-identity makes
stories a powerful tool in the creation and maintenance of the self (Bruner, 1990). Stories are powerful
organizing forces that serve to perpetuate both healthy and unhealthy forms of self-identity. There is evidence
that positive self-narratives aid in emotional security while minimizing the need for elaborate psychological
defenses (Fonagy, Steele, Steele, Moran, & Higgitt, 1991). In the same way, anxious and traumatized parents
pass along their negative experiences in the stories they tell. The recognition of the negative power of personal
narrations containing negative self-statements stimulated the development of rational and cognitive-based
therapies (Ellis, 1962). Having a positive self-narrative can make an immense difference in our ability to cope
with life’s challenges.

Trevor
Being deeply loved by someone gives you strength, while loving someone deeply gives you courage.
—Lao Tzu

Seven-year-old Trevor was brought to see me because his parents were concerned that he might have
“something troubling him.” He was very close to his grandfather, who had passed away six months earlier, but
he didn’t seem to have a reaction to this loss. While his parents felt they had done everything they could to
encourage him to talk about his feelings, he didn’t have that much to say. Trevor seemed to be a normal kid
with interests in science, video games, and computers. As he became comfortable, we hung out, played, and
talked about all kinds of things. During our second session, he mentioned that he liked doing puzzles, so I
purchased a few and brought them to the office.
Before our session, I spread one of the puzzles out on my desk. I put together a few pieces to give him a
jump start and quelled my own compulsive impulse to keep going. He was excited when he noticed it and
asked if he could help me work on it. “Certainly,” I said, and we sat down to a session of puzzling. It didn’t
take long for me to realize that he was having difficulty, and I wondered if I had chosen a puzzle that was too
complex for him. The last thing I wanted to do was to give him a failure experience.
I offhandedly suggested that we didn’t have to work on the puzzle if he would prefer to do something else.
“Maybe this one is too hard for us,” I said. “No,” he replied, “don’t give up. We’ll get it.” Impressed by his
determination, we continued to move pieces around in search of colors and patterns. Every once in a while, I
would leave a piece in front of him that I knew would fit with something he was holding. I became more and
more amazed at his patience and dedication. Many boys his age would have moved on to something else or
just cleared the table with the swipe of an arm.
After a while, I heard Trevor mumbling under his breath. He was repeating something over and over like a
song or mantra. I leaned over, slowly putting my ear closer and closer to him so that I could make out his
words. Finally, I could hear, “I think I can, I think I can.” He was chanting the theme of The Little Engine
That Could. He was the little train that kept on keeping on. I immediately felt my eyes well up and had to
resist the urge to hug him. Sure enough, he slowly got the hang of it and made lots of progress.
I later found out from his parents that Little Engine was his favorite story and one his grandfather loved to
tell him. They told me he wanted to hear it exactly the same way each time and if they made a mistake on any
word, he would stop and correct them. It was clear that this Little Engine was a kind of hero to him, and he
used the story when he was stressed by a challenging situation to regulate his anxiety and keep himself moving
ahead. Part of this heroic story was likely the memory of a loving grandfather whom he carried inside of him.
The Little Engine became a way for us to share his grandfather and the power of a story to soothe and inspire.
I came to realize that his grandfather had done a wonderful job of becoming part of Trevor’s experience of
himself and preparing him for his death. I learned that Trevor’s loss was complicated because, in many ways,
he still had his grandfather with him. I believe that Trevor’s ability to use narrative in this way and his
internalization of his grandfather’s love bode well for his healing.

Integrating Stories and the Self


Everyone is necessarily the hero of his own life story.
—John Barth

To serve their important role in emotional regulation, narratives need to have a brief summary or hook that

106
can be held in mind in the present moment. This summary, which can be a word, a phrase, a visual image, or
even a gesture, can instantaneously evoke the beginning, middle, and end of the narrative, and especially its
message. In Trevor’s case, it was the phrase “I think I can.” This decreased his anxiety, enhanced his problem
solving, and allowed him to discover his true competence.
Putting feelings into words (affect labeling) has long served a positive function for many individuals
suffering from stress or trauma. Labeling emotions correlates with decreased amygdala response and an
increase in right prefrontal activation (Hariri et al., 2000). It has also been found that activation in the
amygdala and the right frontal cortex are inversely correlated and that this homeostatic balance is mediated by
the ompfc (Lieberman et al., 2007). This suggests that the labeling process may require both the lateral and
medial prefrontal regions in order for cognitive processes to have a modulatory impact on our emotional
activation (Johnstone et al., 2007). The narrative, which simultaneously activates an array of networks,
enhances neural balance.
The perception of control has been shown to reduce emotional arousal and stress. It is likely that cognitive
processes involved in prediction and control activate frontal functioning and downregulate amygdala
activation. In other words, thinking we have some control puts us in a state of mind that prepares us to think
and activates prefrontal functioning, which reduces our emotionality. As a self-fulfilling prophecy, believing
you are an efficacious person stimulates frontal activation, making you a more efficacious person (Maier et al.,
2006).
Even writing about your experiences supports top-down modulation of emotion and bodily responses. In a
large series of studies, James Pennebaker (1997) and others have instructed subjects to keep a journal about
emotional issues of personal importance, especially experiences related to close personal relationships. These
studies have revealed increased well-being including a reduction in physical symptoms, physician visits, and
work absenteeism (Pennebaker & Beall, 1986; Pennebaker, Kiecolt-Glaser, & Glaser, 1988). This sort of
writing has also been found to correlate with greater T-helper response, natural killer cell activity, and
hepatitis B antibody levels as well as lower heart rate and skin conductance levels (Christensen et al., 1996;
Petrie et al., 1995; Petrie, Booth, & Pennebaker, 1998). Journaling about emotional issues likely increases
prefrontal activation, downregulating the negative emotional activation of the amygdala (Dolcos &
McCarthy, 2006). Our ability to tame the amygdala (and the HPA axis) using language and narratives results
in a cascade of positive physiological, behavioral, and emotional effects.

Levels of Language and Self-Awareness


The less men think, the more they talk.
—Baron de Montesquieu

Language is not one entity used for a single purpose. During the coevolution of brain and culture, different
languages emerged and expanded along with the sophistication of the brain. Introspection provides us with a
window to shifts in states of mind that reflect the activation and integration of different neural networks.
Through self-reflection, most of us become aware that we seem to shift back and forth among different
perspectives, emotional states, and ways of using language. I am aware of at least four levels of language
processing that take place within my clients and myself during these shifting states of mind; a reflexive social
language, a reflexive internal narrator, a reflective internal language, and a language of self-reflection.
Reflexive social language is a stream of words that maintains ongoing social relatedness and communication.
Primarily a function of left-hemisphere processing, this language mirrors activity within the interpersonal
world and is designed to grease the social wheels. Verbal reflexes, clichés, and overlearned reactions in social
situations spin a web of connections. Most of us experience this whenever we automatically say something
positive to avoid conflict, or tell people we are fine regardless of what’s troubling us. The natural clichés of
reflexive social langauge are as automatic to us as walking and breathing. This level of language serves the
same purpose as grooming in most primates.
In addition to reflexive social language, we are also aware of voices we hear inside our heads. This reflexive
internal narrator often departs in content and tone from what we express to others. While reflexive social
language is driven by current social cooperation, the reflexive internal narrator is shaped by emotions and
inner personal experience based on past learning. The reflexive internal narrator most likely first evolved for
the internalization of tribal mandates—the direction of the leaders—and was conserved by evolution into
what we now call the superego, the voices of our parents inside our heads, or the voices of self-doubt that
make us defer to others and take a lower role in the social hierarchy.

107
Our reflexive inner languages may be primary ways in which right-hemisphere processing and early
implicit memory influences moment-to-moment conscious awareness. Reflexive social language and the
internal narrator are akin to overlearned motor skills, semantic routines, and habits formed during our
learning history that serve to maintain preexisting attitudes, behaviors, and feelings. Together, the two
languages keep us in line with the group and in line with early programming. The conflict between these two
inner languages becomes particularly acute during adolescence if our peer group is discordant with our early
childhood training.
When we find ourselves observing the output of these reflexive forms of language, we move into two
reflective forms of language. The first, reflective internal language, is directed toward our actions in the outer
world and allows us to have private thoughts, to plan and guide future behavior, and to deceive others when
necessary. This is the kind of internal language that we encourage in the young and impulsive when we say,
“think before you act” or “look before you leap.”
Finally, a language of self-reflection emerges in states of openness, low defensiveness, and safety, likely
dependent upon the activation of the DMN. This level of language is not a mechanism of behavioral or social
organization but a vehicle of reflection, self-evaluation, thoughtful consideration, and potential change.
Achieving this state is a target of both talk therapy and meditative practices that attempt to use the mind to
change the brain.
The use of reflective versus reflective forms of language is an important aspect of mature thinking,
empathy, and executive functioning. Much of the early portion of therapy consists of exploring reflexive forms
of language to clarify a client’s conscious self-narrative and uncover early programming. In this process, clients
learn how their past and present behaviors may support and even create the problems for which they are
seeing help. Clients can learn that they can attend to what they say and do, reflect on what is reflexive, and
create experiments and challenges in the service of new ways of being.
As the language of self-awareness is expanded and reinforced, we learn that we are capable of evaluating
and choosing whether to follow the expectations of others and the mandates of our childhoods. The language
of self-reflection most likely requires a higher level of both neural and personality integration. In this
language, cognition is blended with affect so that there can be feelings about thoughts and thoughts about
feelings. At a very deep level, the language of self-reflection provides us with a tool to learn to quiet our
thoughts and move beyond words.
Therapy attempts to use self-reflection to create a metacognitive vantage point from which the shifting
states of mind that emerge during day-to-day life, and the reflexive language they trigger, can be reflected on
and eventually modified. This is accomplished by interweaving the narratives of client and therapist, hopefully
leading clients in a more healthful direction. Clients become aware of one or more of the narrative arcs of
their life story and then understand that change is possible by creating alternative story lines. As the editing
process proceeds, new narrative arcs emerge, as do possibilities to experiment with new ways of thinking,
feeling, and acting.
The importance of the unconscious processes of both parents and therapists is highlighted by their active
participation in the coconstruction of the narratives of their children and patients. This underscores the
importance of proper training and adequate personal therapy for therapists who will be putting their imprint
on the hearts, minds, and brains of their clients. And it certainly wouldn’t hurt for parents to have a little more
self-awareness in their day-to-day interactions with their children.
In essence, therapists hope to teach their clients that they are more than their present story. They can also
be the editors and authors of new stories. When we evolved the capacity to examine our narratives
(metacognition) and to see these stories as one option among many, we also gained the ability to edit and
modify our lives (White, 2001). The narrative process allows us to separate story from self. It is like taking off
your shirt to patch a tear and then putting it back on. This allows us to have the experience of a self that is
separate from our behaviors, feelings, actions, and problems. The fact that someone can say, “I’m not myself
today” implies the capacity for self-reflection and comparison between a current state of mind and our
everyday self-narrative. The ability to take other perspectives also enhances our empathy for others.

Abbey
Do not dwell in the past, do not dream of the future, concentrate the mind on the present moment.
—The Buddha

Abbey, an extremely bright and charismatic woman, came to my office with tears in her eyes and a smile on

108
her face. Even before she sat down, Abbey launched into a description of all the positive events that had
happened to her and her family over the last week. Seeing the pain in her eyes and the rigidity of her body, my
face must have reflected the sadness I was feeling. My expression seemed to make Abbey avoid my eyes and
speak even faster. From time to time, I would attempt to break in and ask her what she was feeling.
Abbey ignored my questions, talking at an ever faster pace. She reminded me of how as a child I would
cover my ears and hum when my mother was about to say, “Bedtime!” I soon realized that all I could do was
sit, listen, and wait. I sat across from her and tried to remain true to my feelings, allowing them to show in my
eyes and facial expressions. Eventually, her speech slowed; she became quiet and hung her head. Her feelings
seemed to have finally caught up with her, the impulsive stream of reflexive social language finally coming to a
halt.
I was considering what to say when she spoke: “I caught myself blabbing on.” It was good to see that
Abbey could employ her language of self-reflection and share her observations with me. I asked her what she
had been thinking about while sitting in silence. Abbey replied, “I was thinking of what an idiot I am and how
I must bore you with endless prattle about my stupid life.” Now she was sharing the content of her internal
dialogue, likely programmed early in life. She seemed deflated, depressed, and ashamed of herself. As a
reaction against her own shame, she attacked. “What a stupid job you have, sitting in this office every day,
listening to people’s problems. Why don’t you get out of here and get a life?” Abbey soon lowered her face
into her hands and began to sob. I could see that she was sharing with me the voices in her head and her fears
and doubts, but she was also projecting onto me her anger, confusion, and frustration. Her internal dialogue
was hurting her, and she wanted me to know how she felt. I said, “Being criticized can be really painful.” She
instantly knew that I was talking about the pain being inflicted by her inner voices and the pain she felt from
her parents as a child.
When she spoke again, she told me of the emptiness she felt from the loss of her husband a few months
earlier (until this point she had denied that it had much impact on her). It had become clear to her over the
last few minutes that she had been coping with her sadness by burying herself in a flurry of words, social
activities, and taking care of others. After a few minutes of silence and deep sighs, Abbey began to talk about
how much she missed his hugs, good advice, and the feeling of safety of having him around. Abbey was now
speaking in the language of self-reflection. In this state of mind, she was able to mourn the death of her
husband.
When clients shift to the language of self-reflection, the changes in their tone, manner, and mood are
palpable. I imagine at this moment that clients have the clearest perspective on their thoughts, behaviors, and
feelings. They speak more slowly because the organization of sentences takes time when they no longer rely on
clichés and semantic habits. Emotions bubble up and clients feel safe enough to express them in a process that
enhances affect regulation. This is when I feel most confident about a client’s ability to join me as a
collaborator in the therapeutic process. These states are usually fleeting and are often not supported by family,
friends, or the day-to-day demands of modern life. Therapy sometimes needs to become somewhat subversive
and conspiratorial as client and therapist attempt to work against all the forces of habit and social momentum
that keep us unhealthy. It has been said that the challenge of increased self-awareness is remembering that we
are more than our reflexes and defenses (Ouspensky, 1954).

Summary

Integration is vital and exists at each level of nature’s complexity from neurons to narratives to nations. As
systems become more complex, it takes more sophisticated mechanisms and increasing amounts of energy to
support their continuing interconnection and homeostatic balance. In this chapter we have explored the axes
of neural integration as well as the narratives that help us coordinate the government of systems that
constitute our brains and construct our conscious experience. The stories we use in psychotherapy also serve to
increase the complexity, coordination, and connectivity between us. This is one of the many connections
between interpersonal relationships and brain functioning that make psychotherapy a neuroscientific
intervention.

109
110
Part IV
Attachment and Connectedness

111
Chapter 11
The Social Brain
Our brains and bodies are designed to function in aggregates, not in isolation.
—John Cacioppo

Using evolution as an organizing principle, we begin with the assumption that our highly social brains have
been shaped by natural selection because banding together in groups enhances survival. The more tightly
interwoven we are as a group, the more eyes, ears, hands, and brains we have available to us. We know that
the expansion of the cortex in primates corresponds to increasingly large social groups and the development of
language, problem solving, and abstract abilities. Our larger and more complex brains not only allow for a
greater variety of responses to challenging situations and across diverse environments, but are also required to
process the vast amount of social information needed to support communication and group coordination
(Dunbar, 2014; Kanai et al., 2011).
Increasingly sophisticated social groups allowed for task specialization such as hunting, gathering, and
prolonged and dedicated caretaking. Caretaking specialization, in turn, produced longer postnatal
development and brains built by lived experience. So, while many animals need to be immediately prepared
upon birth to take on the challenges of survival, human infants have the luxury of years of total dependency as
they learn the complexities of the group. With the expanding size of primate groups, the grooming, grunts,
and hand gestures adequate in small groups were gradually shaped into spoken language. As social groups
grew even larger, more cortical geography was needed to process increasingly complicated social information.
This coevolution of relationships, language, and brain allowed for the development of higher levels of
symbolic and abstract functioning. In other words, early caretaking and intimate relationships are a
fundamental building block in the evolution of the human brain.
Despite the fact that our brains are social organs, Western science studies each individual as a single,
isolated organism rather than one embedded within the human community. This way of thinking leads us in
the West to search for technical and abstract answers to human problems instead of looking at day-to-day
human interactions. Take, for example, how physicians responded to the high mortality rate among children
in orphanages during the last century. Assuming that microorganisms were to blame, they separated children
from one another and ordered their handling to be kept to a minimum to reduce the risk of contamination.
Despite these mandates, the children continued to die at alarming rates, leading staff to fill out admissions
forms and death certificates during intake for the sake of efficiency. It was not until children were held and
played with by consistent caretakers and allowed to interact with one another that their survival rate improved
(Blum, 2002).
The notion of the brain as a social organ emerged in neuroscience during the 1970s as animal researchers
slowly began to appreciate that neuroanatomy, neurochemistry, and social relationships are inextricably
interwoven. The notion that primates possess neural networks specifically dedicated to social cognition was
initially proposed by Kling and Stecklis (1976), who found that damage to certain brain structures in primates
resulted in aberrations of social behavior and a decline in group status. Since then, scientists have been
exploring the varied neural terrain activated during social interactions. Subsequent research in the expanding
field of neuroscience has uncovered multiple sensory, motor, cognitive, and emotional processing streams that
contribute to interpersonal intelligence (Karmiloff-Smith et al., 1995).
Many of these findings have led to the growing realization that the lessons learned during a century of
dynamic psychotherapy may have important neuroscientific implications. The most basic is that we are born
into relationships and only come to our individual identity in the context social connectivity. Another is that
social interactions affect everything from our biology to our intellectual abilities. Neuroscience researchers are
slowly coming to the realization that the scope of their scientific observation needs to expand to include
relationships.
Neuroscientists already possess the perfect model for understanding interdependency—the individual
neuron. We know that neither the individual neuron nor the single human being exist in nature. Without

112
mutually stimulating interactions, people and neurons wither and die. In neurons, this process is called
apoptosis, while in humans it is called anaclitic depression. From birth until death, each of us needs others
who seek us out, show interest in discovering who we are, and help us to feel cared for and safe. Relationships
are our natural habitat, while the isolated brain is an abstract concept. Thus, understanding the brain requires
knowledge of the person embedded within a community of others. Therapists, teachers, and parents
intuitively grasp this profound reality just as laboratory scientists often do not. We are now in a position to
help research scientists know where to look as they explore how the brain grows, learns, and changes
throughout life.

The Social Synapse


Life is the continuous adjustment of internal relations to external relations.
—Herbert Spencer

As we discussed earlier, individual neurons are separated by small gaps called synapses. These synapses are
inhabited by a variety of chemical substances engaged in complex interactions that result in neural
transmission. This activity stimulates neurons to survive and modify themselves and each other. Over vast
expanses of evolutionary time, synaptic transmission has grown increasingly intricate to meet the needs of a
more complex brain. A parallel process has also been occurring in the evolution of the social synapse.
The social synapse is the space between us. It is also the medium through which we are linked together
into larger organisms such as families, tribes, and societies. When we smile, wave, and say hello, these
behaviors are sent through the space between us via sights, sounds, odors, and words. These electrical and
mechanical messages received by our senses are converted into electrochemical impulses within our brains.
These signals stimulate new behaviors, which, in turn, transmit messages back across the social synapse. From
the moment we are born, our very survival depends upon connecting to those around us through touch, smell,
sights, and sounds. If we are able to connect with nurturant others whose brains are primed to accept us as an
extension of themselves, then we can bond, attach, and survive.
The band of communication across the social synapse is extremely broad and includes unconscious
messages sent via posture, facial expression, eye gaze, pupil dilation, and even blushing. As we grow
increasingly interdependent, our inner experience becomes more visible through these and other means of
communication, in order to enhance the strength of our attachments (Cozolino, 2012). Contact with others
across the social synapse stimulates neural activation, which influences the internal environment of our
neurons. This activation in turn triggers the growth of new neurons as well as the transcription of protein,
which builds neurons as they expand, connect, and organize into functional networks. A basic assumption is
that loving connections and secure attachments build healthy and resilient brains, while neglectful and
insecure attachments can result in brains vulnerable to stress, dysregulation, and illness.
Early bonding experiences not only strengthen the networks of the social brain, they also promote the
building of the brain as a whole by stimulating metabolic arousal. Physical and emotional interactions between
mother and child result in a cascade of biochemical processes, enhancing the growth and connectivity of
neural networks throughout the brain (Schore, 1994). Face-to-face interactions activate the child’s
sympathetic nervous system and increase oxygen consumption and energy metabolism. Higher levels of
activation correlate with increased production and availability of norepinephrine, endorphins, and dopamine,
enhancing the child’s pleasure during positive connections (Schore, 1997a). The vital importance of these
early interactions to the building of the entire brain may help to explain the death of institutionalized children
deprived of interaction and love (Spitz, 1946).
You may remember from an earlier chapter that a sensitive period is a window of time when exuberant
growth and connectivity occur in specific neural networks. The onset and conclusion of these periods are
genetically and environmentally triggered, and correspond with the rapid development of skills and abilities
organized by each network. Thus, early experiences have a disproportionately powerful role in sculpting the
networks of attachment and affect regulation due to the strength of learning during these sensitive periods
(Ainsworth, Blehar, Waters, & Wall, 1978). Just as positive experiences equip us with feelings of self-
assurance and optimism, suboptimal bonding experiences become stored within implicit memory, carried into
adulthood, and woven into our adult relationships. Nowhere are these organizing principles more evident than
in psychotherapy.

113
Attunement and Reciprocity
Mirror neurons show how strong and deeply rooted is the bond that ties us to others.
—G. Rizzolatti and C. Sinigaglia

Attunement and reciprocity are aspects of the attachment process that reflect mutual awareness, turn taking,
and emotional resonance. Mother-infant emotional attunement during the first year is predictive of the
toddler’s self-control at two years, even when temperament, IQ, and maternal style are controlled for
(Feldman, Greenbaum, & Yirimiya, 1999). A mother’s ability to resonate with her infant’s internal states and
translate her feelings into words will eventually lead to the child’s ability to associate feelings with words. As
the child grows, the pairing of feelings with words enhances the integration of vertical and horizontal
networks dedicated to language and emotions. Early emotional regulation, established via mother-infant
synchrony, contributes to the organization and integration of neural networks and the eventual development
of self-regulation in the child.
Stage-appropriate attunement maximizes the possibility of neural growth, network coherence, and secure
attachment. The combined sense of safety, freedom from anxiety, and excitement generated via attunement
provides the affective background for the experience of vitality and spontaneous expression. For the newborn,
attunement may be communicated via stroking and cuddling; for a four-year-old, it means helping him or her
learn to share with a sibling. A 16-year-old, on the other hand, may need assistance with creating and staying
focused on goals for the future, while a 30-year-old will benefit from financial advice and some free
babysitting. This safe emotional background created by proper attunement, reciprocity, and loving-kindness
parallels an optimal educational and psychotherapeutic relationship.
The building of the social brain during the first two years is driven by the attunement between the right
hemispheres of the parent and the child (Schore, 2000). It is through this connection across the social synapse
that the unconscious of the mother is transferred to the unconscious of the child. The right-hemisphere-
biased circuits of the social brain come online at birth and appear to have their sensitive periods during the
first two years of life (Chiron et al., 1997). The mother seems to regress to a state of preoccupation with her
infant in the last months of pregnancy, and she continues in this state for a number of months after giving
birth (Winnicott, 1965c).
This maternal preoccupation involves an increased sensitivity to the visceral and emotional experience of
the child in order to attune to his or her primitive means of communication. The mother’s purposeful
regression allows her to lend her capacity to translate bodily states into words and actions that are soothing to
the infant. It is possible that the final trimester activates a shift to right-hemisphere dominance in the mother
to allow her to better attune to her newborn’s primitive communication. Suggestive evidence for this process is
that lateral dominance for emotional processing has been shown to shift during menstruation-related
hormonal changes (Cacioppo et al., 2013).

Jump-Starting Attachment
A mother understands what a child does not say.
—Jewish proverb

Even before birth, mothers and children engage in complex and reciprocal interactions. Communication
occurs through sound, movement, and touch, while their shared biochemical environment informs the child
about his or her mother’s state of mind and body. Prior to the formation of cortically organized social neural
networks, we possess a number of primitive reflexive behaviors that jump-start and stimulate the development
of the more sophisticated forms of attachment behavior to come. These reflexes reach across the social synapse
and allow us to become quickly integrated with our parents. The process of transmitting the communication
style of the mother, family, and culture begins immediately at birth.
Within the first hours after birth, newborns open their mouths and stick out their tongues in imitation of
adults, and after 36 hours they are able to discriminate among happy, sad, and surprised facial expressions
(Field, Woodson, Greenberg, & Cohen, 1982). Seeing happy faces causes newborns to widen their lips, while
sad faces elicit pouting, and surprised expressions result in wide-open mouth movement. Infants look
primarily at the mouth for happy and sad faces, whereas they alternate between the eyes and mouth in
response to expressions of surprise, suggesting they are capable of selecting different visual targets based on
the types of information presented to them (Field et al., 1982).

114
Over 20 involuntary reflexes have been identified in the newborn. Some—like the rooting and sucking
reflexes—help the infant obtain nurturance, while the palmar grasp (automatic hand grasp) and the Moro
reflex (reaching out of the arms) help the child hold onto the caretaker. These early reflexes, controlled by the
brain stem, are gradually inhibited by the cortex and replaced by conscious, flexible, voluntary behavior.
Reflexes such as these increase the newborn’s chances of survival by enhancing his or her physical and
emotional connection to mother and father. The old image of the infant as a passive recipient of stimulation
has been replaced with a view of the infant as a competent participant in the social environment.
One of my clients told me of his first interaction with his son: “A few seconds after he was born, the nurse
handed him to me and told me to put him in a small bed under a heat lamp. I dutifully crossed the delivery
room and gently placed him under the lamp. The light was very bright and he squinted hard, making his face
look like a bunch of wrinkles. I put my hand over his face to shield his eyes and he instinctively reached up
and took my thumb in his left hand and my pinky in his right and pulled my hand onto his cheek. He was
now about 90 seconds old and had become my son. I felt the glow of pride about how clever he was, while
simultaneously feeling a surge of protectiveness. This was obviously a very intelligent child with a bright
future.” In this way, reflexes provide the dual function of creating physical connection and ensuring the
emotional investment of the adults upon whom the infant relies.
Although specific words are meaningless, the tone and prosody of the parents’ voices hold center stage.
Even strangers will instinctively raise the pitch of their voice when talking with babies to match their hearing
abilities. A mother reflexively holds her baby against her body after birth, maximizing skin contact and
helping the infant’s hypothalamus establish a set point for temperature regulation. The infant and mother
gaze into each other’s eyes, linking their hearts and brains, while nursing establishes the lifelong relationship
between nutritional and emotional nurturance (food equals love).
The warm and happy feelings associated with holding, touching, and nursing, the pain of separation and
the joy of reunion, are all stimulated through a variety of primitive neurochemicals that support bonding and
attachment. Through this biochemical cascade, mother-child interactions stimulate the secretion of oxytocin,
prolactin, endorphins, and dopamine, resulting in warm, positive, and rewarding feelings (Love, 2014; Rilling
& Young, 2014). These biochemical processes, in turn, stimulate neural activation and the structural
maturation of the brain while shaping attachment circuitry (Fisher, 2004; Panksepp, 1998).
The secretion of endogenous endorphins results in feelings of well-being and elation. It actually does feel
better when a loved one kisses your boo-boo because endorphins are also natural analgesics. These opiates are
strongly reinforcing and serve to shape our preferences from early in life (Kehoe & Blass, 1989). Research
with primates suggests that the activation of the opioid systems of mother and child propels and regulates the
attachment process. When parent-child primate pairs engage in touching and grooming behavior, endorphin
levels increase in both (Keverne, Martens, & Tuite, 1989). During separation, the administration of
nonsedating morphine has the same soothing effect on the infant as does the reappearance of the mother.
When naltrexone (a drug that blocks the effects of endogenous opioids) was administered to infant primates,
rodents, and dogs, proximity seeking increased (Kalin et al., 1995; Knowles, Conner, & Panksepp, 1989;
Panksepp, Nelson, & Siviy, 1994).
Reflexively orienting the head to the sound of the mother’s voice increases the possibility of eye contact
while the instinct to seek circles and complex figures directs the baby’s attention toward the mother’s eyes and
face. Prolonged mutual gazing stimulates metabolic activity and neural growth, while reflexive smiling evokes
positive feelings and expressions in caretakers, further stimulating the infant’s brain.
Close examination of the bidirectional protoconversation between a mother and her baby demonstrates that
infants have far more influence on their mothers than previously thought (Bateson, 1979). A baby does not
simply react to its mother, but instead learns how to affect her feelings and behaviors. Both mother and infant
adjust to each other’s gestures, behaviors, and sounds in a sort of lyrical song and dance (Trevarthen, 1993). It
is through this language of intersubjectivity that children learn from their mothers about the fundamental
safety or dangerousness of the world. Protoconversation over the first year of life serves as the interpersonal
and emotional scaffolding upon which semantic language and narratives will gradually emerge. The growth
spurt of the right hemisphere provides the neural substrate for the development of the emotional components
of language.

The Importance of Eyes


There is a road from the eye to heart that does not go through the intellect.
—G. K. Chesterton

115
The eyes are a primary point of orientation for infants. They play a significant role in bonding and social
communication. Throughout the animal kingdom, eyes play a crucial role in determining the safety or danger
posed by others. Gaze aversion (visual cutoff) is an important social behavior that indicates dominance
hierarchy in both primates and humans. Direct eye gaze is a threat signal in primates (De Waal, 1989), and
the recognition that we are being looked at results in increased heart rate and amygdala activation (Nichols &
Champness, 1971; Wada, 1961). What must it be like for primates trapped in zoos who have hundreds of
human primates filing by and staring at them each day? Robert De Niro’s “Are you lookin’ at me?” monologue
in Taxi Driver is a dramatic example of the relationship between eye gaze, threat, and dominance.
Learning the language of eyes provides us with valuable information about our environment and what
might be on the minds of others. We reflexively look up when we see other people doing so; in these
situations, the eyes serve as a source of social communication about possible threats in our environment.
Elaborate neural circuits have evolved to monitor the direction of eye gaze of potentially dangerous others in
order to anticipate their next move. On the other end of the spectrum, the connection between the eyes, the
visual system, and emotion can be easily witnessed in the delight a child takes in a game of peek-a-boo.
Thanks to the neurochemistry of bonding, the smiles and laughter elicited from a child during peek-a-boo are
just as addicting for adults. There is a surge of good feelings in both children and caretakers with each
reappearance of the eyes. Similarly, consider the way two people in love can stare endlessly into each other’s
eyes, constantly recharging feelings of happiness.
During infancy, mutual gaze between caretaker and child is a primary mechanism for promoting brain
growth and organization. In their exploration of the environment, toddlers regularly check back to see the
expression on their parent’s face. If the parent looks calm, the child will feel confident to explore further. A
frightened look from the parent may result in the child seeking proximity and decreasing exploration. This use
of the eyes and facial expressions to encourage or inhibit toddler activities is referred to as social referencing
(Gunnar & Stone, 1984).
In therapy, the way a patient experiences your gaze (as caring or threatening) is an aspect of transference
that may provide important cues to early bonding experiences. An identical expression will, for some patients,
lead to a request that the therapist not stare at them, while it will make others feel attended to and cared for.
Although some patients prefer to lie down and look away from the therapist, others want to keep an eye on
you. These reactions reflect the eyes’ ability to elicit emotions from the patient’s interpersonal history stored
within networks of implicit memory. Thus, exploring the clients’ reaction to your gaze may yield valuable
information.

Recognizing Faces and Reading Facial Expressions


Laughter is the sun that drives winter from the human face.
—Victor Hugo

A vital function of the social brain is to recognize faces and assign a value to them; in other words, are they
familiar or strange, friend or foe—should I stay or should I go? This involves both determining identity (who
is this?) and using facial expressions to guess the other person’s emotional state and intentions (what are they
up to?). The first part of this process involves the complex task of recognizing a face from all possible angles,
an analysis that is easy for a child but continues to elude the fastest computers. Although the recognition of
faces involves both hemispheres, it is a function most suited to the visual-spatial mechanisms and holistic
processing strategies of the right hemisphere.
Research with primates has demonstrated that a particular region of the temporal cortex contains cells that
are responsive to faces, their identity, and various facial expressions (Perrett et al., 1984). Neurons activated
specifically by faces have also been detected in the amygdala connecting the reading of others’ faces to our own
autonomic reactions, emotions, and behaviors (Leonard, Rolls, Wilson, & Baylis, 1985; Perrett, Rolls, &
Caan, 1982). The temporal cortex contributes its abilities for complex recognition tasks (e.g., the countless
combinations of facial features), while the amygdala and the ompfc add the emotional elements to processing
social information. Together they give us the ability to approach friendly faces and make us wary of potential
enemies.
Our temporal lobes contain neurons dedicated to faces that are essential to our ability to relate to others.
Besides being able to recognize faces and the behaviors of others, we need to experience other people as being
different from inanimate objects. You are probably familiar with autism, a disorder characterized by profound
deficits in the ability to relate to others. In interacting with individuals suffering from these disorders, I have

116
been left with the feeling that, to them, I am no different from any other object in the room. Not surprisingly,
research has demonstrated that individuals with autism process faces in an area of the right temporal lobe
normally used to process objects (Schultz et al., 2000). This finding reflects one of the many neuroanatomical
mechanisms underlying profound disorders of relationship.

Mirror Neurons
Behavior is the mirror in which everyone shows their image.
—Johann Wolfgang von Goethe

Another way in which we link up across the social synapse is with the help of what are called mirror neurons.
Let me first describe how they were discovered. Using microsensors, neuroscientists are able to record the
firing of single neurons in monkeys’ brains. This recording can take place while they are aware, alert, and
interacting with other monkeys. Through such methods, neurons were discovered in the premotor areas of the
frontal cortex that fire when another primate or the experimenter is observed engaging in a specific behavior,
such as grasping an object with a hand (Jeannerod et al., 1995). Some of these neurons are so specific that they
only fire when an object is grasped in a certain way by particular fingers (Rizzolatti & Arbib, 1998). What is
even more interesting is that these very neurons fire when the monkey itself performs the same action
(Gallese, Fadiga, Fogassi, & Rizzolatti, 1996).
These neurons have been dubbed mirror neurons because they fire both in response to an observation of a
highly specific relationship between an actor and some object and when the action is performed by the
observer. Thus, mirror neurons serve to connect our visual and motor systems with frontal systems responsible
for goal-directed behavior. For obvious reasons, the same sort of studies are not possible in healthy human
subjects. However, noninvasive scanning technologies have been used to extend these findings to human
brains. One such study demonstrated that areas in our brain analogous to those containing mirror neurons in
primates are activated during both the observation and the execution of hand actions (Nishitani & Hari,
2000). Support for the relationship between these areas in the monkey and Broca’s area in humans comes
from positron emission tomography studies showing activation in Broca’s area during the active or imagined
carrying out of hand movements (Bonda, Petrides, Frey, & Evans, 1994; Decety, 1994; Grafton, Arbib,
Fadiga, & Rizzolatti, 1996).
The fact that mirror neurons fire when the same action is observed or performed leads to some interesting
hypotheses about their role in learning and communication. It has always been known that both humans and
primates can learn by observation. Because mirror neurons activate for both observation and action, they may
be the mechanism for one-trial learning. Also, because these neurons have been found in Broca’s area in
humans, mirror neurons may be involved in the imitation, learning, and expression of language (Gallese et al.,
1996). Shared actions and turn taking may have been the genesis of protoconversation and semantic language.
Some language learning may be jump-started by these mirror neurons within Broca’s area, as the sounds and
lip movements of caretakers are imitated. The alternation of mirroring and turn taking seen in mother-infant
interactions may be a contemporary reflection of the early evolution of language (Iacoboni, 2008; Rizzolatti &
Sinigaglia, 2008).
The most interesting application of mirror neurons to psychotherapy is that the facial expressions,
gestures, and posture of another will activate circuits in the observer similar to those which underlie empathy.
Seeing a sad child cry makes us reflexively frown, tilt our heads, say “aww,” and feel sad too. Watching an
athlete walk off the field with his head held high and chest pushed out can lead us to feel energized and
proud. In these and other ways, mirror neurons may bridge the gap between sender and receiver, helping us
understand one another and enhance the possibility of empathic attunement (Wolf, Gales, Shane, & Shane,
2000). The internal emotional associations linked to mirror circuitry are activated via outwardly expressed
gestures, posture, tone, and other pragmatic aspects of communication. Thus, our internal emotional state—
generated via automatic mirroring processes—can become our intuitive theory of the internal state of the
other. These structures are at the core of our ability to develop intimate relationships, be attuned to one
another, and aid our children in shaping a healthy and balanced sense of self.

Winnicott and the Emergence of the Person


Many patients need us to be able to give them a capacity to use us.

117
—Donald Winnicott

Donald Winnicott, an English pediatrician and psychoanalyst, developed some basic principles that provide a
helpful way of thinking about the social processes which shape these neural structures. His work with mothers
and children led him to coin terms such as good-enough mothering, holding environment, and transitional
object, which have become part of the basic lexicon of child development. His ideas have been highly
influential because of both their relevance to everyday experience and their freedom from obscure jargon.
Winnicott described the core of mothering as providing a facilitating and holding environment, which
requires both the mother’s empathic abilities and respect for the autonomy of the child. A mother’s devotion
to her child allows her to offer an expanding scaffolding that constantly adapts to her child’s changing needs
and abilities. Winnicott defined the early and intense focus on the baby as primary maternal preoccupation,
and he understood it to include the mother’s absorption into and attunement to the experiences with her
baby’s primitive developmental state. In this process, she utilizes the biochemistry of attachment and the
circuits of the social brain to bridge the social synapse between herself and her child. The good-enough
mother, in Winnicott’s (1965b) thinking, is a mother who does an adequate job in this difficult, complex, and
constantly shifting process of adaptation.
Winnicott believed that to talk of an infant separately from its mother was a theoretical abstraction. What
actually exists is a symbiotic infant-mother dyad within which the child is nurtured and its social brain is
formed, and from which the infant eventually emerges as an individual psychological being. Because an
internalized mother and the representation of the mother-infant dyad remain as organizing principles of the
social brain, they continue to impact us throughout our lives. In this way, an adolescent or adult with good-
enough mothering is never really alone.
A central component of development from Winnicott’s perspective depends on the mother’s ability to
mirror her child. Mirroring is the process by which a mother attunes to her child’s inner world and gives form
to his or her formless fantasies, thoughts, and needs. Mirroring serves the purpose of taking the disorganized
processes within the child, naming them, and making them a part of the relationship. The child then learns
about his or her inner world through the relationship. Although many decades before their discovery,
Winnicott was describing a process that relies on mirror neurons to support this deep attunement between
mother and child.
It is not uncommon for women in the third trimester of their pregnancy or in the first months after giving
birth to report that they feel they have lost IQ points. Although these changes are often attributed to the
effects of hormones and sleep deprivation, they may also be related to a shift in bias to the right hemisphere.
A shift away from logical and orderly left-hemisphere thinking to right-hemisphere-biased processing may
allow the mother an increased level of emotional and physiological sensitivity that enhances the intuitive
elements of attachment. A shift of brain coherence toward the right hemisphere would explain the decrease in
linear semantic processing and memory abilities reported by new mothers and mothers-to-be. Although such
a shift might be very useful for attunement with an infant, it could be detrimental to functions best performed
by the left, such as finding the right words, remembering appointments, and following logical arguments.
Many new mothers report an increasing need during the first year to get out into the world of adults or back
to work. This need may parallel a shift back to previous levels of left-right hemisphere balance.
As the mother gradually recovers from a deep preoccupation with her infant and again becomes interested
in other areas of life, the child is forced to come to terms with some of his or her own limitations. In an
appropriately attuned parent, a graduated failure of adaptation will parallel the infant’s increasing abilities,
frustration tolerance, and affect regulation. Winnicott used the term impingement to describe the impact on
the child of maternal misattunements. These can take the form of not appropriately anticipating the child’s
needs, interfering with the need for quiet and calm, and even underestimating his or her abilities. Parents have
to fail to adapt in different ways in order for their children to face the challenges necessary for adequate
development.
Minor impingements are challenges that create moderate and manageable levels of stress which the child is
able to cope with and master. These experiences likely promote and may even maximize brain growth and
neural network integration. Major impingements overwhelm the child’s ability to cope and integrate
experience in a cohesive manner, resulting in dissociated networking and functional disabilities. Gradual
minor impingements force the infant to grow, whereas major impingements can result in derailment of
positive adaptation and the solidification of defense mechanisms. Minor impingements are learning-
enhancing experiences, whereas major impingements result in decreased neural integration and hamper the
child’s development.
One of Winnicott’s most clinically useful concepts has been the idea of the development of a true and false
self. Secure attachments and a sense of a safe world create the context for the development of the true self,

118
which represents those aspects of the self that develop in the context of manageable (minor) impingements,
support, encouragement, and proper meaning from the caretaker. Respect for the autonomy and separateness
of the child motivates the parent to discover the child’s interests, instead of imposing his or her own. The true
self reflects our ability to tolerate negative feelings and to integrate them into conscious awareness and to seek
out what feels right for us in our activities, ourselves, and our relationships with others. Winnicott’s true self is
obviously one in which neural network development has been maximized, affect is well regulated, and
emotions and cognition are well integrated. The true self reflects an open and ongoing dialogue among the
heart, the mind, and the body.
What Winnicott called a false self results from major impingements for which the child is unprepared.
Prolonged impingements can result in chronic emotional dysregulation. For example, neglect, abuse, or
continuous states of shame can overwhelm the child’s natural development and lead to the dominance of
emotional defenses. These stressful relationships will also inhibit neurogenesis and proper brain development
(Stranahan, Kahlil, & Gould, 2006). When self-involved or pathological parents use children for their own
emotional needs, the child can become compulsively attuned to the parents, creating a false self designed to
regulate the parents’ needs. Without appropriate assistance in developing his or her self-reflective capacity,
such children live through reflexive social behavior and never learn that they have feelings and needs of their
own that should be expressed and nurtured. Winnicott understood therapy most generally as a process of
controlled regression to a childhood state with the purpose of succeeding in developing a true self in the
present, which was thwarted in early life (St. Clair, 1986).

Shame
Every word, facial expression, gesture, or action on the part of a parent gives the child some message about self-worth. It is sad that so many
parents don’t realize what messages they are sending.
—Virginia Satir

During the first year of life, healthy parent-child interactions are primarily positive, affectionate, and playful.
Due to their limited mobility, infants stay in close proximity to caretakers, who provide for their many bodily
and emotional needs. As the infant transforms into a toddler, a parent’s role comes to include protecting the
child from danger such as falling down stairs, being bitten by stray dogs, or drinking fabric softener. The
emergence of normal, incessant exploratory behavior in toddlers is driven by the brain’s intense need for
stimulation and growth. Due to the toddler’s increasing motor coordination and exploratory drive, parents
find themselves protectively saying “no” beginning at about 18–24 months (Rothbart, Taylor, & Tucker,
1989). Affection and attunement, experienced as unconditional during the first year, come to be tied to limit
setting, control, and early attempts at discipline.
Shame, appearing early in the second year of life, is both a powerful inhibitory emotion and a mechanism
of social control. Thus the positive face-to-face interactions that stimulated excitement and exhilaration
during the first year come to include expressions of disapproval and anger. Shame is represented
physiologically by a rapid transition from a positive to negative affective state and from sympathetic to
parasympathetic dominance. This shift is triggered by the expectation of attunement to a positive state, only
to receive negative emotions from the caretaker (Schore, 1994). While it may be hard to believe, toddlers
expect their parents to be just as excited as they are about covering the floor with milk or plopping their toys
in the toilet. Parental reactions of disapproval or anger are, at first, confusing and difficult to comprehend but
soon come to shape the biology and psychology of the child (Matos, Pinto-Gouveia, & Costa, 2011; Matos &
Pinto-Couveia, 2014).
Behaviorally, people in a shame state look downward, hang their head, and round their shoulders. This
same state (submission) is shown by your pet dog when he hunches over, pulls his tail between his legs, and
slinks away as you upbraid him for some canine faux pas. Similarly, this posture in humans reflects social
exclusion, loss, and helplessness. During early socializing experiences, shame is the emotional reflection of a
lost attunement with the caretaker, drawing its power from the child’s primal need to stay connected for
survival. Prolonged and repeated shame states result in physiological dysregulation and negatively impact
affect regulation, attachment, and the development of networks of the social brain (Schore, 1994).
The return from a state of shame to attunement results in a rebalancing of autonomic functioning,
supports affect regulation, and contributes to the gradual development of self-regulation. Repeated and rapid
return from shame to attuned states also consolidates into an expectation of positive outcomes during difficult
social interactions. These repairs are stored as visceral, sensory, motor, and emotional memories, making the

119
internalization of positive parenting a full-body experience. Thus, the continual experience of attunement,
misattunement, and reattunement creates a kind of body memory that becomes an expectation of a positive
outcome for relationships and life. Children left in a shamed state for long periods of time may develop
permanently dysregulated autonomic functioning along with depression, hopelessness, and despair. As the
child graduates into increasingly complex peer group relations, these same physiological processes are
connected to popularity, social status, and dominance within groups at school and on the playground.
Because shame is a powerful, preverbal, and physiologically based organizing principle, the overuse of
shame in the process of parenting can predispose children to developmental psychopathology related to affect
regulation and identity (Schore, 1994; Schore & Schore, 2008). As part of his therapeutic programs, John
Bradshaw (1990) refers to “inner child work” as addressing the long-standing power of these early shame
experiences, which he calls “toxic shame.” Shame needs to be differentiated from the later-occurring
phenomenon of guilt. Guilt is a more complex, language-based, and less visceral reaction that exists in a
broader psychosocial context. Guilt is related to unacceptable behaviors, whereas shame is an emotion about
the self that is internalized before the ability to distinguish between one’s behavior and one’s self is possible. If
guilt is “I did something bad,” then shame is “I am bad.” We see this often, in individuals who spend their
lives taking care of others and doing good deeds in an attempt to make up for some sin that they cannot recall.

The Consolidation of the Self


Never be afraid to sit awhile and think.
—Lorraine Hansberry

In Winnicott’s view, too many impingements prevent the infant from experiencing what he called formless
quiescence: those moments of safety and calm that teach the child the world can be a safe place. It is in these
quiet moments that the experience of self is consolidated, neural networks integrate, and fantasy and reality
are gently combined. In essence, good-enough parenting results in the belief in a benign world where one is
safe to build an internal experience of self (Winnicott, 1965a). Thus, Winnicott felt that a major achievement
of early attachment was the capacity to be alone, an ability learned by being alone in the presence of a
competent caretaker. These experiences create enough security to allow feelings in the child to spontaneously
bubble up with the confidence that they will be manageable and understandable. In this state of mind, the
need to employ defenses to cope with external threat and inner emotions is at a minimum. At the same time,
parietal-frontal systems involved in imagination and the creation of an inner sense of self become activated.
The manic defenses we often see in our clients result from the lack of the capacity to be alone. Impulsive
behaviors and thoughts, disconnected from self-reflective processes, serve to inhibit emotions because to these
individuals, to feel is to feel badly (Miller, Alvarez, & Miller, 1990). Slowing down stimulates discomfort,
sadness, isolation, and shame, which become background affect throughout life. If manic defenses are
chronically employed, they can become a way of life and keep children and adults from constructing inner
imaginal experience and a sense of self. Sadly, many children with manic defenses are mistakenly diagnosed
with ADHD. They are medicated to help them cope, while the real problem goes unresolved.
People with manic defenses often mask their inability to be alone by stirring up a constant whirlwind of
activities, social interactions, and phone calls. Despite their outward success, and their narcissistic and
grandiose attitudes, they often have great difficulties in relationships and report feelings of despair and
emptiness. Exploration of their histories usually points to patterns of insecure attachments in which
achievement served as the currency for acceptance. Constantly escalating levels of activity are reinforced by
praise from others and the avoidance of the negative feelings that bubble up when the patients are quiet or
alone. These people often have a hard time relaxing or taking a vacation because the lack of distraction leaves
them open to the intrusion of uncomfortable feelings for which they have no effective coping skills.
The inability to be alone is seen most clearly in individuals with borderline personality disorder, who have
a catastrophic reaction to real or imagined abandonment. For these people, separation is experienced as a
threat to their very survival in much the same way as an infant reacts to the absence or loss of a parent. Their
catastrophic reaction in adulthood is likely the activation of an implicit memory of overwhelming
abandonment fears from a time before object constancy or self-regulation. It is as if the child within these
patients is in a holding pattern, awaiting proper parenting. The extremely emotional life-and-death reactions
in borderline clients may be our best window to the chaotic and often frightening emotional world of early
childhood.

120
Summary

The brain is a social organ connected to other brains via the social synapse. Primitive reflexes jump-start the
attachment process and are gradually replaced by voluntary behaviors. The motivation to stay connected is
driven by biochemical systems we share with our primitive ancestors. While there are multiple channels of
communication between us, vision is an important link across the social synapse and the expressive face a focal
point of social information. Theories of psychological development by Winnicott, Freud, and others provide
us with models for the development of mind embedded in these more basic neurobiological processes. The
development of a sense of self requires periods of freedom from external threat and inner turmoil. It also
requires the development of frontal-parietal systems responsible for inner imaginal space. Children constantly
buffeted by external chaos can remain trapped in a selfless state where they are witness to internal impulses
and external behaviors with little or no ability to either understand or control what they are doing.

121
Chapter 12
Building the Social Brain:
Shaping Attachment Schemas
Experience is a biochemical intervention.
—Jason Seidel

While Winnicott observed and worked with mother-infant pairs in his consulting office, John Bowlby was
performing naturalistic observations of primates in the wild and children in orphanages. He was especially
interested in mother-child bonds, the importance of exploratory behavior, and the impact of separation and
loss on healthy development. His experiences led him to develop the concepts of attachment figures, proximity
seeking, and a secure base (Bowlby, 1969). Bowlby’s observations and the subsequent scientific findings in
attachment research are easily integrated with Winnicott’s theories of bonding and attachment.
Bowlby’s work, which highlighted the importance of specific caretakers to a child’s sense of security,
resulted in a major shift in the care of institutionalized children. To encourage bonding, children who had
previously been cared for by whomever was available were now assigned consistent caretakers. In addition, this
change in attitude changed the role of nurses and caretakers from only custodians of small babies into
emotional attachment figures. In essence, they were told that becoming attached should not be avoided.
Subsequently, Mary Ainsworth and her student, Mary Main, developed research methods to test Bowlby’s
theories. Decades of attachment research followed, providing us with some fascinating tools to study the
sculpting of the social brain during childhood, as well as the long-term impact of early experiences later in life.
Bowlby suggested that early interactions create attachment schemas that predict subsequent reactions to
others. Schemas are implicit memories that organize within networks of the social brain, based on experiences
of safety and danger with caretakers during early sensitive periods. A secure attachment schema enhances the
formation of a biochemical environment in the brain conducive to regulation, growth, and optimal
immunological functioning. Insecure and disorganized attachment schemas have the opposite effect, and
correlate with higher frequencies of physical and emotional illness.
Bowlby believed attachment schemas to be a summation of thousands of experiences with caretakers that
become unconscious, reflexive predictions of the behaviors of others. Attachment schemas become activated
in subsequent relationships and lead us to either seek or avoid proximity. They also determine whether we can
utilize intimate relationships for physiological and emotional homeostasis. These implicit memories are
obligatory; that is, they are automatically activated even before we become conscious of the people with whom
we are about to interact. They shape our first impressions, our reaction to physical intimacy, and whether we
feel relationships are worth having. They trigger rapid and unconscious moment-to-moment approach-
avoidance decisions in interpersonal situations. Attachment schemas are especially apparent under stress
because of their central role in affect regulation. Attachment is mediated by the regulation of the autonomic
nervous system by the social brain, and a cascade of biochemical processes that create approach and avoidance
reactions as well as positive and negative emotions. Schemas shape our conscious experience of others by
activating rapid and automatic evaluations hundreds of milliseconds before our perceptions of others reach our
consciousness.
Empirical research into attachment schemas began with Ainsworth’s naturalistic in-home observations of
mothers interacting with their children (Ainsworth et al., 1978). These mothers were found to fall into three
categories: available and effective (free autonomous), dismissing and rejecting (dismissing), and anxious and
inconsistent in their attentiveness (enmeshed/ambivalent). The belief was that these different caretaking styles
would create differing coping and interpersonal styles in their children. So the next step was to determine
whether the children of mothers in each category displayed differences in their attachment behaviors,
especially when stressed or frightened.
The method developed to study the children’s attachment behavior is called the infant strange situation
(ISS). The ISS consists of placing an infant and its mother in a room, then having a stranger join them. After

122
a period of time, the mother exits the room, leaving the child alone with the stranger. Another brief period
follows, and then the mother returns. Children’s reunion behavior, or how they respond to the return of their
mother, is rated to determine their attachment style. This situation was chosen because of Bowlby’s
observation that being left alone with an unknown other evokes distress calls in young primates. The
attachment schema of the child, or the expectation of being soothed by the mother, should be aroused by the
stress of the situation and reflected in his or her reunion behavior. Does the child seek comfort from the
mother or ignore her? This research was begun with a number of questions: Does the child have a hard time
being comforted? Does the child soon feel safe and return to play, or is he or she anxious, clingy, or
withdrawn? These and other behaviors are the focus of the ISS scoring system and are thought to reflect the
child’s experience and expectation of the mother’s soothing capacity.
Four categories of the infants’ reactions to their mothers’ return have been derived from the ISS: secure,
avoidant, anxious-ambivalent, and disorganized. Furthermore, a relationship was found between ISS categories
and the maternal behavior originally derived from in-home observations. The general findings were as follows:
children rated as securely attached sought proximity with the mother upon her return, were quickly soothed,
and soon returned to exploratory or play behavior. These children, who were approximately 70% of the
sample, seemed to expect that their mothers would be attentive, helpful, and encouraging of their continued
autonomy. Securely attached children appeared to have internalized their mothers as a source of comfort,
using them to feel safe. These mothers were seen as effective in their interactions with their children and had
become “a background context for seeking stimulation elsewhere” (Stern, 1995, p. 103).
Avoidantly attached children tended to ignore their mothers when they returned to the room. They would
glance over to the mother as she came in, or shun contact altogether. These children tended to have
dismissing and rejecting mothers and appeared to lack an expectation that she would be a source of soothing
and safety. Avoidantly attached children behaved as though it was easier to regulate their own emotions than
seek comfort from their mothers, whose misattunement or dismissal might well compound their stress.
Children rated as anxious-ambivalent sought proximity, but they were difficult to soothe and slow to
return to play. Anxious-ambivalent children, who often had enmeshed or inconsistently available mothers,
may have had their stress worsened by their mothers’ distress. Their slow return to play and emotional
regulation may be a reflection of their mothers’ anxiety and lack of internalized safety. These children tended
to cling more and explore their environment less.
Finally, there was a group of children who engaged in chaotic and even self-injurious behaviors. On
reunion with their mothers, they demonstrated odd behavior such as turning in circles or falling to the
ground. They would freeze in place or be overcome by trancelike expressions. During later research, these
children were included in a fourth category called disorganized attachment. These chaotic behaviors were
demonstrated in conjunction with secure, avoidant, and anxious-ambivalent behaviors and were often present
in children whose mothers suffered from unresolved grief or trauma. Parents of children in this category
demonstrate frightened and frightening behavior to their children, inducing an alarm state in the child. In this
biological paradox, the child’s brain has an innate drive to move toward the mother. However, since the parent
is also a source of alarm, the child is faced with an approach-avoidance conflict. The resulting inner turmoil
dysregulates the child to the point that his or her adaptation and coping skills—even motor abilities—appear
to become disorganized. The fear and chaos of the mothers’ internal worlds can be observed in their children’s
behavior.
The transmission of trauma from parent to child is both powerful and insidious. A traumatized mother
who creates alarming experiences for her child leaves the child no choice but to stay with and depend on the
source of the alarm. The child’s safe haven is replaced with repetitive trauma by proxy and emotional
dysregulation (Olsson & Phelps, 2007). This process creates a new generation of victims. In research with
Holocaust survivors, indications of parental trauma were found to be reflected in the biochemistry of their
nontraumatized children (Yehuda et al., 2000; Yehuda & Siever, 1997). Further compounding the child’s
dysregulating environment, the trauma-related behaviors of victims will lead them to be avoided by other
children in normal social interactions. It is not surprising that children with avoidant and disorganized
attachment schemas are also shown to have higher levels of stress hormones and other biological markers of
trauma and sustained stress (Spangler & Grossman, 1993).

Parents Talk of Their Childhoods


A Freudian slip is when you say one thing but mean your mother.
—Unknown

123
The relationships discovered between attachment schemas and parenting style raised the question of whether
a parent’s early attachment experiences influenced parenting style decades later. While it was assumed that the
parenting styles of adults are somehow shaped by childhood experiences, there was no empirical support for
this transfer from one generation to the next. Because implicit memory is inaccessible to our conscious mind,
and explicit memories of childhood are shaped by so many emotional factors, a measure was needed that could
bypass the usual distortions of memory and all of our defense mechanisms. An extremely interesting research
tool that appears to have succeeded in this task is the Adult Attachment Interview (AAI) (Main & Goldwyn,
1998).
The AAI consists of a series of open-ended questions about childhood relationships and early experiences
such as:

• I’d like you to try to describe your relationship with your parents as a young child . . . if you could start
from as far back as you can remember.
• Choose adjectives that reflect your relationship with your mother, father, and so on.
• Which of your parents did you feel closest to and why?

Although the AAI gathers information about what individuals remember of their childhood, it also
provides the data for a linguistic analysis of the coherence of the narrative’s organization and presentation.
Coherence analysis is conducted based on what are called Grice’s maxims and include an examination of both
the logic and understandability of the narrative based on the following four principles:

1. Quality: Be truthful, and have evidence for what you say.


2. Quantity: Be succinct, and yet complete.
3. Relevance: Stick to the topic at hand.
4. Manner: Be clear, orderly, and brief.

Scoring takes into account the integration of emotional and experiential materials, gaps in memory and
information, and the overall quality of the presentation (Hesse, 1999).
The AAI bypasses the left-hemisphere interpreter by examining the quality of the brain’s synthesis of the
various cognitive and emotional components of explicit and implicit memory. Siegel (1999) proposed that the
coherence of the AAI narrative parallels the level of neural integration attained during childhood, providing a
window to early attachment experiences and emotional regulation. In essence, the AAI gets at how individuals
put feelings into words, resolve traumatic experiences, and integrate the various networks of information
processing across emotion, sensation, and behavior in making sense of their lives. It does all of this while
simultaneously bypassing the problems inherent in self-report measures about the past.
Four categories emerge from the AAI that appear to correspond to the findings of the in-home
observations and the ISS. Mothers and fathers with securely attached children tended to have more detailed
memories, as well as a realistic and balanced perspective of their parents and childhood. Adhering well to
Grice’s maxims, they were able to describe these experiences in a coherent narrative that was understandable
and believable to the listener (Main, 1993). This group, called autonomous, demonstrated an integration of
cognitive and emotional memories, had processed their negative experiences, and was therefore more fully
available to their children.
The second group of parents, associated with avoidantly attached children, demonstrated a lack of recall
for childhood events and large gaps in memory for their childhood. This lack of recall is believed to reflect a
disruption of the integration of cognitive and emotional elements of autobiographical memory. This could be
due to trauma, chronic stress, or a lack of assistance in learning to regulate affect from their own parents early
in life. They also demonstrated an overall dismissing attitude toward the importance of their early
relationships, just as they were dismissive of their own children now. The narratives of these parents were
incoherent both due to missing information and a tendency to either idealize or condemn their parents. They
gave the impression that they were defending against fully acknowledging their histories through denial and
repression.
A third group of parents, rated as enmeshed or preoccupied, tended to have anxious-ambivalently attached
children. Their narratives contained excessive, poorly organized verbal output that lacked boundaries between
the past and present. They appeared preoccupied and pressured and had difficulty keeping the perspective of
the listener in mind.
Last, the unresolved/disorganized group of parents had highly incoherent narratives disrupted by emotional

124
intrusions and missing or fragmented information. Their narratives not only reflected the disorganization of
verbal and emotional expression, but also the devastating impact early stress had on the development and
integration of their neural networks. The content of their narratives confirmed chaotic and frightening
childhood experiences which we can assume were devastating to the integration and homeostatic balance of
both body and brain. See Table 12.1 for a summary of attachment findings.
The power of the relationship between parent and child attachment patterns was demonstrated by Fonagy
and his colleagues when they administered AAIs to expectant first-time parents (Fonagy, Steele, & Steele,
1991). Over a year later, when the children reached their first birthday, their attachment patterns were
assessed using the ISS. In 75% of these cases, the child’s attachment pattern was predicted by the coherence of
the parents’ narratives and their attachment style many months before birth. Parents of infants who came to
be securely attached were able to provide a fluid narrative with examples of interactions, had few memory
gaps, and presented little idealization of the past. These parents did not seem to have significant defensive
distortions and were able to express negative feelings without being overwhelmed, and listeners tended to
believe what they were saying. It is not a big stretch to see that these parents were best able to provide the
kind of good-enough social environment which provides a balance between safety and challenge, attunement
and autonomy.
We now have some evidence that parents’ capabilities for attachment to an infant begin to take shape in
their own childhoods. Their skill as parents will depend on their empathic abilities, emotional maturity, and
neural integration: in essence, how they were parented as children. As a child, a young girl may begin to
imagine someday having children of her own. The shaping of her virtual children will be influenced by both
her fulfilled and unfulfilled needs. The empathy and care each parent received as well as the assistance they
experienced in articulating and understanding their inner worlds will influence future parenting abilities. A
mother’s childhood can determine whether she is prepared to provide emotionally for her newborn or if she
will unknowingly require her child to give her the attention she failed to receive when she was young (Miller,
1981).

TABLE 12.1
Summary of Attachment Findings

In-Home Observations of Mothers Infant Strange Situation Interview Adult Attachment

Free Autonomous Secure Autonomous


Emotionally available Infant seeks proximity Detailed memory
Perceptive and effective Easily soothed/returns to play Balanced perspective
Narrative coherency

Dismissing Avoidant Dismissing


Distant and rejecting Infant does not seek proximity Dismissing/denial
Idealizing
Infant does not appear upset Lack of recall

Enmeshed-Ambivalent Anxious-Ambivalent Enmeshed-Preoccupied


Inconsistent availability Infant seeks proximity Lots of output
Not easily soothed Intrusions, pressured, preoccupied
Not quick to return to play Idealizing or enraged

Disorganized Disorganized Unresolved/ Disorganized


Disorienting Chaotic Disoriented
Frightening or frightened Self-injurious Conflictual behavior
Unresolved loss
Traumatic history

Because attachment schemas are part of implicit memory, this level of caretaking occurs automatically and
connects our unconscious childhood experiences across the generations. In this way, a parent’s unconscious is
a child’s first reality. Interestingly, negative events in childhood are not necessarily predictive of an insecure or
disorganized attachment schema or future parenting style. Working through, processing, and integrating early
experiences, and constructing coherent narratives, are more accurate predictors of a parent’s ability to be a safe
haven for his or her children. This earned autonomy, through the healing of childhood wounds, appears to
interrupt the transmission of negative attachment patterns from one generation to the next.
The inference that parents who are rated as autonomous have higher levels of neural integration is based
on the fact that they are able to access and connect cognitive and emotional functioning in a constructive and
useful manner. They do not appear to be suffering the effects of unresolved trauma or dissociative defenses

125
and have attained a high degree of affect regulation, as demonstrated by their ability to meet the demands of
parenting with ongoing grace. They are able to remember and make sense of their own childhoods and are
available to their children both verbally and emotionally. Their children develop attachment schemas that
make them secure in the expectation that their parents are a safe haven and will soothe and assist them when
threats arise. Not surprisingly, parents’ emotional insight and availability to themselves appears to parallel
their emotional availability to their children.
The three insecure patterns of attachment research all reflect lower levels of psychological and neurological
integration. They also correlate with the use of more primitive psychological defenses associated with
disconnections among streams of processing within the brain. The lack of recall and black-and-white thinking
of the dismissing parent likely reflect blocked and unintegrated neural coherence. This brain organization
results in decreased attention and emotional availability to the child. The enmeshed parent has difficulty with
boundaries between self and others, as well as between past memories and present experiences. These internal
and interpersonal issues then lead to inconsistent availability and a flood of words that dysregulate the child.
Thus the child, who is also anxious and ambivalent, will seek proximity but have a difficult time returning to
play because of the unpredictable availability, as well as the confusing and emotionally dysregulating nature of
the parent’s messages and emotions. The internalized mother, instead of being a source of security and
autonomic regulation, becomes organized as a destabilizing state of mind and body.
Maternal and paternal instincts—in fact all caretaking behaviors—are acts of nurturance that depend upon
the successful inhibition of competitive and aggressive impulses. Too often, however, that inhibition is
incomplete and some of us are unable to be good-enough parents. When a parent abuses, neglects, or
abandons a child, the parent is communicating to the child that he is less fit for survival. Consequently, the
child’s brain may become shaped in ways that do not support his long-term survival. Nonloving behavior
signals to the child that the world is a dangerous place and tells him to not explore, discover, or take chances.
When children are traumatized, abused, or neglected, they are taught that they are not among the chosen.
They grow to have thoughts, states of mind, emotions, and immunological functioning that are inconsistent
with well-being, successful procreation, and long-term survival. With all due respect to the old adage, we
could say that what doesn’t kill us makes us weaker.
The tragedy of this lies in the fact that early experiences have such a disproportionately powerful impact on
the development of the infrastructure of the brain. As highly adaptive social organs, our brains are just as
capable of adjusting to unhealthy environments and pathological caretakers as they are to good-enough
parents. While our brains become shaped to survive early traumatic environments, many of these adaptations
may impede health and well-being later in life. Negative interpersonal experiences early in life are a primary
source of the symptoms for which people seek relief in psychotherapy.
Secure attachments represent the optimal balance of sympathetic and parasympathetic arousal, whereas
their imbalance correlates with insecure attachment patterns such as fight or flight and splitting (Schore,
1994). The balance of these two systems becomes established early in life and translates into enduring patterns
of arousal, reactivity to stress, and possible vulnerability to adolescent and adult psychopathology. Poor
attachment patterns lead to long-lasting emotional and physical over- or underarousal throughout the body
and the brain.
Secure and insecure attachment schemas are quite different. Securely attached children do not produce an
adrenocortical response to stress, suggesting that secure attachment serves as a successful coping strategy. On
the other hand, individuals with insecure attachment schemas do show a stress reaction, demonstrating that
insecure attachment is better described by a model of arousal rather than of successful coping (Izard et al.,
1991; Nachmias, Gunnar, Mangelsdorf, Parritz, & Buss, 1996; Spangler & Grossman, 1993; Spangler &
Schieche, 1998). In other words, the behavior of insecurely attached individuals is an expression of the state of
their autonomic arousal in response to fear.

Narrative Coconstruction
The wise man must remember that while he is a descendant of the past, he is a parent of the future.
—Herbert Spencer

Parent-child talk, in the context of emotional attunement, provides the ground for the coconstruction of
narratives. These narratives, in time, become the mass of our inner experience and the parameters of our
personal and social identities. When verbal interactions include references to sensations, feelings, behaviors,
and knowledge, they provide a medium through which the child’s brain is able to integrate the various aspects

126
of its experience in a coherent manner. The organization of autobiographical memory that includes input from
multiple neural networks enhances self-awareness and increases the ability to solve problems, cope with stress,
and regulate affect. This integrative process is what psychotherapy attempts to establish when it is absent.
Coconstructed narratives form the core of human groups, from primitive tribes to modern families. The
combined participation of caretakers and children in narrating shared experiences organizes memories,
embeds them within a social context, and assists in linking feelings, actions, and others to the self. The
creation and repetition of stories help children to develop and practice recall abilities and have their memories
shaped in relationships (Nelson, 1993). This mutual shaping of memory between child and caretaker can serve
both positive and negative ends. Positive outcomes include teaching the importance of accuracy of memory,
imparting of cultural values, and shaping the child’s view of herself based on her role in the story. Negative
outcomes include the transfer of the caretakers’ fears and anxieties into the child’s narratives so that they
become central themes in the experience of the child (Ochs & Capps, 2001).
When caretakers are unable to tolerate certain emotions, they will be excluded from their narratives or be
shaped into distorted but more acceptable forms. The narratives of their children will come to reflect these
editorial choices. At the extreme, parents can be so overwhelmed by the emotions related to unresolved
trauma that their narratives become disjointed and incoherent. On the other hand, narratives that struggle to
integrate frightening experiences with words can serve as the context for healing by simultaneously creating
cortical activation and increasing descending control over subcortically triggered emotions. Parental narratives,
both coherent and incoherent, become the blueprint not only for the child’s narratives, but for the
organization and integration of neural circuitry. As it turns out, there appears to be a relationship between the
complexity of a child’s narratives, self-talk, and the security of the child’s attachment.
Main, Kaplan, and Cassidy (1985) studied a group of six-year-old children who, at one year, were assessed
in the infant strange situation. They discovered that securely attached children engaged in self-talk during
toddlerhood and spontaneous self-reflective remarks at age six. They also tended to make comments about
their thinking process and their ability to remember things about their history. These processes of mind,
which insecurely attached children often lack, reflect the utilization of narratives in the development of self-
identity. They also point to a more sophisticated ability to metacognize (think about thinking), which
represents a high level of neurolinguistic self-regulation. What we are witnessing appears to be the
internalization of their parents’ self-regulatory mechanisms. As you might expect, child abuse correlates with
less secure attachment patterns in children and a decreased ability to talk or think about their internal states
(Beeghly & Cicchetti, 1994).
Fonagy, Steele, Steele, Moran, et al. (1991) studied the relationship between infant security and reflective
self-functioning in mothers and fathers. They found a strong correlation between measures of self-reflection
and narrative coherence. In fact, when reflective self-function was controlled for in the statistical analysis,
coherence no longer related to infant security. This suggests that the relationship between coherence and
reflective self-functioning is powerful, and that self-reflection plays an important role in the integration of
multiple processing networks of memory, affect regulation, and organization. In discussing these results, the
researchers suggested, “The caregiver who manifests this capacity at its maximum will be the most likely to be
able to respect the child’s vulnerable emerging psychological world and reduce to a minimum the occasions on
which the child needs to make recourse to primitive defensive behavior characteristic of insecure attachment”
(p. 208).
When the parents’ inability to verbalize internal and external experiences leaves the child in silence, the
child does not develop a capacity to understand and manage his or her own inner and outer world. The ability
of language to integrate neural structures and organize experience at a conscious level is left unutilized. When
children with nonhealing parents experience trauma early in life, the stress of each new developmental
challenge is multiplied. In the same way, language, in combination with emotional attunement, is a central
tool in the therapeutic process, creating the opportunity for neural growth and neural network integration.
A child who is able to achieve this ability with the help of someone other than the primary caretaker may
be able to earn a higher level of integration and security than would be predicted by his or her parents’ rating
on the AAI. This may come from other significant people in the child’s environment who are able to attune to
the child’s world and assist in the child’s articulation of his or her emotional life. This might explain some of
the earned autonomy seen in parents with negative childhood experiences but with coherent narratives and the
ability to provide a safe haven for their children. Earned autonomy is convincing evidence that early negative
experiences can be reintegrated and repaired later in life. Personal growth has the ability to heal because the
social brain remains plastic.
Attachment patterns formed in childhood can be relatively stable into adulthood and have been shown to
impact romantic love, interpersonal attitudes, and psychiatric symptoms (Brennan & Shaver, 1995; Hazan &

127
Shaver, 1990). Adult children of anxious parents repeatedly return to them throughout life, still seeking
comfort and a safe haven. Many of these children become parents to their parents, taking care of the people
that should take care of them. They continue to return to an empty well; each time the bucket is lowered,
there is the hope that it will contain the nurturance they need. Each returning empty bucket reinforces their
lack of safety.

Child Therapists
Treat people as if they were what they ought to be and you help them to become what they are capable of being.
—Goethe

A question that commonly arises among therapists, adoptive parents, and mental health legislators is, “When
is it too late?” At what age do the negative effects of early abuse, trauma, and neglect become permanent?
Getting to the heart of the issue, the true question becomes: Who is worth seeing as a client, adopting as a
child, or investing public funds in for rehabilitation? In my mind, these are moral rather than scientific
questions. I have become very skeptical of experts who think that they have found answers to any issue in
neuroscience. My bias is to trust in plasticity and our own ingenuity to discover new solutions to these
problems. Here is a study that may give us some guidance as we consider these issues.
To test the impact of maternal deprivation, Harry Harlow isolated newborn monkeys not only from their
mothers, but from any and all contact with other monkeys. Beyond minimal contact related to taking care of
their basic needs, these young monkeys were left alone in a cage with a few toys and little else. Pictures of
these isolate monkeys are heartbreaking—they huddle in corners, rock, bite themselves, and stay curled up in a
fetal position. It is as if they are trapped in hell waiting to be born into the social world.
When these isolate monkeys were then introduced into a colony at six months of age, they were
understandably terrified. They didn’t seem to understand what was going on, retreated from curious others,
and did their best to avoid interaction. At first it was tempting to think of this six-month period as a cutoff
point for attachment plasticity. Perhaps attachment circuitry had gone through a hard-wired critical period
and, by six months, it was too late to learn how to be social. But as with every conclusion in neuroscience,
there needs to be caution.
Harlow and Suomi (1971) wondered if therapy could help these isolate monkeys overcome their fear and
allow them to join the social world of the colony. But how do you do therapy with a monkey? Would you
choose Gestalt, cognitive-behavioral, or psychoanalytic treatment? Ultimately what was chosen was a
combination of play and attachment therapy. The therapists chosen for the job were normal three-month-old
monkeys, who were selected because they were smaller, craved playful contact, were less aggressive, and were
probably less threatening than same-aged peers.
Therapy consisted of three 2-hour sessions per week for four weeks—a total of 24 hours of treatment.
When the therapists arrived for the session, the isolates were terrified and retreated. The therapists
approached, touched, and climbed on their older clients. The isolates tried to retreat while their anxiety and
self-stimulating behavior increased. Again the therapists engaged, touched, climbed on, and got in the face of
their clients. Apparently, when it comes to play and social engagement, a three-month-old monkey won’t take
no for an answer. As the sessions continued, it was reported that the isolates gradually came to habituate and
accept their “therapists.” These interactions interrupted autistic, self-stimulating behavior, and the clients
eventually began to initiate physical contact and interact with their younger therapists. Therapy was so
successful that the authors stated, “By 1 year of age, the isolates were scarcely distinguishable from the normal
therapists in terms of frequencies of exploratory, locomotive, and play behavior” (Harlow & Suomi, 1971, p.
1537).
After a course of treatment, when these former isolates were introduced to the colony, they did much
better, and were able to find a role within the group and social hierarchy. Were they impaired? Most likely,
early deprivation had a long-lasting impact, but it appeared to the researchers that they had attained a
functional social recovery. These results surprised Harlow and Suomi because of their prior assumptions about
critical periods. They also help remind me to keep an open mind, and remember that giving up on a child or
client is not something I am ever willing or prepared to do.

Summary

128
Neuroscience suggests that an important aspect of love is the absence of fear. If therapists and adoptive
parents can create an environment that minimizes fear and maximizes the positive neurochemistry of
attachment through human compassion, attachment circuitry can be stimulated to grow in ways which are not
only healing, but that also allow victims of abuse and neglect to risk forming a bond with another.
Because the process of attachment is, at heart, a way in which social animals initially regulate fear, and
later their affective lives, modifying insecure attachment, first and foremost, requires the establishment of a
safe and secure relationship. Therapists work diligently to establish this type of relationship for each client and
to create an experience similar to what the three-month-old monkeys were able to give their senior isolates:
the experience of social connection in the absence of threat or rejection.
There are probably thousands of studies supporting what we all intuitively know—childhood experience
affects emotional and physical health later in life. While plenty of psychological and social theories attempt to
explain this relationship, we are beginning to put together the biological mechanisms of action of these
findings. The general question is, how do early social experiences shape our neurobiology in ways that can
influence us decades later?

129
Chapter 13
The Neurobiology of Attachment
Offspring inherit, along with their parents’ genes, their parents, their peers, and the places they inhabit.
—Leon Eisenberg

Reflected in the architecture of each of our brains is the coming together of our evolutionary history, the
generations preceding our birth, and our unique relationship with our parents (Eisenberg, 1995). Hundreds of
studies have demonstrated the ways in which early experience is correlated with physical and emotional health
later in life. Research in psychoanalysis, epidemiology, developmental psychology, and psychiatry have all
supported what we think of as common sense: A good childhood is better than a bad one; positive parental
attention is important; and less stress early in life is a good thing. Of course, each field explains these findings
from its own theoretical model and tends to see other perspectives as secondary.
Research in molecular biology offers us a groundbreaking view into the underlying mechanism of the
effects of early experience on genetic expression, that is, how early experience triggers gene expression to guide
our brains onto particular adaptational trajectories. In contrast to the correlations found in other fields of
study, this new work explores causal biological links between maternal behavior and the building of the brains
of children.
When we think about having an internalized mother, our thoughts usually stray to images of a kind smile,
a warm hug, feeling safe, and being loved. Depending on the culture, you may remember your mom serving
Thanksgiving dinner, stirring tomato sauce, or frying chicken. Those of us who are less fortunate may have
images of enraged behavior, endless criticism, or a mother passed out on the couch after a long day of
drinking. These conscious autobiographical memories are but one layer of an internalized mother. Another
level, deeper and just as meaningful, is how these early experiences shaped the neurobiological processes of our
brains.
In this chapter we climb up and down the evolutionary ladder from humans to rats and back again to
humans. It turns out that we have learned a great deal by exploring how the behavior of mama rats influences
the brains of their pups. The conservation of structures and functions during evolution provides us with a
good animal model of the effects of maternal behavior on the brain. Although this research has yet to be done
with humans, the behavioral and neurobiological parallels between rats and humans are striking, making rats
very helpful in understanding the interpersonal aspects of neurobiology. We also explore networks of the
human brain that rely on similar shaping during early experience, as well as some other ways in which
epigenetic factors may impact everything from the timing of menopause to human longevity.

The Evolution of Complexity


Who takes the child by the hand takes the mother by the heart.
—German proverb

Our best guess is that larger and more complex brains allow for more diverse responses in challenging
situations across environments. Our brains allow us to fashion clothing, build houses with heating systems,
and create sophisticated farming techniques that allow us to expand our habitats and sources of food. But does
this explain why we have relationships?
We know that the expansion of the cortex in primates correlates with increasingly large social groups. The
benefit of living in a tribe is not just in the safety of numbers, but also in the ability of groups to have task
specialization such as hunting, gathering, and caretaking. So, while many fish and reptiles need to be born
immediately prepared to take on the challenges of survival, human infants have years of total dependency
during which their brains can grow and adapt to very specific environments. This larger window of time for

130
adult-child interactions allows for increasingly sophisticated postnatal brain development and increased
investment in each child (Kaplan & Robson, 2002). This expanded social investment enhances the child’s
chances of survival, which in turn increases the chances of our genes surviving into the next generation (Allen,
Bruss, & Damasio, 2005; Charnov & Berrigan, 1993). Thus, the development of the brain, group
organization, caretaking, and social communication coevolved in a mutually interdependent manner.
There are many interesting theories about how humans evolved to be the complex social creatures we are.
The big story probably goes something like this:

• Larger group size enhanced the probability of survival, but required bigger and more complex brains to
process social information, so larger brains were selected.
• More complex brains require longer periods of development and result in prolonged periods of child
dependency.
• Longer periods of dependency require more attention and dedication to nurturance, caretaking
specialization, and social structures that can support this specialization.
• As the size of primate groups expanded, grooming, grunts, and hand gestures became inadequate and
were gradually shaped into spoken language.
• Complex social structures encouraged the development of more sophisticated communication, leading
to the development of oral and written language.
• As social groups grew in size and language became more complex, a larger cortex was needed to process
increasingly complex information.
• Language and culture provide our expanding brains the ability to record and accumulate history and
information, and develop technology.
• The accomplishments of culture allow for even greater group size and more sophisticated brains.

It is very likely that some version of this evolutionary narrative shaped the contemporary human brain. But
despite some remarkable advances, we have continued to be governed by the basic biological principles of
homeostasis, fundamental approach-avoidance choices, and flows of electrical and chemical information
throughout our brains and bodies. Like every living system, from a single neuron to complex ecosystems, the
brain depends on interaction with other brains for its survival. Because increasing complexity requires greater
interdependency, our brains have come to exist more and more profoundly within a matrix of other brains.
At birth, humans are dependent on caretakers for survival and growth. The prolonged and sophisticated
parenting that has evolved in primates scaffolds an increasing amount of postnatal growth and development.
This allows each human brain to be a unique blending of nature and nurture as it builds its structures through
interactions and molding itself to its environment. Parents’ nonverbal communications and patterns of
responding to the infant’s basic needs also shape the baby’s perceptions of the world and its sense of self.
Because the first few years of life are a period of exuberant brain development, early experience has a
disproportionate impact on the development of neural systems.
Genes first serve to organize the brain and trigger sensitive periods, while experience orchestrates genetic
transcription in the ongoing adaptive shaping of neural systems, so that experience becomes the actual
hardware of our brains. This structure, in turn, organizes other brains, allowing experiences to be passed
through a group and carried forward across generations. While being embedded in a group comes with many
challenges, it also comes with an ability to interactively regulate each other’s internal states and assist in neural
integration.

Environmental Programming
The group consisting of mother, father, and child is the main educational agency of mankind.
—Martin Luther King Jr.

The transduction of bonding experiences into neurobiological structure is a fascinating area of study. It carries
deep implications for how relationships throughout life impact our experience and thereby shape our brains.
Michael Meaney and his colleagues have been studying this question in great depth for many years. Their
work has taken advantage of naturally occurring variation in the maternal behavior of mother rats (dams) to
explore the impact of their ministrations on the brains of their pups. Mother rats lick, nurse, and retrieve their
pups when they roll out of the nest. These three behaviors are easily observed and counted by willing
undergraduates, and correlated with behavioral and biological variables in the brains of both mothers and

131
children.
The work of Meaney and others has provided us with ample evidence that mother rats pass on their genes
through DNA and shape genetic expression through their behavior. Environmental programming is a term
used to describe the expression of epigenetic factors during development (Fish et al., 2004; Meaney & Szyf,
2005; Sapolsky, 2004). Thus, two mechanisms of inheritance exist: slow changes across many generations
through mutation and natural selection, and rapid changes in genetic expression during each generation
(Clovis et al., 2005; Cameron et al., 2005; Meaney & Szyf, 2005; Zhang, Parent, Weaver, & Meaney, 2004).
Their research has thus far revealed three primary ways in which maternal behavior impacts variations in brain
structure—learning and plasticity, the ability to cope with stress, and later maternal behavior in adulthood. A
mother’s impact on the way her daughter will mother her children serves as a parallel channel of inheritance
that is highly sensitive to environmental conditions.
Genetic expression is programmed by experience through the alteration of the chromatin structure and the
methylation of DNA (Szyf, Weaver, & Meaney, 2007). In effect, the genome is like a keyboard while these
processes select the notes to be played. Methylation is a process by which a methyl group is added to DNA.
This has been shown to be a reversible but stable modification to DNA that is passed along to daughter cells
and can lead to long-term gene silencing. Low licking/grooming mothering results in increased glucocorticoid
receptor (GR) methylation, decreased GR expression, and an increased stress response. Licking/grooming
decreases methylation, increases GR expression, and downregulates the stress response (Weaver et al., 2007).
So as we show affection and kindness to our children, we may be building more resilient brains, an expression
of genetic variation that would likely have made Lamarck smile.
Three different research methods have been employed to study the effects of maternal behavior on genetic
expression. In the first model, the amount of attention is measured and the behaviors and brains of pups in
high- and low-attention groups are compared. The second examines the effects of periods of maternal
deprivation, while the third uses handling of the pups by human researchers as the experimental manipulation.
Because it has been found that human handling stimulates more maternal attention, the first and third
categories may turn out to be one and the same (Garoflos et al., 2008).
Levels of maternal attention have been shown to either stimulate or silence gene expression in the domains
of neural growth and plasticity, modulation of hypothalamic-pituitary-adrenal (HPA) activity, and
programming of future maternal behavior (Szyf, McGowan, & Meaney, 2008). Neural growth is stimulated
via the activation of brain-derived neurotrophic factor (BDNF) and messenger RNA expression in a variety of
brain areas, processes tied to the biochemistry of neuroplasticity and learning. Stress reactivity is controlled via
levels of benzodiazepine, oxytocin, and GRs in many regions of the brain. Higher levels of maternal attention
result in more of these receptors being formed, allowing for the dampening of fear and anxiety and an increase
in exploratory behavior. Maternal behavior is governed by the growth and activation of medial optic areas (the
rats’ version of our ompfc) as well as the regulation of oxytocin and estrogen receptors (Neumann, 2008). See
Table 13.1 for some of the specific findings in each area of study.
In essence, rats who receive more maternal attention have brains that are more robust, resilient, and
nurturing of others. They are able to learn faster and maintain memories longer. They are less reactive to
stress, making them able to use their abilities to learn at higher levels of arousal and across more difficult
situations. They will also suffer less from the damaging effects of cortisol by downregulating it sooner after a
stress response.
And finally, females growing up with more attentive mothers pass these positive features on to their
children. The mechanisms for the association in humans between early secure attachment and healthier minds
and bodies is likely similar but far more complex.
Maternal attention stimulates the expression of BDNF, the most abundant neurotrophin in the brain.
Among its many functions, BDNF modulates glutamate-sensitive NMDA receptors which, in turn, regulate
long-term potentiation, long-term depression, and neuroplasticity (Alonso et al., 2002; Bekinschtein et al.,
2008; Monfils, Cowansage, & LeDoux, 2007). While cortisol inhibits the production of BDNF (and new
learning), higher levels of BDNF appear to both buffer the hippocampus from stress and encourage ongoing
plasticity (Pencea, Bingaman, Wiegland, & Luskin, 2001; Radecki, Brown, Martinez, & Taylor, 2005;
Schaaf, de Kloet, & Vreugdenhil, 2000). And because the production of BDNF (and other neurotrophins) is
under epigenetic control, physical, emotional, and interpersonal experience all influence its production and
availability (Berton et al., 2006; Branchi, Francia, & Alleva, 2004; Branchi et al., 2006).

TABLE 13.1
The Impact of Maternal Attention

Maternal Action Study Findings

132
Neural Growth and Plasticity
Licking Increased synaptic density, longer dendritic branching, and increased neuronal survival1
Licking Increased neuronal survival in the hippocampus2
Licking Fos expression in the hippocampus and parietal and occipital cortex3
Licking/nursing Increased NMDA and BDNF expression and increased cholinergic innervation of the
hippocampus4

Modulation of HPA Activity


Licking Increased medial prefrontal cortex dopamine in response to stress and increased startle inhibition5
Licking/nursing Decreased fear reactivity6
Licking/nursing Increased epigenetic expression of GR gene promoter in the hippocampus7

Licking/nursing Increased mRNA expression in medial prefrontal cortex, hippocampus, and the basolateral and
central regions of the amygdala8
Licking/nursing Increased levels of benzodiazepine receptors in the lateral, central, and basolateral regions of the
amygdala and the locus coeruleus as well as increased levels of alpha-2 adrenoreceptor density and
decreased corticotropin-releasing hormone receptor density in the locus coeruleus9

Modulation of Future Maternal Behavior


Nursing call Enhanced metabolic activation in precentral medial cortex, anterior cingulate cortex, and lateral
thalamus10
Licking Elevated levels of estrogen mRNA and more maternal behavior later in life11

Licking/nursing Increased levels of oxytocin and estrogen receptors in medial preoptic areas (and increased maternal
behavior when they have their own pups)12

Licking Less sexual behavior in females and less likely to become pregnant after giving birth13

Many researchers have found correlations between hippocampal volume and symptoms of depression.
While most stressful illnesses correlate with reductions in hippocampal volume, there is speculation that
depression may be a result rather than a cause of hippocampal reduction. In other words, the symptoms of
depression are an experiential expression of a shutdown of neuroplasticity. Therefore, if our neurons become
depressed, so do we. Since depression is often a natural consequence of prolonged stress, one mechanism of
action linking the two may be the catabolic impact of high levels of cortisol on the neurons within the
hippocampus. It is suspected that antidepressant SSRIs and physical activity work to reverse the negative
impact of cortisol in the hippocampus by triggering BDNF synthesis (Fernandes et al., 2008; Russo-Neustadt,
Beard, Huang, and Cotman, 2000; Warner-Schmidt & Duman, 2006). Direct administration of BDNF has
also been shown to have long-lasting antidepressant effects (Hoshaw, Malberg, & Lucki, 2005).
While more maternal attention results in increased growth and enhanced functioning throughout the pups’
brains, separation from mothers proves to have the opposite effects. The same three areas that are upregulated
with more maternal attention are all downregulated by her absence. Deprivation of maternal attention
increases neural and glial death, while reducing gene expression, impairing pups’ ability to learn. Maternal
separation also results in reduced inhibitory gamma-aminobutyric acid (GABA) receptors in the locus
coeruleus, increasing adrenaline secretion in reaction to stress while reducing the antianxiety properties of
benzodiazepine receptors in the amygdala. Decreased cortisol receptors in the hippocampus also impair the
inhibitory feedback to the stress system to shut down cortisol production. See Table 13.2 for the specific
studies from which this information is taken. So again we see results that parallel findings with human
subjects where early maternal deprivation through separation or depression results in decreased brain
functioning, higher levels of anxiety, and difficulty with subsequent attachment (Brennan et al., 2008; Tyrka
et al, 2008).

TABLE 13.2
The Impact of Maternal Separation

Neural Growth and Plasticity


Increased neuronal and glial death1

Decreased neurotrophin levels in ventral hippocampus2


Decreased glial density3

133
Modulation of HPA Activity
Reduced GABA receptors in the locus coeruleus
Decreased GABA receptor maturity
Reduced benzodiazepine receptors in the central and lateral amygdala and increased mRNA expression in the amygdala4
Increased anxiety, fearfulness, and response to stress5
Increased LTP and LTD in amygdalo-hippocampal synapses6
Decreased exploratory behavior, avoidance of novelty, and greater vulnerability to addiction7
Reduced gene expression8

Greater cortisol secretion in response to mild stress and increased startle response and startle-induced sounds9
Reduced somatic analgesia and increased colonic motility, mimicking irritable bowel syndrome in humans10
Upregulation of glutamate receptors11

Modulation of Future Maternal Behavior


Decreased synaptic density in the medial prefrontal cortex
Decreased cell survival in maternal neural networks12
Decreased activation in the bed nucleus of the stria terminalis and nucleus accumbens13

The evidence from the handling studies is essentially the same as for highly attentive mothers in neural
health and the modulation of anxiety, supporting the idea that they may both represent the effects of greater
amounts of maternal attention. More GRs, lower cortisol levels, and greater brain activity reflect a brain
geared toward less anxiety, helplessness, and fear. In turn, these pups are more resilient, engage in more
complex exploratory behavior, and are better learners than their nonhandled siblings. Similar results have also
been discovered in parrots and pigs (see Table 13.3).

TABLE 13.3
The Impact of Human Handling on Rats, Pigs, and Parrots

Modulation of HPA Activity

Rat Pups
Increased concentrations of GRs in the hippocampus and frontal lobes1
Increased GR binding capacity in the hippocampus2

Increased corticotrophin-releasing factor mRNA3


Decreased inhibitory avoidance and increased object recognition4

Lower levels of stress in reaction to a predatory odor5

Increased neurotrophin-3 expression and neuronal activation in hippocampus and parietal lobes6
Low cortisol secretion in response to stress and high exploratory behavior7

Protection against neuroendocrine and behavioral decline with age8

Decreased helplessness behaviors9

Baby Pigs
Lower basal and free plasma levels of cortisol10

Amazon Parrots
Decreased serum cortisol levels in response to stress11

These and other studies support the idea that the reaction of the brain to maternal attention is not an
abstract theory, but a well-documented phenomenon with a mechanism of action that is slowly being
uncovered. The consistency of behavioral, emotional, and biological findings across species is too powerful to
be discounted. In fact, over 900 genes have been discovered that are differentially expressed based on the
amount of maternal behavior (Rampon et al., 2000; Weaver, Meaney, Szyf, 2006). And there is no reason to
believe that the maternal control of epigenetic expression has not been conserved in primates and humans.
Rhesus monkeys deprived of maternal contact demonstrate reduced transcriptional efficiency of serotonin

134
and its receptors in the brain (Bennett et al., 2002). We do know that low levels of caring maternal behavior
in humans correlate with more fearful behavior, less positive joint attention, and right-biased frontal
activation, all of which are related to higher levels of stress and arousal (Hane & Fox, 2006). Self-esteem and
locus of control have been found to correlate with hippocampal volume, which we know is tied to cortisol
regulation (Pruessner et al., 2005). In my mind, the parallels as well as the tendency for evolution to conserve
such mechanisms form a strong case for the theory that what Meaney and his colleagues are finding in rats is
at work in humans.
In an exciting twist, it has been found that biological interventions and enriched social and physical
environments can reverse the effects of low levels of maternal attention and early deprivation on both HPA
activity and behavior (Bredy et al., 2004; Francis, Diorio, Plotsky, & Meaney, 2002; Hood, Dreschel, &
Granger, 2003; Szyf et al., 2005; Weaver et al., 2005). Unfortunately, chronic stress or trauma in adolescence
and adulthood can also reverse the positive effects of higher levels of attention earlier in life, shaping a brain
that resembles one that was deprived of early maternal attention (Ladd, Thrivikraman, Hout, & Plotsky,
2005). These studies all support the notion that our brains are capable of continual adaptation in both positive
and negative directions and that successful psychotherapy, one that establishes a nurturing relationship, may
well be capable of triggering genetic expression in ways that can decrease stress, improve learning, and
establish a bridge to new and healthier relationships.
Keep in mind that the amount of attention that a mother rat shows her pups exists in a broad adaptational
context. Highly stressed mothers demonstrate lower rates of licking and grooming, which prepare the pups’
brains for living in a stressful environment. In other words, under adverse conditions, maternal behavior
decreases, which programs the offspring for enhanced reactivity to stress. This likely increases the probability
of survival while simultaneously elevating the risk of physical and emotional pathology later in life (Diorio &
Meaney, 2007). The impact of neonatal handling is also different for male and female pups, reflecting their
divergent adaptational roles and contributions to the survival of their species (Park, Hoang, Belluzzi, & Leslie,
2003; Ploj, Roman, Bergstrom, & Nylander, 2001; Stamatakis et al., 2008). All of this suggests that the level
of maternal behavior is interwoven into a matrix of adaptation choices that vary based on external factors. The
fact that processes that are set in motion early in life can be modified by subsequent experience demonstrates
the ability to adapt to a changing environment.
This work with rats has established guidelines for future exploration into environmental programming in
humans. There are obvious limitations to research with humans that requires physical examination of the
brain. We will have to rely on samples of opportunity and utilize careful methodological controls to be certain
of the quality of results. One such study compared the brains of suicide victims with normal controls and
found lower mRNA levels of BDNF and trkB in the suicide victims, both of which are involved in neuronal
health and plasticity. These data raise the possibility that early environmental programming may have made
them susceptible to depression and suicide (Dwivedi et al., 2003). A more recent study compared the brains of
suicide victims with and without histories of child abuse and found that those with histories of early abuse
demonstrated decreased levels of GR mRNA, receptor expression, and growth factor transcription when
compared to those without abuse (McGowan et al., 2009). These studies are highly supportive of our ability
to apply animal research to humans.

Attachment and the Human Brain


What I do and what I dream include thee, as the wine must taste of its own grapes.
—Elizabeth Barrett Browning

It turns out that gene expression in response to close contact is a two-way street. Giving birth and exposure to
children changes the brains of parents and caretakers in ways that support bonding, attachment, and
nurturance. A primary discovery in mother rats has been the remodeling and expansion of the hippocampus in
preparation for locating, storing, and retrieving greater amounts of food (Pawluski & Galea, 2006). Contact
with her pups causes increased growth in the medial preoptic area, basolateral amygdala, and parietal and
prefrontal cortex as the brain expands to incorporate these new beings into the self-experience of the mother
(Fleming & Korsmit, 1996; Kinsley et al., 2006; Lonstein, Simmons, Swann, & Stern, 1998). Even virgin rats
who are given pups to care for experience increased dendritic growth and neuronal excitation in superoptic
areas (Modney, Yang, & Hatton, 1990; Modney & Hatton, 1994; Salm, Modney, & Hatton, 1988). Thus,
just as with children, interpersonal contact changes the brains of parents.
It has been shown that when human mothers hear their infants cry, activity increases in their right medial

135
prefrontal cortex and anterior cingulate—regions known to mediate maternal response (Lorberbaum et al.,
1999). Watching a video of their own infant will stimulate activity in the right anterior temporal pole, left
amygdala, and both right and left ompfc (Minagawa-Kawai et al., 2008; Nitschke et al., 2004; Ranote et al.,
2004).
It is likely that just as maternal attention triggers epigenetic factors in children, caring for children may
also change genetic expression in their caretakers. The grandmother gene hypothesis suggests that human
women experience early menopause to be available to help nurture their grandchildren and avoid the risks
inherent in mating and childbirth, which rise with age for both mother and child (Lee, 2003; Rogers, 1993;
Turke, 1997). In essence, the grandmother gene hypothesis suggests that early menopause has been shaped by
natural selection to enhance the survival rate of a woman’s children’s children.
Sear, Mace, and McGregor (2000) studied tribal life in rural Gambia where families live at the level of
subsistence and depend upon one another for basic survival. Their lifestyle is likely similar to the social and
environmental context of most of our evolutionary history. The results of their research show that nutritional
status, height, and the probability of survival of their young were significantly correlated with the presence of
postmenopausal maternal grandmothers. On the other hand, the presence of fathers, paternal grandmothers,
or other male kin had a negligible impact on either the nutritional status or survival rates of offspring. Similar
findings have been obtained in studies of hunter-gatherer populations of Tanzania, premodern Japan, Canada,
and Finland, and contemporary urban populations in the United States (Hawkes, O’Connell, & Jones, 1997;
Lahdenperä et al., 2004; Pope et al., 1993).
Besides the timing of menopause, is it possible that women live longer because they have been traditionally
more involved in caring for their young? That is, is there a health benefit involved in taking care of children
because of the neurobiological processes it stimulates in our brains and bodies? Some interesting evidence that
may support a connection between child care and longevity comes from looking at parenting responsibilities
across different species of primates. A longevity advantage for females does exist in gorillas, orangutans, and
humans, in which they are the primary caretakers. On the other hand, in species such as the owl and titi
monkeys, where the male is the primary caretaker of infants, males tend to survive longer. In Goeldi monkeys,
where caretaking is shared, the longevity for both genders is equivalent (see Table 13.4).
Given the amount of data supporting the beneficial effects of secure attachment, caretaking, human touch,
and social support, it is plausible that nurturing, emotional attunement, and physical contact have salubrious
effects that may provide primary caretakers with a survival advantage. It is possible that attachment bonds,
caretaking experiences, and the neurochemicals and epigenetic phenomena they impact may well enhance our
health and survival. Perhaps caring for our children and grandchildren may be more supportive of health and
longevity than cholesterol medication and treadmills.

TABLE 13.4
Child Care and Longevity Across Primate Species
Primate Female/Male Survival Ratio Male Care
Chimpanzee 1.418 Rare
Spider monkey 1.272 Rare
Orangutan 1.203 None
Gorilla 1.199 Pair living, little direct role
Gibbon 1.125 Protects and plays with offspring
Human 1.052–1.082 Supports economically, some carecare
Goeldi monkey 0.974 Both parents carry infant
Siamang 0.915 Carries infant in second year
Owl monkey 0.869 Carries infant from birth
Titi monkey 0.828 Carries infant from birth

Adapted from Allman, Rosin, Kumar, & Hasenstaub, 1998.

Interestingly, women who give birth after the age of 40 are almost four times more likely to live to be 100
years old (Perls, Alpert, & Fretts, 1997). While this is usually explained in terms of the protective nature of
birth-related hormones, their enhanced longevity may be part of broader biological and psychological
processes involved in caretaking (King & Elder, 1997). It is a good bet that taking care of a child tells the
brain and body to trigger epigenetic and biochemical processes that enhance health and slow down aging.

The Human Social Brain

136
It is good to rub and polish our brains against those of others.
—Montaigne

We have seen a great deal of evidence of the impact of early nurturance on the shaping of the social brain and
its emotional circuitry. So when our early relationships are frightening, abusive, or nonexistent, our brains
dutifully adapt to the realities of our unfortunate situations. However, there is reason to believe that these
circuits retain experience-dependent plasticity throughout life, especially in close relationships (Bowlby, 1988;
Davidson, 2000). Experience-dependent plasticity has been found in many areas of the brain, including the
prefrontal cortex and hippocampus (Kolb & Gibb, 2002; Maletic-Savatic, Malinow, & Svoboda, 1999).
These structures, central to learning and memory, are also key in shaping attachment schema. Further,
research is emerging that in the transition from dating to marriage, there is a broad tendency to move from
insecure and disorganized attachment schema to increasingly secure patterns (Crowell, Treboux, & Waters,
2002). On the other hand, social stress inhibits cell proliferation and neural plasticity, while social support,
compassion, and kindness support positive neural growth (Czéh et al., 2007; Davidson, Jackson, & Kalin,
2000).
While rats possess the basic mechanisms of bonding and maternal behavior, our brains have more
elaborate and sophisticated mechanisms for attachment. In fact, the human brain is crisscrossed with neural
networks dedicated to receiving, processing, and communicating messages across the social synapse. The
difference in humans is that the environmental programming of these experience-dependent circuits is far
more lengthy and complex. Networks of our complex social brains include brain regions, neural systems, and
regulatory networks listed in Table 13.5.
These are the same circuits that therapists attempt to influence in reshaping the brain in ways which lead
to more positive adaptation later in life. The idea that psychotherapy is a kind of reparenting may be more
than a metaphor; it may be precisely what we are attempting to accomplish at the level of the epigenome. This
research establishes attention, care, and nurturance as a way to influence the very structure of our brains and
places psychotherapy at the heart of biological interventions. It is odd to think that Carl Rogers may someday
find a place next to Crick and Watson in the pantheon of biologists.
Now, let’s shift to the structures of the human brain that organize attachment, affect regulation, and the
modulation of stress. For a more in-depth exploration of the social brain, see The Neuroscience of Human
Relationships, second edition (Cozolino, 2012). Keep in mind that, just as in rats, these systems are also built
by the attachments they come to control. Thus, our learning history comes to be reflected in the architecture
of our neural systems.

TABLE 13.5
Structures and Systems of the Social Brain

Cortical and Subcortical Structures


Orbital and medial prefrontal cortices (ompfc)
Cingulate cortex and spindle (Von Economo) cells
Insula cortex
Somatosensory cortex
Amygdala, hippocampus
Hypothalamus

Sensory, Motor, and Affective Systems


Face recognition and expression reading
Imitation, mirroring, and resonance systems

Regulatory Systems
Attachment, stress, and fear regulation (orbital medial prefrontal cortex–amygdala balance)
Social engagement (the vagal system of autonomic regulation)
Social motivation (reward representation and reinforcement)

Cortical and Subcortical Structures


The prehistorical and primitive period represents the true infancy of the mind.
—James Baldwin

The ompfc, insula, and cingulate cortices—the most evolutionarily primitive areas of the cortex—lie buried

137
beneath and within the folds of the later-evolving cortex. In fact, some neuroanatomists see these contiguous
structures as creating a functional system called the basal forebrain (Critchley, 2005; Heimer & Van Hoesen,
2006). The ompfc sits at the apex of the limbic system. As a convergence zone for polysensory, somatic, and
emotional information, it is in the perfect position to synthesize information from both our internal and
external worlds. In its position at the apex of the limbic system, the ompfc’s inhibitory role in autonomic
functioning highlights its contribution to the organization of behavior and affect regulation.
The ompfc allows us to translate the punishment and reward values of complex social information, such as
facial expressions, gestures, and eye contact, into meaningful information, associate it with our own emotions,
and thus organize attachment schema (O’Doherty et al., 2001; Tremblay & Schultz, 1999; Wang et al., 2014;
Zald & Kim, 2001). The ompfc also mediates emotional responses and coordinates the activation and balance
of the sympathetic and parasympathetic branches of the autonomic nervous system (Hariri et al., 2000; Price,
Carmichael & Drevets, 1996).
The cingulate cortex is a primitive association area of visceral, motor, tactile, autonomic, and emotional
information that begins to participate in brain activity during the second month of life (Kennard, 1955). It
first appeared during evolution in animals exhibiting maternal behavior, play, and nursing, and when making
sounds became involved with communication between predator and prey, potential mates, and mother and
child (MacLean, 1985). The caretaking and resonance behaviors made possible by the cingulate also provide
an important component of the neural infrastructure for social cooperation and empathy (Rilling et al., 2002;
Vogt, 2005). Destruction of the anterior cingulate in mammals results in mutism, a loss of maternal responses,
infant death due to neglect, and emotional and autonomic instability (Bush, Luu, & Posner, 2000; Bush et al.,
2002; Paus, Petrides, Evans, & Meyer, 1993).
The anterior cingulate contains spindle-shaped neurons that appear to have evolved in humans and great
apes to connect and regulate divergent streams of information (Nimchinsky et al., 1995, 1999). These cells
may provide the neural connectivity necessary for both the development of self-control and the ability to
engage in sustained attention to difficult problems (Allman et al., 2001, 2005). Spindle cells are especially
fascinating because they emerge after birth and are experience dependent. Early neglect, stress, and trauma
may negatively impact the development and organization of the anterior cingulate and spindle cells, resulting
in lifelong cognitive deficits, and emotional functioning may be based on the construction and health of these
structures (Cohen et al., 2006; Ovtscharoff, Helmeke, & Braun, 2006).
The insula begins life on the lateral surfaces of the brain, only to become hidden by the rapid expansion of
the frontal and temporal lobes (Uddin, 2014). The insula is sometimes described as the limbic integration
cortex because of its massive connections to all limbic structures, and its feed-forward links with the frontal,
parietal, and temporal lobes (Augustine, 1996). It provides the brain with a means to connect primitive bodily
states with the experience and expression of bodily awareness, emotion, and behavior (Carr et al., 2003; Phan,
Wager, Taylor, & Liberzon, 2002; Wiech et al., 2010).
In tandem with the anterior cingulate, the insula allows us to be aware of what is happening inside of our
bodies and reflect on our emotional experiences (Bechara & Naqvi, 2004; Critchley et al., 2004; Gundel,
Lopez-Sala, & Ceballos-Baumann, 2004). Damage to the right insula can result in anosognosia, a condition
where a patient seems unaware of and unfazed by severe paralysis on the left side of the body (Garavan, Ross,
& Stein, 1999). More recent research suggests that the insula is involved with mediating the entire range of
emotions from disgust to love (Bartels & Zeki, 2000; Calder et al., 2003; Menon & Uddin, 2010).
The somatosensory cortex, located along the front of the parietal lobes, processes information about bodily
experiences. It lies just behind the central gyrus and wraps deep within the Sylvian fissure that divides the
parietal from the frontal lobes. Along with the insula and anterior cingulate cortices, it contains multiple
representations of the body that process and organize our experience of touch, temperature, pain, joint
position, and our visceral state. These different processing streams combine to create our experience of our
physical selves. It also participates in what we call intuition or gut feelings by activating implicit memories
related to our experiences and helping us to make decisions guided by feelings (Damasio, 1994). The
experience of our own bodies becomes the model for our understanding and empathizing with others
(Damasio et al., 2000).
Working in concert with the ompfc, the subcortical amygdala is another core component of the social
brain (Bickart et al., 2010). The amygdala achieves a high degree of maturity by the eighth month of
gestation, allowing it to associate a fear response to a stimulus prior to birth (LaBar, LeDoux, Spencer, &
Phelps, 1995; Ulfig, Setzer, & Bohl, 2003). As a primitive organ of appraisal, the amygdala closely monitors
signals of safety and danger and mediates the fight-or-flight response via the autonomic nervous system
(Davis, 1997; Ono, Nishijo, & Uwano, 1995; Phelps & Anderson, 1997; Rudebeck, Mitz, Chacka, &
Murray, 2013). The ompfc can inhibit the amygdala based on conscious awareness and feedback from the

138
environment (Beer et al., 2003). By the same token, when we are frightened and our amygdala is activated, it
inhibits the ompfc, and we have more difficulty being rational, logical, and in control of our thoughts. The
amygdala also appears to contribute to our conscious experience of both emotional and physical pain (Mitra &
Sapolsky, 2008; Neugebauer, Li, Bird, & Han, 2004). Since the networks connecting the ompfc and the
amygdala are shaped by experience, our learning history of what is safe and dangerous, including our
attachment schema, is thought to be encoded within this system.
The hippocampus is situated at the junction between the cortex and limbic system on both sides of the
brain. In lower mammals like the rat, the hippocampus is a specialized spatial map of foraging territory. In
humans, the parietal lobes evolved from the hippocampus and assist it in complex visual-spatial processing.
The human hippocampus, along with its adjacent structures (parahippocampal gyrus, dentate gyrus), have
come to be specialized in the organization of spatial, sequential, and emotional learning and memory
(Edelman, 1989; McGaugh et al., 1993; Sherry et al., 1992; Zola-Morgan & Squire, 1990). In contrast to the
amygdala, the hippocampus is a late bloomer, continuing to mature into early adulthood along with the dlpfc
connections upon which it relies (Benes, 1989). Our lack of conscious memory for early childhood, known as
childhood amnesia, is likely due to the slow developmental course of the hippocampus (Fuster, 1996; Jacobs et
al., 2000; McCarthy, 1995).
The hypothalamus is a small and ancient structure that sits at the center of the brain below the thalamus
and halfway between the cortex and the brain stem. It has extensive connections with the structures of the
social brain within the frontal lobes, limbic system, and brain stem. I include the hypothalamus as part of the
social brain because it is centrally involved with the translation of conscious experience into bodily processes
and thus, the transduction of early experience into the building of the brain and body. Its various nuclei
organize many bodily functions such as temperature regulation, hunger, thirst, and activity level. The
hypothalamus is also involved in the regulation of sexual behavior and aggression. As the head of the HPA
axis, it translates brain processes into hormonal secretions from the anterior pituitary. Among the hormones
produced by the pituitary, follicle-stimulating hormone and prolactin are involved in reproduction and
nursing. Adrenocorticotropic hormone is sent to the adrenal glands via the bloodstream, and stimulates the
production of cortisol, which we discuss in depth later as it relates to caretaking and early stress.

Sensory, Motor, and Affective Systems


Common Sense is that which judges the things given to it by other senses.
—Leonardo da Vinci

It is in the temporal lobes that our senses are integrated, organized, and combined with primitive drives and
emotional significance in a “vertical” linkup across all three levels of the triune brain (Adams, Victor, &
Ropper, 1997). For example, the recognition of faces and reading their expressions occurs in the top-down
networks. Cells involved in both reading and identifying facial expressions are located in adjacent areas of the
temporal lobes (Desimone, 1991; Hasselmo, Rolls, & Baylis, 1989). When we see faces, the areas of the brain
that become activated lie in a processing stream dedicated to the identification of visual stimuli (Lu et al.,
1991). The association region of the occipital lobe dedicated to the identification of faces is the fusiform face
area (Gauthier et al., 2000; Halgren et al., 1999). These areas, in turn, are interconnected with other clusters
of cells that are responsible for eye gaze, body posture, and facial expression as the brain constructs complex
perceptions and social judgments from basic building blocks of visual information (Jellema, Baker, Wicker, &
Perrett, 2000).
Regions in the anterior (front) portions of the superior temporal sulcus integrate information about various
aspects of the same person (form, location, and motion), allowing us to identify others from different angles,
in various places, and while they are in motion (Jellema, Maassen, & Perrett, 2004; Pelphrey et al., 2003;
Vaina et al., 2001). The superior temporal sulcus also contains mirror neurons, which activate either when we
witness others engaging in behaviors or when we ourselves subsequently engage in these actions. By bridging
neural networks dedicated to perception and movement, mirror neurons connect the observed and the
observer by linking visual and motor experience. Resonance behaviors, which are based on mirror systems, are
the reflexive imitation responses we make when interacting with others. It is hypothesized that mirror systems
and resonance behaviors provide us with a visceral-emotional experience of what the other is experiencing,
allowing us to know others from the inside out.

139
Regulatory Systems
We gain our ends only with the laws of nature; we control her only by understanding her laws.
—Jacob Bronowski

The body’s regulatory systems are involved in the maintenance of internal homeostatic processes, balancing
approach and avoidance, excitation and inhibition, and fight and flight responses. They also control
metabolism, arousal, and our immunological functioning. It is through these systems that we regulate each
other’s biological and emotional states.

The Stress, Fear, and Attachment System


I’m not afraid of storms, for I’m learning how to sail my ship.
—Louisa May Alcott

The HPA system regulates the secretion of hormones involved with the body’s response to stress and threat.
The immediate reaction to stress is vital for short-term survival, while the rapid return to normalization after
the threat has passed is essential for long-term survival. Prolonged stress results in system damage and
breakdown. The long-term effects of negative parenting experiences, failures of attachment, and early trauma
are mediated via the HPA system. In terms of fear, we turn again to the amygdala, as it alerts a variety of
brain centers that a fight-flight response is required. In turn, the activation of the sympathetic branch of the
autonomic nervous system results in symptoms of anxiety, agitation, and panic. The prime directive of the
amygdala is to protect us by pairing a stimulus with a fear response, and it works so fast that this pairing can
be created far ahead of our conscious perception of threat. Throughout our lives, but especially during
childhood, relationships with others regulate our stress and fear. A secure attachment indicates that we have
learned to successfully utilize our relationships with others to quell our fears and modulate our arousal.

The Social Engagement System


Communication leads to community, that is, to understanding, intimacy and mutual valuing.
—Rollo May

The tenth cranial nerve, also called the vagus, is actually a complex communication system between the brain
and multiple points within the body including the heart, lungs, throat, and digestive system. Its afferent
(sensory) and efferent (motor) fibers allow for rapid continuous feedback between brain and body to promote
homeostatic regulation and the optimal maintenance of physical health and emotional well-being (Porges,
Doussard-Roosevelt, & Maiti, 1994). The vagal system is a central component of the autonomic nervous
system. In the absence of external challenge, the vagus works to enhance digestion, growth, and social
communication. When a challenge does arise, a decrease in vagal activation facilitates sympathetic arousal,
high energy output, and the fight-flight response. Between rest and all-out activation, the vagus allows us to
maintain continued engagement by modulating arousal during emotional interpersonal exchanges. The vagal
system accomplishes this by modulating and fine-tuning sympathetic arousal.
Like the attachment system described earlier, the development of this engagement system and the fine-
tuning of the vagal brake to regulate affect appear to depend on the quality of attachment relationships in
early childhood. This allows us to internalize what we learn from experience with caretakers into moment-to-
moment somatic regulation. The vagal system also controls the primary facial, mouth, and throat muscles
involved in communication, and links them with an awareness and control of internal states, coordinating the
cognitive and emotional processing necessary for relationships.
The “tone” of the vagus refers to the vagal system’s dexterity in regulating the heart and other target organs
(Porges et al., 1996). Inadequate development of this experience-dependent vagal tone can impact all levels of
psychosocial and cognitive development (Porges et al., 1994). Children with poor vagal tone have difficulty
suppressing emotions in situations demanding their attention, making it difficult for them to engage with
their parents, sustain a shared focus with playmates, and maintain attention on important material in the
classroom (see Table 13.6 for more details).
Vagal regulation allows us to become upset, anxious, or angry with a loved one without withdrawing or

140
becoming physically aggressive. We can hypothesize that many who engage in domestic violence, child abuse,
and other forms of aggressive behavior may not have had the kinds of early attachment relationships required
to build an adequate vagal system. Thus, good parenting not only teaches appropriate responses in challenging
interpersonal situations but also builds vagal circuitry required to stay engaged.

TABLE 13.6
Correlates of Vagal Tone

Higher Vagal Tone Correlates With Lower Vagal Tone Correlates With
The ability to self-regulate irritability Irritability
Self-soothing capacity by three months of age Behavioral problems at three years of age
The range and control of emotional states Emotional dysregulation
More reliable autonomic responses Distractibility
Suppression of heart rate variability Hyperreactivity to environmental and visceral stimuli

Enhanced attentional capacity and the ability to take in information


Positive social engagement Withdrawal
Increased behavioral organization Impulsivity/acting out
Consistent caretaking/secure attachment Insecure attachment

The Social Motivation System


Interdependence is and ought to be as much the ideal of man as self-sufficiency. Man is a social being.
—Mohandas Gandhi

Nelson and Panksepp (1998) postulated the existence of a social motivation system modulated by oxytocin,
vasopressin, endogenous endorphins, and other neurochemicals related to reward, decreased physical pain, and
feelings of well-being. While conserved from more primitive approach-avoidance and pain regulation
circuitry, the social motivation system extends into the amygdala, anterior cingulate, and ompfc. These circuits
and neurochemicals are thought to regulate attachment, pair bonding, empathy, and altruistic behavior
(Decety & Lamm, 2006; Schneiderman et al., 2012; Seitz, Nickel, & Azari, 2006). In other words, as Fisher
(1998) suggested, the social motivation system can be divided into three categories: those involved in bonding
and attachment (regulated by peptides, vasopressin, and oxytocin), attraction (regulated by dopamine and
other catecholamines), and sex drive (regulated by androgens and estrogens). The production of these various
biochemical elements, as well as the creation of their receptors, is subject to the influences of experience early
in life.
In addition, the dopamine reward system of the subcortical area known as the ventral striatum is involved
with more complex analysis of reward and social motivation. The ventral striatum becomes activated with an
expectation of a social reward, such as when we anticipate being given candy or positive attention (Kampe,
Frith, Dolan, & Frith, 2001; Pagnoni, Zink, Montague, & Berns, 2002; Schultz, Apicella, Scarnati, &
Ljunberg, 1992). For example, once the cortex has determined that we find someone attractive, the ventral
striatum becomes activated when they look our way, giving the signal that the possibility for being rewarded
with a desirable outcome has increased (Elliott, Friston, & Dolan, 2000; Schultz, Dayan, & Montague, 1997;
Schultz, 1998). The activation of the ventral striatum translates the anticipation of reward into a physical
impulse to approach. In this way, those whom we find attractive actually exert what feels like a gravitational
pull on us.

Summary

Research has provided us with new ways of understanding how early experience builds the brain. Maternal
attention has been linked to the neurobiology of systems related to learning and memory, stress regulation,
and attachment behavior. Although the human brain is far more complex than those of the animals upon
whom this research has been conducted, findings in human research across a broad array of disciplines
demonstrate consistent results in the areas of learning, resilience, and attachment. The neural hubs and
regulatory networks described here are all built in an experience-dependent manner. Early relationships shape
the building of neural circuitry, which guides how we are able to learn, react to stress, and attach to others in

141
ways that parallel those seen in the animal research discussed earlier. As we learn more about the complexities
of the human brain, we will understand how relationships build the brain, and how love becomes flesh.

142
Chapter 14
Altruism and Psychotherapy:
Leveraging the Social Brain
in the Service of Change
The best way to find yourself is to lose yourself in the service of others.
—Mahatma Gandhi

Most of us have experienced the positive thoughts and warm feelings associated with giving and receiving acts
of kindness. Giving to others warms our hearts, connects us to others, and makes us feel better about
ourselves. We are a social species. Taking care of others has evolved because it is essential for both individual
and group survival. As evidence of altruism’s deep history, we observe these behaviors among family, friends,
and strangers, even in our primate cousins (de Waal, 2008; Warneken et al., 2007). The neurobiological
mechanisms of attachment have come to drive the altruistic behaviors that are now sequenced in our genes,
wired into our brains, and woven into our cultures.
For countless eons, humans lived in small bands of hunter-gatherers in relatively stable environments. In
the face of climate change and geological upheaval, early humans had to adapt to new challenges in order to
survive. Our brains evolved into increasingly social organs as natural selection favored those groups that were
more exploratory, adaptive, and cooperative. Groups with more “team players” were able to hunt, fight, and
breed with more success. If you played for the team, your team would be more likely to survive, as would your
children and your genetic legacy (Boehm, 2009; Bowles, Choi, & Hopfensitz, 2003; Wilson, 2012).
Over time, what began with our instincts to bond with and care for our children allowed us to attach to
mates, extended family, and members of our tribes. The brain’s expanding ability to identify with larger and
larger groups enabled tribes to come together to form kingdoms, city-states, and nations. To this day, groups
that are unable to grow larger and work together are easy prey for those who can, leading to their exploitation
and potential annihilation. Given our evolutionary history, it is easy to understand why we possess the dual
motivations of selfishness and selflessness. What we call morals, ethics, and the golden rule are the cultural
expressions of this delicate self-other negotiation. What we admire as altruism, self-sacrifice, and acts of
bravery reflect a communal survival strategy that was a necessary precursor to our present stage of evolution.
As evidence of its importance, research has demonstrated that altruistic behaviors correlate with greater life
satisfaction, longevity, and better physical health (Post, 2005). Altruistic individuals report being happier,
exhibit fewer mental disturbances, and have fewer negative thoughts (Brown, Consedine, & Magai, 2005;
Nakamarua & Iwasab, 2006; Post, 2005; Tankersley, Stowe, & Huettel, 2007). This means that altruistic
people will be around longer to contribute to their tribe and that they will be rewarded for their contributions
by having a more enjoyable life.
For social animals like us, attachment, emotional attunement, and connection have also come to serve the
role of healing psychological distress. This is ultimately why the therapeutic relationship is our strongest lever
of change in psychotherapy. Long before the advent of psychotherapy, the work of healing was accomplished
within relationships with family, friends, clergy, and other wise elders in the context of everyday life. As the
talking cure converted natural healing social interactions into a professional intervention, healing became
radically separated from the real world. The professionalization of interpersonal healing left concepts like
altruism behind.
Altruism, as a one-way relationship, is already at the heart of therapy. Research has consistently
demonstrated that much of the success of psychotherapy depends upon the therapeutic alliance—support,
generosity of spirit, respect, and positive regard. The traditional therapeutic relationship is a form of
professionalized altruism—a nonreciprocal relationship established in order to allow our clients to explore and
develop their inner worlds. While therapists can’t define what we do as altruistic because we are paid for our
services, all of us understand that our work goes beyond a straightforward fee-for-service relationship. The

143
caring, loving, and reparenting that go on in the consulting room require a personal investment; we become
vulnerable to our clients’ suffering, joy, and heartbreak.
Neurobiological research reveals that the ways in which the brain becomes activated in the face of altruistic
behavior may also underlie central aspects of positive change in psychotherapy. The interactive nature of
caring for others and self-growth raises two questions: 1) Do these processes reflect the underlying value of a
positive therapeutic relationship? 2) Could engaging in altruistic activities between sessions have a positive
impact on therapeutic change? Put in another way, can altruism serve as a tool to enhance the therapeutic
process? In this chapter, I explore what happens in the altruistic brain and some ideas about how and why
guided altruistic activity could serve to benefit therapy.

Altruism and the Brain


Individual versus group selection results in a mix of altruism and selfishness, of virtue and sin, among the members of a society.
—E. O. Wilson

There is a direct relationship between the use and development of neural systems. In other words, the brain
grows through stimulation, challenge, and use. This is why experienced meditators have larger brain regions
involved with self-awareness, why pianists have expanded cortical areas controlling their fingers, and why
those who engage in challenging careers have more complex neural networks (Stern, 2002). Might altruistic
behavior develop neural networks in a way that will support therapeutic success?
It is becoming clear that networks dedicated to the experience of self and other overlap in our brains. This
is why being more aware of ourselves helps us to be more mindful of others and why self-compassion leads to
greater compassion for others. For these reasons, altruistic behaviors could potentially drive the development
of brain regions dedicated to attachment, self-efficacy, perspective-taking, and compassion, all of which would
be extremely beneficial to positive change in psychotherapy. Support for this idea lays not only in brain
research, but also in folk wisdom, religious traditions, and the philosophies of many social organizations.
Humans are not the only species to have developed social brains; elephants, dolphins, and many other
species also possess social brains. Long before homo sapiens arrived on the scene, attachment and sustained
pair bonding blossomed from the underlying biology of the mother-child relationship. As primate and later
human brains became more sophisticated, these instincts were expanded in ways that allowed us to connect
with a larger number of others. Today, our social affiliations are driven by a wide array of neurobiological
processes from primitive neurohormones to sophisticated cortical networks. Together, these neurochemicals
and brain regions comprise what might be called an altruism circuit.
The evolution from maternal instinct to acts of extreme altruism, like giving a kidney to a total stranger,
reflects the radical socialization of the human brain over eons of evolution. Although I am using the term
altruism circuit as shorthand to discuss the neural drivers of altruistic behavior, there is no dedicated altruism
circuit. Each of the neurochemicals and neural networks that contribute to altruism evolved for multiple
purposes and come together to enhance our ability to connect with others and feel united as a family, tribe, or
species. Whether by miracle or natural selection, altruism is an amazing evolutionary achievement.

Neurochemicals: Oxytocin and Dopamine


Although there is a full orchestra of neurochemicals underlying our drive to bond and attach, I focus here on
two primary molecules, oxytocin and dopamine. When mothers and children come together, separate, and
reunite, their oxytocin levels rise, fall, and rise again. Separation anxiety in a child and the distress we
experience when we can’t find our child are both driven by precipitous falls in oxytocin. The amygdala, the
executive center of fear circuitry, is rich in oxytocin receptors and is inhibited by the presence of oxytocin. By
inhibiting the amygdala, oxytocin allows us to feel safer, which enhances neuroplasticity, new learning, and
adaptive change. As we evolved into more sophisticated beings, oxytocin became involved with our
experiences of pleasure, pain, anxiety, irritability, and aggression. As you might expect, feelings of well-being
and diminished fear go along with increased levels of oxytocin, the reward for staying close to and taking care
of the ones we love.
Many of us enter relationships to regulate negative emotions such as anxiety, loneliness, and depression.
This is because increasing oxytocin makes us feel safer and triggers a wide range of positive physical and
emotional experiences. In contrast, decreases in oxytocin make us feel abandoned and lost and put us at risk
for self-destructive and suicidal behavior. Being addicted to love really means that we become addicted to the

144
neurochemicals it activates. This is also why critical, hostile, and abusive relationships take such a negative toll
on our minds, bodies, and spirits.
People who report engaging in more altruistic behavior have higher levels of oxytocin, which suggests that
altruistic acts may be both stimulated and reinforced by oxytocin which may account for how good it feels to
give to others (Zak, Stanton, & Ahmadi, 2007). Higher levels of oxytocin also correlate with greater trust,
empathy, cooperation, and in-group bias (De Dreu et al., 2010; Luo et al., 2015; Zak et al., 2007). When we
feel lost, frightened, or depressed, oxytocin activation is likely to decrease negative feelings and make us seek
out others for comfort. Stimulating oxytocin activation through guided altruism could enhance both the
therapeutic relationship and the beneficial effects of psychotherapy.
Working in connection with oxytocin is dopamine, the brain’s primary reward neurochemical. If natural
selection has shaped us to do something to enhance survival, the behavior will likely be paired with dopamine.
Dopamine activation creates a positive association to activities, objects, and other people. Dopamine drives
our shopping patterns and eating behaviors and who we hang out with. Altruistic behavior feels so good at
least in part because the ventral tegmental area and mesolimbic pathway, which make and deliver dopamine to
the rest of the brain, become more active (Moll et al., 2006).
Oxytocin and dopamine have a very special and important relationship. When we see something that
stimulates our dopamine reward center, say a new car, buying the car makes us feel good. After a month, the
car is just another car because we’ve habituated to it. But if we see something that activates both dopamine
and oxytocin release, say a new baby, after a month he or she is still very special. This is because the
simultaneous activation of oxytocin and dopamine inhibits dopamine habituation. This is why fads come and
go and we grow tired of possessions, but we continue to look forward to seeing those we love.
While oxytocin and dopamine are modulating and reinforcing our connection with others, an array of
neural networks allows us to understand, attune to, and empathize with the internal states of those around us.
A variety of brain regions that become active during altruism are dedicated to social intelligence, emotional
regulation, and the experience of self. These neural networks make it possible to recognize significant others
and experience their feelings in our own bodies, establishing the groundwork for attunement and empathy.

The Right Amygdala


Although the amygdala is the executive center of our fear circuitry, it is much more than that. It is a structure
that appraises both good and bad things in our world, signaling the body to approach what is good and to
avoid what is bad. Like the cerebral hemispheres, the amygdala are on both sides of the brain and have
lateralized specialties. The left amygdala focuses more on the external world while the right specializes in
internal experience. The right amygdala, coupled with the functional expertise of the right cortex, is involved
with the internal emotional experiences of self and other. When we struggle with emotional pain, shame, and
loneliness, it is likely that the right amygdala is involved.
People referred to as extraordinary altruists, such as those who have donated their kidneys to strangers,
have been found to have significantly larger right amygdala, which become more active than the left in
response to fearful facial expressions (Marsh et al., 2014). This suggests that these extraordinary altruists may
have a deeper recognition of and a stronger emotional reaction to the pain of others. In contrast, psychopaths,
who lack empathy and altruism, have significantly reduced right amygdala volume. Because of its functional
specialties, the right amygdala is a strong candidate for membership in an altruism circuit.

The Orbitomedial Prefrontal Cortex


The orbitomedial prefrontal cortex (ompfc) is a key structure in the organization of social connection and
when our attention turns to the experience of self and empathizing with others (Frith & Frith, 1999; Gusnard
et al., 2001; Gusnard & Raichle, 2001; Lane et al., 1997; Reiman et al., 1997). Thus, it is not surprising that
we see considerable activation of the ompfc during altruistic behavior (Moll et al., 2006; Skuse & Gallagher,
2008). At the apex of attachment circuitry, the ompfc is able to regulate the amygdala and other subcortical
structures, allowing proximity to caretakers and loved ones to modulate our anxiety and fear. What we call
secure attachment is essentially the ability to experience stress in the absence of autonomic arousal (e.g., a
flight-flight response).
As you might guess, securely attached individuals have been found to be more helpful to others and engage
in more altruistic behaviors. They also express more compassionate feelings and provide others with more
caretaking (Post, 2005). The activation of ompfc circuitry through altruistic behavior will likely stimulate
bonding and attachment in situations where the altruist feels safe. This may be why taking care of animals

145
often serves as a valuable bridge to relationships with people for those with social anxiety or posttraumatic
stress disorder. This could be a step toward connection that might allow clients to experiment with more risky
relationships with people. The circuits connecting the ompfc and amygdala are a central neural target for
change in psychotherapy.

The Temporal-Parietal Junction


The area of the cortex where the temporal and parietal lobes meet, the temporal-parietal junction (TPJ), is
involved with the storage of social and emotional memory and experiencing the interface of self and other.
This may be the area that activates our personal memory when observing others that leads to projection and
transference. As with the other candidates for membership in an altruism circuit, the TPJ becomes more
active than other cortical regions when we make a decision to engage in an altruistic act and in response to
expressions of empathy and sympathy from others (Jackson, Meltzoff, & Decety, 2005). Altruistic individuals
have more gray matter in the TPJ, suggesting that they may be touched more deeply by others and experience
deeper self-other interactions (Morishima et al., 2012).
Research has demonstrated that this region is involved in the identification and processing of social cues
from others as well as in the exercise of self-agency (Decety & Jackson, 2004). It is likely that the TPJ
automatically connects the distress we see in others with memories of our own distress, allowing us to
reflexively use our personal history to have sympathy for others and take action to assist them. It would make
sense to assume that the TPJ is an important area of brain activation in both clients and therapists during
states of attunement, transference, countertransference, and projective identification, as both conscious and
unconscious memories become activated during sessions.

The Superior Temporal Zone


Adjacent to the TPJ is the superior temporal zone (STZ), another brain region that becomes active during
altruistic behavior (Lutz et al., 2008; Tankersley et al., 2007). These areas (the posterior superior temporal
cortex and superior temporal sulcus) are dedicated to different aspects of social connectedness. While the
superior cortex is responsible for receptive language, the superior sulcus helps us to understand facial gestures
and expressions (Perrett et al., 1984; Jellema et al., 2000; Jellema & Perrett, 2003). Together they support
social processes such as imitation, joint attention, modeling, and theory of mind. The STZ is particularly
active during face-to-face interactions in the psychotherapeutic relationship.

Mirror Neurons
Mirror neurons within the frontal and parietal lobes fire both when we engage in a behavior and when we
watch another engage in the same behavior. Mirror neurons are thought to be central to the brain’s ability to
learn through observation and imitation as well as to the development of gestural communication and spoken
language. Due to their connectivity with other social and emotional neural networks, mirror neurons are also
assumed to play a significant role in allowing us to feel both the pain of others in need and their joy when they
receive our help, central aspects of altruism and the therapeutic connection.

The Default-Mode Network


What has come to be called the default-mode network (DMN) has been found to be a coherent functional
network that becomes active when we are having reflective experiences of ourselves and others (Beckmann et
al., 2005; Mitchell, Banaji, & Macrae, 2005). It is likely that the DMN also serves as a neural platform for
social cognition, empathy, and altruism (Schilbach et al., 2008). The DMN, along with other neural
networks, appears to combine a sensory experience of the body and external world that allows us to have a
conscious experience of ourselves in imaginal space within the flow of time. Within this imaginal space we can
engage in relationships, take the perspective of others, solve moral dilemmas, and imagine future outcomes of
social behaviors (Maguire & Mummery, 1999; Iacoboni et al., 2005; Uddin et al., 2004; Spreng & Grady,
2010). This function is vital to the development of the experience of subjectivity and self, and our appreciation
of the experience of others (Morcom & Fletcher, 2007).

TABLE 14.1

146
The Neurobiology of Altruism

Altruism Activates Supports


Oxytocin Bonding and attachment
Decreased stress and increased neuroplasticity
Dopamine Reward of helping others
Reward of seeing others punished for asocial behaviors
Right amygdala Attention to the feelings of others
Orbitomedial prefrontal cortex (ompfc) Secure attachment, affect regulation, and emotional safety
Theory of mind and comprehending the experiences of others
Temporal-parietal junction (TPJ) Self-agency, self-efficacy, self-esteem
The sense of personal responsibility
An increased consideration/understanding of feedback
An enhanced ability to understand the experience of others
Empathy
Superior temporal zone (STZ) Connection and behavioral synchrony with others
Mirror neurons Imitation, attunement, empathy, theory of mind
Default mode network (DMN) A coherent sense of self
The ability to comprehend the internal states of self and others

We are most likely attempting to activate this circuit when we encourage clients to shift from a posture of
defensiveness to one of open self-reflection or from thoughts about the external world to reflect on distressing
emotions and memories. This may well be the circuit that becomes most active when an interpretation hits
home and a client’s reflexive defenses stop short (see Table 14.1). The facial expressions that appear during
this state often reflect the emotions clients have been defending against.
All of these neural circuits that shape the brain into a social organ serve as the infrastructure for language
and culture. Religious beliefs and political institutions that emerge as part of culture reflect the challenges of
navigating our interwoven drives toward self-preservation and caring for others. Our religions and political
beliefs are therefore important clues as to how our brains and minds struggle to process social information. As
therapists, we need to be open to and accepting of many of our clients’ beliefs while being mindful that these
beliefs hold valuable clues about how their brains might be organized.

Altruism in Religion and Politics


The modern conservative is engaged in one of man’s oldest exercises in moral philosophy: the search for a superior moral justification for selfishness.
—John Kenneth Galbraith

The coevolution of love and aggression increased our ability to attach to and protect those closest to us (Choi
& Bowles, 2007). The juxtaposition of these sometimes conflicting instincts may help us to understand our
vulnerability to domestic violence and jealous rage or why some people can act so kindly toward some, but so
cruelly toward others. The battle between empathy and selfishness, played out in everyday life, is the focus of
our secular laws and religious commandments.
Most religious traditions promote harmony and unity among all people, and many place altruistic acts at
the center of their belief systems. Despite the diversity of spiritual traditions, the insight of human
connectedness—species consciousness—is a consistent theme. To this day, those who consider themselves to
be religious both feel and are seen by others as more altruistic, empathic, and prosocial (Saroglou, 2013;
Saroglou et al., 2005). This suggests to us that species consciousness is part of a broad adaptational process
that resides deep within us.
Our conflicting instincts toward selfishness and altruism are reflected in the polarization between
conservatives and liberals. Conservatives are commonly characterized as being self-centered and protective of
their own interests. In contrast, liberals are seen as being too concerned with issues of justice and equality
while not attending enough to fiscal realities. The broadening divide between conservatives and liberals that
has stalled our political system has even drawn the attention of neuroscientists. While just a few studies have
compared the brains of conservatives and liberals, some interesting findings have emerged.
As you may have already anticipated, the brains of liberals have a larger anterior cingulate, a region
dedicated to both attachment and fear regulation (Jost & Amodio, 2012). Liberals also have higher levels of
activity in their left insula when assessing risk, suggestive of a greater sensitivity to their own internal states
and the internal states of others (Kanai et al., 2011). Put in a slightly different way, they feel more themselves

147
and feel more for others, making empathy a strong guiding factor when making political decisions. Liberals
may also be better able to regulate their fear in interpersonal situations. In contrast, the brains of conservatives
appear biased toward seeing the world as dangerous while being less focused on the experience and needs of
others. Conservative brains have greater amygdala volume, right amygdala activity, and greater fear reactivity
in the face of a threat. At the same time, conservative brains show decreased volume in the anterior cingulate,
suggesting both a lack of activation in attachment circuitry and a decreased ability to be alert to more subtle
differences in arguments (Amodio, Jost, Master, & Yee, 2007; Schreiber et al., 2013). For conservatives, issues
are seen as more a function of right and wrong with clear solutions.
The difference between self-other focus found in the brains of conservatives and liberals when faced with
political issues impacts the way they frame problems and search for solutions. Fear (driven by amygdala
activation) makes all of us more self-protective, causing us to think in black and white terms and become more
rigid in our beliefs. Thinking about others requires the activation of attachment circuits, which inhibit the
amygdala and activate the anterior cingulate, allowing those individuals to distinguish shades of gray and
tolerate the anxiety of uncertainty. Thus, a client’s political beliefs may give us a clue as to how his or her brain
is organized and how we might adapt therapy to fit them. Politics appears to tap into the altruistic drive of
liberals and inhibit it in conservatives. Of course, the real-world adaptive value of either position depends on
the situation. It would have been much better for Neville Chamberlin to be more suspicious of Adolf Hitler,
while the British would have been well served if they had had more empathy for Gandhi’s call for an
independent India.
The research has yet to be done to address whether conservatives and liberals behave differently when it
comes to engaging in altruistic acts. We don’t know which group, if either, would be more likely to go out of
their way to help a neighbor, give to charity, or sacrifice their own needs for the needs of others. Perhaps
conservatives are more generous with people they know while liberals have more altruistic attitudes toward
strangers and are less generous with intimates. The important issue for our present discussion is that these
differences in attitudes may give us insight into the biases built into the organization of our clients’ neural
circuitry.

Guided Altruism as an Adjunct to Psychotherapy


Every man must decide whether he will walk in the light of creative altruism or in the darkness of destructive selfishness.
—Martin Luther King Jr.

Including guided altruism as an adjunct to psychotherapy may result in a cascade of positive effects that could
enhance therapeutic success by activating specific neurobiological and psychological processes:

1. Decreased stress, a deepened relationship with the therapist, and greater neural plasticity mediated via
increases in oxytocin and dopamine, activation of the attachment circuitry of the ompfc, and amygdala
inhibition.
2. An increase in self-awareness, self-reflective capacity, perspective taking, openness to feedback, and a deeper
internalization of the therapeutic experience mediated via activation in the relevant zones of the temporal
lobes, prefrontal lobes, and the DMN.
3. An increase in self-esteem, self-agency, personal responsibility, and empathy mediated via activation of the
TPJ, STZ, ompfc, right amygdala, and mirror neurons.

Initially, altruism can activate the regions associated with bonding and secure attachment. This enhances the
platform from which client and therapist collaborate and creates states of brain and mind that are more open
to neural growth and learning. The next step, psychological mindedness, allows clients to use language and
self-reflection to expand awareness of themselves and the world. The third step is to combine emotional
security with enhanced perspective by changing the way they experience themselves in the world—from victim
to actor, from helpless to helpful, from beta to alpha status.
We all develop a personal narrative that consists of the stories we tell about ourselves. This becomes the
center of gravity of our own identity and serves as a blueprint for future behavior. Given that our self-identity
is embedded in the context of relationships, having stories to tell about ourselves in which we are the giver,
caretaker, or benefactor creates a narrative that will naturally enhance efficacy and self-esteem. We are more
likely to engage with others and garner positive feedback, which also becomes part of our story. Being
altruistic makes us a contributor instead of a victim, efficacious instead of helpless, and allows us to connect to

148
the social world, creating opportunities for reconnection, transformation, and redemption.
Alcoholics Anonymous (AA) is an excellent example of how altruism has long been utilized as a tool of
personal transformation in a nontherapeutic environment. Values within AA, articulated in the Big Book and
the Twelve Steps, include a commitment to the common welfare of the group, bringing the message of AA to
other alcoholics, and serving as a sponsor to others who are newer in their sobriety (Alcoholics Anonymous,
2015). Another utilization of altruism is the ninth step, during which individuals make amends to those they
have hurt in the past. This involves those aspects of the altruism circuit that include self-reflection, perspective
taking, empathy, and humility. Most members of AA recognize the importance of giving to others and
making amends to those they have hurt as key components of their sustained sobriety.
I would suggest that guided altruism become part of psychotherapy in the middle stages of therapy—after
any initial crises and presenting problems have been addressed and the therapeutic relationship is established.
It should be considered with clients who seem especially obsessive about their own emotions, past
victimization, or feelings of powerlessness in their own lives. The choice of where, when, and how they
become engaged in altruistic activities should be thoughtfully considered to optimize the match with their
strengths, weaknesses, and personal needs. For example, volunteering at a children’s hospital, homeless
shelter, or hospice community can be inspiring for some but overwhelming for others. Time spent reading to
children at a public library or being a dog walker at an animal shelter may be as much as some clients can
handle. The goal of guided altruism is to help clients to feel like they are making a positive contribution to
society, to experience feelings of self-efficacy, and to see that they are not alone in the struggle for meaning
and love.
The idea that serving others is a path to self-growth is far from new. Many religious traditions, including
Christianity and Buddhism, believe that service to others leads to health, healing, and spiritual growth.
Naikan therapy, a school of Japanese healing, equates mental illness with a lack of appreciation of others
(Reynolds, 1980). Treatment consists of two weeks of guided meditation and journaling focused on
remembering examples, no matter how small, of what others have done for you. From the Naikan perspective,
self-centeredness leaves us feeling isolated, angry, and disconnected from the group mind, resulting in a kind
of altruism deficit disorder. Naikan therapy is designed to reactivate your appreciation of others in order to
reestablish a balanced mind. Adding guided altruism to psychotherapy is well in line with these long-standing
practices and traditions.

Looking at Satan by the Light of Paradise


Altruism is innate, but it is not instinctual. Everybody’s wired for it, but a switch has to be flipped.
—David Rakoff

Literature and folklore contain numerous stories in which acts of altruism lead to the kinds of personal
transformation that we often strive to attain in psychotherapy. One of my favorites serves as the launching
point for Victor Hugo’s (2013) Les Miserables when the criminal Jean Valjean encounters a kindly and
generous bishop. When they meet, Valjean is struggling with the shame, anger, and trauma born of his
painful past. Recently released from prison after serving time for stealing bread, Valjean feels spurned by
society, angry, and struggling to survive.
Expecting no pity or grace, he is surprised to be welcomed into the rectory of a bishop where he is offered
friendship, food, and lodging. Suspicious of the bishop’s intention, as he has only known cruelty, Valjean eats
dinner and retires with the rest of the household. Untouched by the bishop’s kindness, Valjean waits for
everyone to fall asleep, quietly collects the silverware, and steals away into the night. The following day, he is
quickly apprehended and brought back to the rectory for identification. When the bishop responds to a knock
on his door, he finds Valjean in the clutches of the gendarmes, silverware in hand.
Instead of identifying Valjean as a thief and sending him back to prison, the bishop tells the gendarmes
that there must have been a mistake; Valjean is his friend and the silverware was given to him as a gift. The
bishop then tells Valjean that he forgot to take the matching candlesticks as his confused captors walk away.
The bishop then pulls Valjean close to him and says: “You have promised me to become an honest man. I am
purchasing your soul. I withdraw it from the spirit of perdition, and I give it to God!” As Valjean walks away
from the bishop’s house, something in him breaks and his previous understanding of the world crumbles;
Valjean is in crisis.

He dimly felt that this priest’s pardon was the hardest assault, the most formidable attack he had ever sustained; that his hardness of

149
heart would be complete, if it resisted this kindness; that if he yielded, he would have to renounce the hatred with which the acts of
other men had for so many years filled his soul, and in which he found satisfaction; that, this time, he must conquer or be conquered,
and that the struggle, a gigantic and decisive struggle, had begun between his own wrongs and the goodness of this man. (Hugo, 2013,
p. 110)

What might be happening in Valjean’s brain? I can imagine that the bishop’s compassion activated oxytocin
that resulted in feelings of being cared for and positive social emotions that Valjean had not experienced for
years. The feelings evoked in him made Valjean question the need for his defense of hardness and hatred.
Activity within the TPJ, STR, and DMN may have allowed Valjean greater self-reflective capacity and a
broader perspective on himself, giving him the ability to imagine that he was capable of change. Because
oxytocin activation results in a decreased vigilance to external threat, there is greater neuroplastic flexibility
making it more likely that his experiences would be translated into ongoing changes within his brain.
In this state of crisis, Valjean encounters a little boy, Petit Gervais, and steals a coin from him. While in
the past he would have done this without a second thought, he now possesses an internal perspective that is
watching and considering his own behavior. Instead of a simple reflexive behavior, he now sees his impulse to
steal as an evil instinct. He reenvisions his petty thievery as the act of a brute, a beast inside him, and thinking
about his behavior makes him recoil in horror. His actions, driven by the anger and hatred for how he has
been treated, now feel alien and repulsive.

It was a strange phenomenon, possible only in his current condition, but the fact is that in stealing this money from the child, he had
done a thing of which he was no longer capable. He could see his life, and it seemed horrible; his soul, and it seemed frightful. There
was, however, a gentler light shining on that life and soul. It seemed to him that he was looking at Satan by the light of Paradise. (p.
111)

The point of this case study is to suggest that the bishop’s altruistic act led to a cascade of changes in emotion,
thought, and behavior that led Valjean to evolve from an angry beast into a compassionate human being. He
now had an inner world that included the ability to reflect on his experiences, to consider new ways of being,
and to have empathy for himself and others. No longer a prisoner of his impulses, there now existed the
possibility of joining the group mind and creating positive human relationships. The result was a new and
more adaptive narrative of the world that included self-reflection and personal responsibility.
If the bishop were a therapist, he would certainly have counted his brief intervention with Valjean as a
treatment success. If you don’t know the rest of the story, Valjean went on to become the champion of a
young orphan, Cossette, whom he raised and protected from the harsh world that had taken such a toll on
him. As for so many others, a core component of his healing was giving a child what he never received as a
child. He also went on to become a man of character and an upstanding citizen who helped many others.

Summary
When will our consciences grow so tender that we are moved to prevent suffering, not just avenging it?
—Eleanor Roosevelt

There is a good possibility that having clients engage in guided altruistic acts as an aspect of treatment will
enhance neuroplasticity and positive change via altruism’s natural impact on the brain, mind, and
relationships. Activating altruism circuitry appears to target many of the brain regions we need to optimize the
impact of psychotherapy. Activating the DMN during altruistic behavior, as an example, will likely support
the development of self-reflective capacity in many clients. As we know, being able to see yourself as others do
is a critical component of social intelligence. It is the other side of being able to grasp that others have an
internal world, feelings, and a perspective that may be quite different from our own.
I strongly suspect that Carl Rogers activated these regions through the use of warmth, caring, and
unconditional positive regard. This may be why the quality of the therapeutic relationship is consistently the
most influential factor in achieving a successful outcome. Further, it makes sense that simultaneously being on
either end of altruism, as a client and as a positive social agent, will serve to enhance the activation and growth
of our altruism circuitry and therapeutic successes.
Using guided altruistic behavior between sessions may serve to integrate the often isolated experience of
therapy into the community and enhance a client’s sense of self-efficacy, competence, and inner strength. It

150
could counterbalance the risk of dependency on the therapist and an associated decrease in self-efficacy.
Therapists should also consider the potential value of having groups that are organized around community
service where self-exploration is balanced with working as a team for a common cause. In this way, the group
can grow into a tribe through shared goals, teamwork, and group achievement.
Evolution is an extremely uneven process. While many individuals are still engaged in tribal warfare,
others are beginning to look beyond these proximate affiliations to a broader world. As life on earth becomes
increasingly crowded, complex, and conflictual, the need for meaningful and sustained prosocial behavior will
grow more pressing. “Paying it forward” may have to grow from a catchphrase into a prominent form of social
behavior, and altruistic behavior toward our species and the planet as a whole may become necessary for our
continued survival.

151
Part V
The Disorganization of Experience

152
Chapter 15
The Anxious and Fearful Brain
Fear is the oldest and strongest emotion of mankind.
—H. P. Lovecraft

All animals have been shaped by evolution to approach what is life sustaining and avoid what may be
dangerous. The success of rapid and accurate approach-avoidance decisions determines if an organism lives
long enough to carry its genes forward to the next generation. Because vigilance for danger is a central
mechanism of natural selection, evolution may favor an anxious gene (Beck et al., 1979). Some anxieties
appear to be hard wired in all primates and linked to our deep past. Fear of spiders, snakes, and heights all
harken back to the survival needs of our forest-dwelling ancestors. Our complex neural systems have all been
sculpted to serve the prime directive of survival.
The neural circuitry involved in fear and anxiety, although biased toward the limbic system and right
hemisphere, involves both hemispheres and all levels of the brain and spinal cord. The most primitive fight-
or-flight circuitry shared with our reptilian ancestors is interwoven with the most highly evolved regions of the
cortex. This connectivity results in the capacity to experience anxiety about everything from an unexpected tap
on the shoulder to an existential crisis and supports the theory that, at its core, all anxiety is connected to a
fear of death (Tillich, 1974).
Anxiety and fear are the conscious manifestations of the body’s ongoing appraisal of threat, telling us to be
prepared to take action. Anxiety can be triggered by countless conscious or unconscious cues and has the
power to shape our behaviors, thoughts, and feelings. At its most adaptive, anxiety encourages us to step back
from the edge of a cliff or to check to see if we signed our tax forms before sealing the envelope. At its least
adaptive, anxiety steers us away from taking appropriate risks or engaging in new and potentially beneficial
behaviors.
The response to stress results in a range of physiological changes designed to prepare the body for fight or
flight (Selye, 1979). Energy is mobilized through increased cardiovascular and muscular tone, whereas
digestion, growth, and immune responses are inhibited. A cascade of biochemical changes occurs in the
hypothalamus, pituitary, and adrenal glands (the HPA axis) and throughout the autonomic nervous system
mediating the physical and psychological manifestations of stress. Increased levels of glucocorticoids,
epinephrine, and endogenous opioids are particularly relevant to the psychological impact of stress and trauma
due to their impacts on attention, cognition, and memory. We experience the effects of stress across a range of
situations both negative and positive, from auto accidents to marriage, at crucial moments during sporting
events, and when engaging in public speaking. The dangers can be real, imagined, or experienced vicariously
as when we watch others in dangerous situations.
With the expansion of the cerebral cortex and the emergence of imagination, we have become capable of
being anxious about situations we will never experience. We can now worry about monsters living under our
beds and the incineration of the earth resulting from the sun’s expansion a billion years in the future. Because
our imaginal capabilities have allowed for the construction of the self, we can also become anxious about
potential threats to our psychological and social survival. Psychotherapists deal with a wide variety of anxiety
disorders based in the fear of a social death. The expectation of rejection by another can result in social
withdrawal; the fear of forgetting one’s lines in a play can result in stage fright. Systems of physical survival
have been conserved in the evolution of consciousness and the ego, to be triggered when threats to these
abstract constructions are activated.
We see this, for example, with adults who were sexually or emotionally abused as children who come to
therapy with severe weight problems. They do well on diets until they begin to be seen as attractive by others,
which is associated in implicit memory with the pain and shame of their childhood experiences. These
reactions lead them to eat to avoid their feelings, feel nurtured, and gain the weight that protects them against
sexual advances.
Consciously experienced anxiety provides the opportunity to face and work through one’s fears. The

153
common wisdom of getting back on the horse that threw you is advice clearly aimed at preventing the use of
avoidance to control anxiety. In fact, the reduction of anxiety through avoidance reinforces the power of the
feared stimulus. Unfortunately, anxiety becomes paired with all kinds of sensations, emotions, and thoughts,
unconsciously shaping our behavior. The avoidance of thoughts and feelings associated with feared stimuli
both reflects and perpetuates a lack of integration among neural networks. Compounding the problem, the
left-hemisphere narrator provides a rationale supporting and reinforcing avoidance: “It’s inhuman to ride
horses!” “Who needs planes?” “Why go out when it’s so comfortable at home?” Facing one’s fears is a core
component of all forms of psychotherapy.
Thus, what started out as a straightforward survival-based alarm system in our ancestors with small
cortices has become a nuisance for us. This makes an understanding of the neuroscience of anxiety and fear
helpful in the conceptualization and treatment of many clinical disorders. In the following pages, we will look
at the two loops of fear circuitry outlined by Joseph LeDoux, the role of the amygdala in the regulation of fear
and anxiety proposed by Michael Davis, and Robert Sapolsky’s work on the negative impact of long-term
stress.

Fast and Slow Fear Networks


Fear is an emotion indispensable for survival.
—Hannah Arendt

Through his research with animals, LeDoux (1994) demonstrated the existence of two separate yet
interrelated neural circuits that regulate fear. The conservation of these more primitive systems during primate
evolution allows us to apply many of LeDoux’s findings to human experience (Phelps, Delgado, Nearing, &
LeDoux, 2004). The two systems (which we will call fast and slow) each play a somewhat different role in our
reaction to danger. This model can be useful to describe to anxious and fearful clients to help them understand
the underlying neurobiological mechanisms of their unsettling experiences.
The reflexive fast system acts immediately, sending information directly from the sense organs (eyes, ears,
skin, nose, tongue) through the thalamus to the amygdala. And when I say fast, I mean fast: All of this
processing can occur in one-twelfth of a second. The amygdala evaluates the sensory input and translates it
into bodily responses via its many connections with the autonomic nervous systems. The thalamus may aid in
this rapid evaluation by maintaining crude representations of potentially dangerous things historically
encountered in the environment such as spiders, heights, and dangerous predators (Brosch, Sander, &
Scherer, 2007). The amygdala and subcortical structures that play an executive role in these rapid appraisals
and reactions have retained this role in the face of massive cortical development because the increased time it
would take to include the cortex must have had too large a survival cost.
As the fast system is reacting, the information is being sent through the slow system, the hippocampus and
cortex, for further evaluation. This system is slower because it contains more processing hubs and more
synaptic connections, and involves very slow conscious processing. Cortical circuits examine the information
more carefully, compare it to memories of similar situations, and decide how to proceed. The slow circuit aids
in fear processing by contextualizing the information in time and space. This slow system in humans has the
additional task of making sense of the behavioral and visceral reaction already set into motion by the fast
systems. In this way, our conscious executive functions discover the decisions that have already been made by
our primitive executive. We find ourselves already scared when we initially perceive what is frightening us, or
ecstatic as our loved one comes into view. Figure 15.1 depicts the neural circuits of the fast and slow fear
networks.
This fast and slow dual circuitry helps us to understand why we often react to things before thinking and
then have to apologize later on. In therapy, we often attempt to utilize the conscious linguistic structures of
the slow circuit to modify or inhibit dysfunctional reflexes and emotional appraisals of the fast circuit.
Coupled with relaxation techniques and conscious awareness and understanding of the process, exposure to a
feared stimulus can serve to enhance the regulatory input of the slow cortical circuits by building new neural
connections that increase the ability of the cortex to inhibit the amygdala.
As an example of these two systems in action, I walked into my garage one day to look for a tool when, out
of the corner of my eye, I saw a small brown object near my foot. There are plenty of little critters in my
neighborhood, which often crawl, burrow, or fly into my house. I immediately jumped back, my heart rate
increased, my eyes widened, and I became tense, ready to act. Moving backward, I oriented toward the shape,
saw that it looked more like a piece of wood than a rodent, and began to relax. After a few seconds, my heart

154
rate and level of arousal were back to normal; the potential danger had passed.

FIGURE 15.1
Fast and Slow Fear Circuits
A depiction of the two pathways of information to the amygdala—one directly from the thalamus and the other through the cortex and hippocampus
(adapted from LeDoux, J. “Emotion, memory, and the brain.” Copyright ©1994 by Scientific American, Inc. All rights reserved.)

Analyzing this experience in terms of the two systems, my peripheral vision saw the object and my
amygdala appraised it in an overgeneralized fashion as a threat. My amygdala activated a variety of
sympathetic responses including startle, increased respiration, adrenalin output, and behavioral avoidance. As
my body was reacting, I reflexively oriented toward the shape, which brought it to the fovea of my retina,
providing my hippocampus and cortex with more detailed visual information. This allowed my slow system to
appraise the situation more accurately than my skittish amygdala. Our species-specific fear of dangerous
insects and small creatures accounts for such a strong reaction to an animal weighing just a few ounces. This
example, trivial as it may be, leads to a more serious application of LeDoux’s theory to interpersonal
relationships.
As the core of our social brain, the amygdala organizes the appraisals of what we have learned from our
relationship history. In interpersonal situations, our amygdala reflexively and unconsciously appraises others in
the context of our past experiences. In intimate relationships, this is called attachment schema. From moment
to moment, the reflexive activations of our fast systems (organized by past learning) shape the nature of our
present experience (Bar et al., 2006). This is a powerful mechanism by which our early social learning
influences our experience of the present. So, by the time we become conscious of others, our brain has already

155
made decisions about them. In the case of prejudice, skin color triggers a set of assumptions upon which we
evaluate other people (Olsson, Ebert, Banaji, & Phelps, 2005). At the opposite extreme, love at first sight is a
positive prejudice triggered by oxytocin-driven emotional memories projected onto another person.

The Amygdala’s Role in Anxiety and Fear


No passion so effectually robs the mind of all its powers of acting and reasoning as fear.
—Edmund Burke

Although the amygdala is an organ of appraisal for both positive and negative stimuli, its most primitive and
basic function is to translate danger into the activation of the fight-flight response. Electrical stimulation of
the amygdala’s central nucleus results in the experience of fear, whereas destruction of the amygdala will
eliminate prior fear reactions and result in an inability to acquire a conditioned fear response (Carvey, 1998).
It has been conserved, expanded, and more widely connected with other brain regions during evolution in
order to process increasingly complex cognitive, emotional, and social input. It is central to memory
processing, emotional regulation, attachment, and sophisticated social interactions.
Although we are genetically programmed to be fearful of things like snakes and social exclusion, a fear
response can be quickly learned by pairing any thought, feeling, or sensation with a noxious or painful
stimulus (Corcoran & Quirk, 2007). This one-trial learning is the amygdala’s appraisal mechanism at work.
The amygdala is capable of pairing any stimulus (even positive physical approach, affection, or praise) with
fear. Like the hippocampus, the lateral areas of the amygdala are capable of long-term potentiation (LTP)
involved in reinforcing connections among neurons. Remember that LTP is believed to be the process
through which the association among neurons becomes strengthened and learning is established.
As we saw earlier, the hippocampus and amygdala organize interacting but dissociable systems of memory.
Bechara and colleagues (1995) reported that a patient with bilateral (left and right) amygdala damage was
unable to acquire a conditioned autonomic response to sensory stimuli. The patient was, however, able to
consciously remember the conditioning situation because his hippocampi were still intact. Another patient
with bilateral damage to the hippocampus showed no conscious memory for the conditioning situation but
did acquire autonomic and behavioral conditioning. In other words, the patient experienced a fear reaction but
had no memory for learning it. The authors concluded that the amygdala is “indispensable” for coupling
emotional conditioning with sensory information while the hippocampus is required for conscious recollection
and episodic memory (Bechara et al., 1995).
The neural projections from the amygdala to numerous anatomical targets cause multiple physical
expressions of anxiety, fear, and panic. Projections from the amygdala to the lateral hypothalamus result in
sympathetic activation responsible for increased heart rate and blood pressure. The amygdala’s stimulation of
the trigeminal facial motor nerve even causes the facial expressions of fear (Davis, 1992). The amygdala is also
essential in reading the fearful facial expressions of others (Baird et al., 1999). As you can see from Figure
15.2, the amygdala is well connected, making the fear response a powerful whole-body experience.
Panic attacks result from triggering of the autonomic nervous system by the amygdala in the absence of an
apparent external danger. Victims of panic attacks have been shown to have increased neural activity in the
amygdala (Reiman et al., 1989). The experience of panic attacks is so powerful that sufferers often go to
emergency rooms, certain that they must be dying of a heart attack. Psychologically, victims report a sense of
impending doom, feelings of unreality, and a fear of going crazy. Panic attacks are often precipitated by stress
or other conflicts in the sufferer’s life, but he or she seldom makes the connection between these events and
the panic attacks. Because the neural connections are contained within hidden neural neural layers that
operate below conscious awareness, they are experienced as coming out of the blue, leaving victims feeling
confused and overwhelmed, resulting in a narrowing of thoughts, feelings, and activities.

FIGURE 15.2
Some Targets of the Amygdala in the Fear Response
Some of the many anatomical targets of the amygdala in the fear response, and their biological and behavioral contribution.

156
Because the job of the amygdala is to protect us from danger, it tends to generalize from one fearful
stimulus to a broader array of internal and external cues (Douglas & Pribram, 1966). For example,
agoraphobia, a fear of open spaces, develops as victims of panic attacks associate fear with a broader variety of
situations. Hoping to avoid these attacks, sufferers restrict their activities to the point where they eventually
become housebound. Simultaneously, the amygdala becomes conditioned to respond faster in people who
become phobic, creating a vicious cycle of anxiety and fear (Larson et al., 2006). On the other hand, those of
us with a slower and less active amygdala experience greater psychological well-being (van Reekum et al.,
2007). One gift of aging is that the amygdala also appears to become less sensitive to fear as we grow older
(Mather et al., 2004).
The development and connectivity of the amygdala have many implications for both early child
development and psychotherapy. Because the amygdala is operational even before birth, the experience of fear
may be the strongest early emotion. Without the inhibitory impact of the later-developing hippocampal-
cortical networks, early fear experiences are unregulated, overwhelming full-body experiences. Part of the
power of early emotional learning may be the intensity of these unregulated negative emotions in shaping our
early neural infrastructure. The infant is totally dependent on caretakers to modulate these powerful negative
bodily states and emotions. Amygdala and hippocampus-mediated memory systems are dissociable from one
another, which means that early and later traumatic memories can be stored without conscious awareness or

157
cortical control. They will, however, emerge as sensory, motor, and emotional memories in the form of
flashbacks, nightmares, and panic attacks.
Another limbic structure closely connected to the amygdala is the bed nucleus of the stria terminalis
(BNST). In fact, many consider the BNST to be part of what is referred to as the extended amygdala. Like
the rest of the amygdala, it is connected upward to the prefrontal cortex, as well as down into the autonomic
nervous system. Unlike the core structures of the amygdala, the BNST is sensitive to abstract cues mediated
by the prefrontal cortex and capable of long-term activation (Lebow & Chen, 2016; Somerville et al., 2010;
Walker, Miles, & Davis, 2009). These two abilities are suggestive of its later evolution and its role in
anticipatory anxiety (Davis, 1998; Kalin, Shelton, Davidson, & Kelley, 2001). It may be that the central
components of the amygdala specialize in fear while the BNST evolved to deal with the more complex
triggers for anxiety that emerged as we became more capable of imagining future outcomes. Interestingly, the
BNST in rats is a structure that grows in response to maternal responsibilities. We have to wonder if, as the
brain became specialized for caretaking, the scaffolding we need to create around our children to protect them
also pushed the evolution of our ability to anticipate potential dangers. It is also interesting to reflect on the
statement of so many parents who say that to have children is to worry and that you always worry about your
children no matter how old they are.

The Locus Coeruleus and Norepinephrine


Worry gives a small thing a big shadow.
—Swedish proverb

One important descending projection from the amygdala and BNST connects them with the locus coeruleus
(LC). The LC is a small structure with extensive projections throughout the brain stem, midbrain, and
cerebral cortex. It is, in fact, connected with more parts of the brain than any other discovered so far (Aston-
Jones, Valentino, VanBockstaele, & Meyerson, 1994). The LC is the brain’s primary generator of
norepinephrine (NE), which drives the activity of the sympathetic nervous system and the fight-or-flight
response. One effect of NE is to enhance the firing of neurons that are highly relevant to a present experience
based on a past fear response, while inhibiting activities involved in baseline activities.
This means that stimulation of the LC prepares us for danger by activating circuits dedicated to attention
and preparation for action. NE activation makes us scan for danger and maintain a posture of tense readiness.
It also heightens our memory for danger, creating a sort of screen shot for amygdala memory circuits
(Livingston, 1967). The pathways containing these traumatic memories become hyperpotentiated, meaning
that they are more easily triggered by less severe subsequent stressors. This allows us to be reminded in the
future of similar dangers. During times of lowered hippocampal-cortical involvement (e.g., intoxication or
near-sleep states), these stressful traumatic memories may become disinhibited as intrusive images and
flashbacks. In English, this means that surges of NE during periods of safety may result in past traumatic
associations (anxiety, startle, visual images, etc.) being brought to awareness, which overshadow current
experiences.
Electrical stimulation of the LC in animals results in a disruption of ongoing behavior and triggering of an
orienting reflex, like the one I had to the small piece of wood. This is seen in patients with PTSD who
experience amygdala activation and autonomic arousal responses to trauma-related cues after many decades.
LC activity in primates results in a high degree of vigilance while interrupting sleep, grooming, and eating.
Through a series of connections, the central nucleus of the amygdala stimulates the LC, which, in turn, is
thought to be a major control area of the sympathetic nervous system (Aston-Jones et al., 1994). An
understanding of the biochemistry and functioning of the LC is an important component of an understanding
of anxiety disorders (Svensson, 1987).

Stress and the Hippocampus


Anxiety is a thin stream of fear trickling through the mind. If encouraged, it cuts a channel into which all other thoughts are drained.
—Arthur Somers Roche

The human brain is well equipped to survive brief periods of stress without long-term damage. In an optimal
state, stressful experiences can be quickly eluded or resolved. However, people often come to psychotherapy

158
with long histories of anxiety, which have taken a toll on both their coping abilities and their brains. Working
with rats and vervet monkeys, Sapolsky and his colleagues demonstrated that sustained stress results in
hippocampal atrophy and a variety of functional impairments (Sapolsky, 1990; Sapolsky, Uno, Rebert, &
Finch, 1990). His research is particularly important because it may help explain some of the negative long-
term effects of childhood stress.
The biological link between prolonged stress and hippocampal atrophy appears to be mediated via the
catabolic influence of stress hormones. Glucocorticoids (GCs) such as cortisol are secreted by the adrenal
gland to promote the breakdown of complex compounds so that they can be used for immediate energy. The
first of these hormones was found to break down complex sugars, hence the name glucocorticoid. It was later
found that they also block protein synthesis, inhibiting both new neural growth and the construction of
proteins involved in immunological functioning. Overall, long-term learning and biological well-being are
sacrificed for the sake of immediate survival. This makes great sense when stressors are short-lived. But when
stress is chronic, sustained high levels of cortisol put us at risk of physical illness as well as learning
dysfunctions and memory deficits. A number of roles of cortisol are listed in Table 15.1, along with its impact
on the brain and its relationship to a variety of illnesses.
During a fight-flight response, the focus on immediate survival supersedes all long-term maintenance, akin
to burning the furniture to survive freezing in winter. This system was designed to cope with brief periods of
stress in emergency situations; it was not designed to be maintained for weeks or years at a time. These
biological processes need to be reversed as soon as possible after the danger has passed to allow the body to
return to restoration and repair. The complexities of cortical processing and anticipatory anxiety are poorly
matched with our primitive stress system.
Prolonged stress inhibits protein production, diverting the energy needed to maintain higher levels of
metabolism. Proteins are, of course, the building blocks of the immunological system (leukocytes, B-cells, T-
cells, natural killer cells, etc.) and the suppression of protein synthesis also suppresses our body’s ability to
fight off infection and illness. This is one of the primary reasons for the relationship between stress and
illness. The resulting higher levels of metabolism cause sodium to continue to be pumped into neurons,
eventually overwhelming the cell’s ability to transport it out again. This results in destruction of the cell
membrane and consequent cell death. This process has been found to be particularly damaging to
hippocampal neurons, resulting in a variety of memory deficits and depression. Loss of volume in the
hippocampus has been found to be related to cumulative GC exposure (Sapolsky et al., 1990). Sustained high
levels of stress partly explain why early negative experiences in parenting and attachment have a lifelong
impact on physical health, mental health, and learning.
The hippocampus, rich in GC receptors, plays a negative feedback role with the adrenal gland to inhibit
GC production. If the hippocampus detects too many GCs together, it sends a message (via the hypothalamus
or the pituitary) to the adrenal gland to slow down GC production (Sapolsky, Krey, & McEwen, 1984). The
more receptors we have, the greater our feedback abilities to decrease cortisol production. Prolonged high
levels of GCs increase the vulnerability of the hippocampus to a number of potential metabolic insults
(Sapolsky, 1985; Woolley, Gould, & McEwen, 1990). High-dose cortisol administration for autoimmune
diseases may also result in hippocampal damage (Sapolsky, 1996). At this point, it is unclear if decreased
volume reflects permanent damage to the hippocampus or a reversible inhibition of the growth of new
neurons. In either case, less hippocampal mass means fewer GC receptors, which, in turn, means less negative
feedback to the adrenal gland.

TABLE 15.1
Stress and the Hippocampus

The Role of Cortisol


Breaks down fats and proteins for immediate energy
Inhibits inflammatory processes
Inhibits protein synthesis within the immune system (leukocytes, B and T-cells, natural killer cells, etc.)
Suppresses gonadal hormones that support neural health, growth, and learning

Chronic High Levels of Cortisol/Glucocorticoids Result In


Decreased plasticity1

Dendritic degeneration2
Deficits of remyelination3

Cell death4

Inhibition of neurogenesis and neural growth5

159
High Levels of Cortisol Correlate With
Impaired declarative memory and spatial reasoning6

Compromised Hippocampi Result In


Deficits of new learning7
Deficits in short- and long-term memory8

Individuals With Smaller Hippocampi Include or Have


Adult victims of early trauma9
Posttraumatic stress disorder10
Temporal lobe epilepsy11
Schizophrenia12
Cushing’s disease (hypercortisolism)13

Early trauma results in hippocampal impairment, which decreases our ability to regulate and inhibit
emotions. Further, deficits in short-term memory and reality testing will make the process of integrating
traumatic experiences into conscious awareness more difficult. This means that longer periods of relationship
building and pragmatic stress reduction techniques may be necessary when working with victims of early
trauma. The hippocampus is also exquisitely sensitive to oxygen deprivation, so patients who have a history of
cardiovascular problems, head trauma, or seizures may have hippocampal compromise, as well as mountain
climbers and divers, who are apt to experience periods of oxygen deprivation (Lombroso & Sapolsky, 1998;
Regard, Oelz, Brugger, & Landis, 1989). All of these factors should be kept in mind when taking histories of
patients with cognitive and neurological symptoms.
Impairment of the hippocampus from early chronic stress may make the therapeutic process more difficult
for many clients. For example, Stein, Koverola, Hanna, Torchia, and McClarty (1997) found that adult
women who had experienced childhood abuse had significantly reduced left hippocampal volume. They also
found that the amount of reduction was significantly correlated with increased dissociative symptoms. This
relationship supports the role of the left hippocampus in semantic memory narratives and the flexible
incorporation of new information into existing structures of memory (Eichenbaum,1992). If this is the case,
early abuse may result in damage to neural structures required to update old narratives and create new ones.
Rats and humans differ in a number of ways besides whisker length. The increased size of the human brain
and its additional processing capacity make it possible for us to worry about many more potential dangers,
both real and imagined. In addition, our brains allow us to create complex situations such as traffic jams and
overburdened schedules, generating ever-increasing levels of stress. Stressors such as these, especially when
experienced as inescapable, tend to result in greater cortisol activation and negative impact on the brain.
Although we like to think of childhood as a time of innocence and play, many children grow up in a state of
constant distress. We saw this clearly in the attachment research where adults with anxious attachment
demonstrated a lack of recall for long periods of their childhoods. Parental physical or mental illness,
community violence, poverty, and many other factors can contribute to chronic cortisol release. Prolonged
childhood stress can have lifelong effects on functioning related to hippocampal damage, immunological
suppression, and other stress-related impairments.

Learning Not to Fear


Courage is acting in spite of fear.
—Howard W. Hunter

It is an unfortunate twist of evolutionary fate that the amygdala is mature before birth while the systems that
inhibit it take years to develop. This leaves us vulnerable to overwhelming fear with little to no ability to
protect ourselves. On the other hand, evolution has also provided us with caretakers who allow us to link into
and utilize their developed cortices until ours are ready. The way they protect us from fear and modulate our
anxiety becomes a template within which our own brain develops. Thus, we use proximity to our parents as
our key method of fear regulation, just as cold-blooded animals use locomotion and change of location to
regulate their internal temperature. Our attachment schemas come to reflect the success or failure of how we

160
and our parents navigate this process. We have seen from the research with rats that more maternal attention
results in a brain that is better equipped to regulate stress, learn, adapt, and survive.
While the hippocampus is constantly remodeled as we learn new information and skills, the amygdala’s
role is to remember each and every threat and carry them into the future. Because the amygdala exhibits
persistent dendritic modeling, we are unable to completely forget painful and traumatic experiences (Rainnie
et al., 2004; Vyas, Bernal, & Chattarji, 2003; Vyas & Chattarji, 2004). The power of the amygdala and its
dedication to protecting us from all possible danger leads us to be biased toward anxiety and fear.
Getting past our fears and phobias does not entail forgetting to be afraid; rather, the extinction of fears
represents new learning organized by our slow systems of the cortex and hippocampus. In other words,
extinction learning represents the formation of new neural associations that somehow keep the memory stored
in the amygdala from triggering the sympathetic nervous system (Milad & Quirk, 2002; Rau & Fanselow,
2007). The reappearance of old fears and phobias demonstrates how the fear we hoped was long gone was
stored in our amygdala all along (Vansteenwegen et al., 2005).
The ability of the prefrontal cortex to modulate the amygdala is accomplished through the development of
descending inhibitory circuitry (Akirav & Maroun, 2007; Ochsner et al., 2004). This cortical-amygdala
network exhibits a reciprocal activation pattern; more cortical activation results in less amygdala activation and
vice versa. This is why our problem-solving abilities decline when we are afraid and also why being well
prepared for a situation lessens our fears. When we successfully use cognitive techniques to decrease anxiety,
we are building these descending cortical networks (Schaefer et al., 2002).
The central nucleus of the amygdala is an output region that projects to sites in the midbrain and
hypothalamus responsible for generating different aspects of the fear response. The connections of the ompfc
to the central nucleus of the amygdala are particularly strong, especially to GABAergic (inhibitory) neurons
called intercalated cells (Freedman, Insel, & Smith, 2000; McDonald et al., 1999; Royer, Martina, & Paré,
1999). It is now believed that it is within the descending networks from the ompfc to the amygdala’s central
nucleus that extinction learning is remembered and carries out its inhibitory influences (Gottfried & Dolan,
2004; Quirk, 2004). Because learning in this neural circuit conforms to what is known about the neurobiology
of learning in general, the role of NMDA receptors, protein synthesis, cortisol, and other factors that
modulate learning are likely involved in extinction learning (Elvander-Tottie et al., 2006; Santini et al., 2004).
Learning not to fear, just like secure attachment, is an important accomplishment of the ompfc (Morgan,
Romanski, & LeDoux, 1993; Phelps et al., 2004). Electrical stimulation of the homologous region in the
cortex of rats results in both amygdala inhibition and a reduction of conditioned fear (Milad, Vidal-Gonzalez,
& Quirk, 2004; Pérez-Jaranay & Vives, 1991; Quirk, Likhtik, Pelletier, & Paré, 2003). In humans, the size of
the ompfc is positively correlated with our ability to inhibit a fearful response (Milad, Quinn, et al., 2005).
Thus, it appears that cortical control of the amygdala allows us to discontinue a fearful response to something
that makes us afraid. Subjects taught to interpret a surprised face as negative had greater amygdala activation
while those who were guided to see it as positive had more ompfc activation (Kim et al., 2003). These studies
support the notion that the ompfc modulates the activation of the amygdala based on contextual and
motivational factors (Kim et al., 2005; Myers & Davis, 2007; Ochsner, Bunge, Gross, & Gabrieli, 2002; Phan
et al., 2005). Put in another way, the slow system is capable of modulating the fast system.
These top-down circuits organize, modulate, and direct attention in ways that shape experience and
reinforce the existing emotional state (Bishop, 2007; Christakou, Robbins, & Everitt, 2004). Anxiety is
associated with a reduced top-down control of threat cues just as there is a reduction of control over negative
stimuli in depression (Bishop et al., 2004; Brewin & Smart, 2005). In other words, anxious people tend to
find danger while depressed people discover the negative aspects of their environments. And while those of us
with more attentional control will still have a bias to orient toward the threat, we will exert more top-down
control as we become conscious of the stimulus (Derryberry & Reed, 2002). Once again, the slow system
modulates the fast system.
The balance of activation between the prefrontal cortex and amygdala has been shown to guide visual
attention based on relevance, emotion, and motivation (Gazzaley et al., 2007; Geday, Kupers, & Gjedde,
2007). This is one of the many networks that may become dysregulated in PTSD, resulting in disturbances of
sensory and memory processing (Gilboa et al., 2004; Rauch, Shin, & Phelps, 2006). PTSD patients with
dissociative symptoms have greater activation in neural networks involved in the representation of bodily
states, suggesting a lack of adequate top-down modulation of these networks by the prefrontal cortex (Lanius
et al., 2005). As one would expect, the severity of PTSD symptoms is positively associated with amygdala
activation and negatively correlated with ompfc size and activation (Shin, Rauch, & Pitman, 2006; Williams
et al., 2006).
Research shows that subjects involved in the cognitive appraisal of fearful faces show both a decrease in

161
amygdala activation and an increase in prefrontal activation (Hariri et al., 2000, 2003). This same amygdala-
prefrontal activity shift occurs during activation of a placebo effect (Wager et al., 2004) and recovery following
the presentation of negative emotional material (Jackson et al., 2003). Individuals who manage to control their
fear tend to have more activation in right frontal regions than those who do not (Johanson et al., 1998).
A deficit of extinction learning could be an alternative description of PTSD. Sufferers with PTSD show
amygdala dysinhibition, making them vulnerable to the hallmark symptoms of intrusion and arousal (Akirav
& Maroun, 2007). Patients with PTSD have also been shown to have smaller subregions within the ompfc in
regions where intercalated cells are assumed to reside (Rauch et al., 2003). Also the thickness and activity
levels of these prefrontal regions in patients with PTSD during extinction training correlate with their
symptoms, supporting the association between their symptoms and deficits of cortically based extinction
learning (Milad, Orr, Pitman, & Rauch, 2005; Phelps et al., 2004).

The Recovery of Fears and Phobias Under Stress


Dangers bring fears, and fears more dangers bring.
—Richard Baxter

Jacobs and Nadel (1985) proposed the existence of two systems of learning and memory involved in fears and
phobias. These two systems predicted LeDoux’s model of fast and slow fear circuitry. The taxon system (fast
or amygdaloid system) is responsible for the acquisition of skills and rules, and the conditioning of stimulus-
response connections. This system is context free, meaning that it contains no information about the time or
location in which the learning took place. Taxon learning generalizes broadly and is primarily nonconscious.
This is the system in which early learning of fear, safety, and attachment is organized and stored. The taxon
system is represented in what cognitive psychologists call implicit and procedural memory.
The locale system—with the hippocampus and the cortex at its core—is responsible for cognitive maps,
mental representations and episodic memory. The gradual developmental course of the locale memory system
parallels that of hippocampal-cortical circuits. Thus, although a great deal of social and emotional learning
occurs during the early years of life, no source attribution or autobiographical narrative is associated with these
memories. For example, a mother’s fearful look when strangers approach may cause her child to develop a
general wariness of the world, but not recognize the source of this apprehension in similar situations later in
life.
We enter middle childhood with neural networks programmed by early learning that we experience as
emotional givens. In the absence of trauma, learning in adults involves a balanced integration of taxon and
locale systems that connect sensory, motor, and emotional aspects of memory to its semantic and
autobiographical components. For children and traumatized adults, the taxon system may function
independently, resulting in an adaptive dissociation among various systems of memory and conscious
awareness.
Extreme stress changes the brain’s biological environment in ways that activate the taxon system and
inhibit the locale system. Stress impairs or downgrades the functioning of the locale system, causing us to fall
back on the more primitive organization of taxon (amygdaloid) systems. These changes result in the
emergence of earlier fears or frightening experiences that had been successfully inhibited. This theory certainly
parallels the voluminous research demonstrating the contribution of stress to the emergence or worsening of
psychiatric and physical disorders. From a psychoanalytic perspective, this process may be understood as
regression to more primitive self-states and defense mechanisms. This process also parallels the disinhibition
of neonatal reflexes in patients with dementia or other forms of brain damage.
Despite the apparent extinction of a phobia or fear, the original memory is maintained and can become
reactivated under stress. This neural explanation addresses how the amygdala holds on to past fears and the
Freudian notion of symptom substitution, where one source of fear may be replaced by another. In other
words, a new trigger reactivates the still-intact underlying neural circuitry. Because of this, Jacobs and Nadel
(1985) suggested that the therapist treating fears and phobias may need to continue treatment well after the
symptoms appear to be eliminated and include stress management training. If the overall level of stress can be
decreased, the likelihood of inhibiting primitive fear memories increases.
Successful psychotherapy for anxiety, fears, and phobias has been shaped by the necessity of integrating
fast and slow circuits, and taxon and locale systems and keeping negative amygdala memories inhibited.
Cognitive therapy is all about utilizing the slow locale systems to inhibit and modulate fast taxon systems that
have been shaped in maladaptive ways. Exposure, response prevention, and relaxation training lead to an

162
increasing opportunity for descending inhibition of the amygdaloid circuits by the hippocampal-cortical
networks. This model of memory can be applied to most clinical situations, regardless of the presence of panic
or anxiety disorders.

Drowning in a Sea of Doom


There is no greater hell than to be a prisoner of fear.
—Ben Johnson

Tina’s cardiologist suggested that she see a psychotherapist after a third visit to the emergency room. Each
time, seemingly out of nowhere, Tina would become breathless and lightheaded; her heart would race until
she felt as if it were going to burst from her chest. Convinced she was having a heart attack, Tina would call
the paramedics. As she waited for the ambulance, Tina reported feeling like she was drowning in a “sea of
doom.” She would imagine her teenaged children growing up without her, and vividly recollect her own
mother’s death when she was a child. These feelings and images, together with the fear of death, would make
her even more frightened. She told me that waiting for the ambulance felt like “an eternity.”
Tina, who was actually in excellent health, was repeatedly told she was having panic attacks. It took three
of these embarrassing episodes to convince her to seek therapy. She came to my office feeling defeated and
very frightened, as if she was losing a lifelong battle to stay in control. During our first session, I learned that
Tina had a difficult childhood, including abandonment by her father, prolonged financial difficulties, and the
death of her mother when she was 15. Tina finished high school while living with an aunt, put herself
through college, and became a successful real estate agent. A four-year marriage had left her with two
children, now in their teens, to raise on her own. Tina’s identity was that of a survivor and hard worker who
did not allow herself to depend on others. The panic attacks had shaken her self-confidence and created a fear
of returning to the chaos, pain, and dependency of her childhood. She had hoped for a medical explanation to
avoid revisiting her past.
I began treatment by educating Tina about her body’s fear response and why it felt to her like she was
having a heart attack. Her racing heart, lightheadedness, rapid breathing, and sense of danger were the result
of the amygdala’s multiple signals to prepare to fight or run. Gaining conscious regulation of her amygdala’s
alarm circuitry was the first order of business. We discussed strategies to ward off these attacks by slowing her
breathing and employing relaxation techniques. During sessions, I would have Tina make herself anxious, and
then assist her in calming down. This provided her with a sense of mastery in regulating amygdala activation.
Understanding what was happening in her body and knowing that her life was not in danger relieved some of
her fear.
The second phase of treatment focused on addressing the long-standing lifestyle issues that kept her in a
chronic state of stress. We examined the heavy burden of responsibilities she carried and her lack of relaxation
and recreation. Tina’s financial fears led her to overbook her work schedule to the point of exhaustion. I
learned that Tina constantly criss-crossed the Los Angeles freeway system, traveling between 30,000 and
40,000 miles each year. Between showing homes and shuttling her children from school to their various
activities, we calculated that she was fighting traffic up to six hours a day, usually behind schedule. She began
to understand the panic attacks as her body’s way of telling her to make some changes to reduce her level of
stress. Regular exercise, decreasing her sales territory, and making alternative arrangements for some of her
children’s transportation proved to be the most helpful solutions in these areas.
As these interventions became more routine, we explored the impact of her childhood experiences on both
her self-image and lifestyle. Tina harbored the fear that she would die like her mother, leaving her children
alone in the world. She tried to do everything she could for them, and save all the money they would need to
go to college, all the time thinking that she would not be around much longer. Her financial planning was
detailed and over the years she had followed through with it almost to the letter. The problem was that it had
originally been created for two incomes; now she was doing it on her own. She came to see that her fear of
death might become a self-fulfilling prophecy. Tina also came to realize that her heart was still broken over
her mother’s death, and that she had never allowed herself to grieve her loss, a luxury she had felt she could
not afford. Opening herself to these feelings of loss was the beginning of her therapy.

Summary

163
The fearful brain has two interconnected systems responsible for different aspects of fear processing. The fast
or taxon system—with the amygdala at its core—makes rapid, reflexive, and unconscious decisions to provide
for immediate survival. This system develops first and organizes learning related to attachment and affect
regulation. It involves sensory, motor, and affective memories typical of early life and later traumatic
memories. The slow or locale system, based in hippocampal-cortical networks, contextualizes and makes
conscious what is being processed. The slow system’s job is to regulate the activity of the amygdala by
modulating its output based on a more complex appraisal of potentially dangerous situations. This system
contextualizes experience in time and space, and supports conscious awareness via cortical connectivity.
These two systems, reflecting both top-down and left-right circuits, can become dissociated during
prolonged periods of stress or trauma. In psychotherapy, we attempt to activate fast and slow circuits, taxon
and locale systems, and implicit and explicit forms of memory to inform and educate each about the other.
When emotional taxon networks are inhibited, we use techniques to trigger them so that they can be activated
and integrated with slow locale circuits. When these networks are out of control, we recruit locale circuits to
contextualize them in time and space and allow them to be tamed by the descending, inhibitory capabilities of
cortical processes. The overall goal is the activation and integration of both systems.

164
Chapter 16
Early Traumatic Stress:
The Fragmentation of Self and Other
Our brains have been designed to blur the line between self and other. It is an ancient neural circuitry that
marks mammals, from mouse to elephant.
—Franz de Waal

After a century of research linking early childhood stress to adult psychopathology, we may have finally
uncovered an underlying mechanism of action. Most scientists now think that early experiences are converted
into the neurobiological and biochemical structures of our brains via the transcription of specific sections of
our genetic code through a process called epigenetics (Hoffman & Spengler, 2014). It is through epigenetics
that experience shapes our brains in such a manner as to either enhance or diminish our resilience and
adaptive capacities in the face of subsequent stress. Early deprivation, poverty, dysfunctional households,
physical and sexual abuse, in fact all early experiences impact the infrastructure of our nervous system and our
ability to survive and thrive (Engle & Black, 2008; Felitti et al., 1998).
The diagnostic category of PTSD has historically been understood to be the result of a normally
functioning adult experiencing a traumatic event. This is where a physical or emotional threat triggers a fight-
or-flight response, the threat is dealt with, the flight-fight response subsides, and the brain and body return to
a state of energy conservation and maintenance. Children, on the other hand, are not equipped to cope
autonomously with threats. Their limited mobility, strength, and experience make fighting or fleeing
impossible. Children need to hold tight to their caretakers, who are supposed to provide this survival function
for them. Human caretaking and aggression toward others coevolved to better protect those who cling to us.
All too often, a child is either not afforded adequate protection or experiences trauma at the hands of
caretakers. This results in a cascade of neuroanatomical and neurochemical processes similar to those of
traumatized adults. However, the absence of adequate coping strategies and self-soothing abilities results in
impairments in the development, organization, and integration of the brain, the mind, and attachment
strategies (Perry et al., 1995). When trauma occurs before a sense of self is solidified, the adaptation to the
trauma can distort or even prevent the development of a coherent sense of self.
The agitation shown by neglected and traumatized children can be misdiagnosed as hyperactivity or an
attention deficit, while the numbing response in infants may be misinterpreted as a lack of awareness of
feelings. Until recently, surgery was often performed on infants without anesthesia—their lack of protest and
shock misinterpreted as an insensitivity to pain (Marshall, Stratton, Moore, & Boxerman, 1980; Zeltzer,
Anderson, & Schecter, 1990). Survey research as recently as the 1990s showed that less than 25% of
physicians used anesthesia when performing circumcision on newborns (Hoyle et al., 1983; Wellington &
Rieder, 1993). These practices appear to be a holdover of beliefs that a newborn’s nervous system is too
immature to experience pain. Our lack of appreciation for the possibility of traumatic reactions in neonates
and children can lead us to create unintentional trauma.
Severe neglect may be the most damaging of all forms of early stress. As organs of adaptation, our brains
depend on stimulation to grow. When a child is physically abused, the brain is challenged to adapt, which
stimulates neural growth and connectivity. But with extreme neglect, the resulting neural atrophy spreads
from the cortex down to the striatum, caudate, and cerebellum (Bauer et al., 2009; Hanson et al., 2013a,
2013b; Takiguchi et al., 2015). Victims of severe neglect consistently show ongoing deficits in sensory,
cognitive, and social functioning (Kaler & Freeman, 1994; Pollak et al., 2010).
Although we tend to think in terms of specific clusters of symptoms within diagnostic categories, early
trauma can result in a spectrum of highly diverse symptom patterns and an array of overlapping
neurobiological and behavioral alterations to normal development. In turn, these alternations impact how
children experience and cope with their worlds—how they do in school, how they interact with their parents,
and whom they choose as friends. Some of the effects of early stress impact specific structures such as the

165
amygdala and hippocampus while others effect the development of the entire brain. For example, early stress
results in diminished organization of left-hemisphere auditory processing and decreased gray matter in the
visual cortex suggesting that even our sensory systems are shaped by negative emotional experience—perhaps
to protect us from overwhelming stress (Choi et al., 2009; Shimada et al., 2015; Tomoda et al., 2009;
Tomoda, Polcari, Anderson, & Teicher, et al., 2012).
The connection between borderline personality disorder (BPD) symptomatology and childhood abuse has
long been recognized (Sansone, Sansone, & Gaither, 2004; Wingenfeld et al., 2011). Adults with BPD tend
to report higher levels of childhood trauma and lower levels of parental care than those with other psychiatric
diagnoses (Machizawa-Summers, 2007). Borderline outpatients in Australia were found to have a high rate of
physical and emotional abuse in childhood (Watson, Chilton, Fairchild, & Whewell, 2006). A large sample of
college students in Turkey found a highly significant correlation between BPD and childhood trauma (Sar et
al., 2006). Borderline personality disorder is characterized by a variety of disturbances of the experience of self
and the ability to maintain social relatedness.
The histories and symptoms of these individuals strongly suggest that early attachment was experienced as
highly traumatic and even life-threatening (Fonagy, Target, & Gergely, 2000). In line with these experiences,
the world is seen as more malevolent and themselves as less fortunate than others (Giesen-Bloo & Arntz,
2005). The trauma resulting from interpersonal betrayal appears to be particularly salient in their minds
(Kaehler & Freyd, 2009, 2011). Whether these perceptions are a result of negative experiences, a genetic bias,
or some combination of the two is unknown (Judd, 2005). It is interesting to note that all of the clients with
borderline symptomatology that I have worked with have had siblings who did not demonstrate the same
symptoms, even though they shared parents and family environments.
Attachment trauma can result from physical or sexual abuse, neglect, or profound misattunement between
parent and child. Affective disorders have also been shown to occur at above-average rates in these patients
and their parents, a likely contributing factor to difficulties in emotional regulation. Whatever the cause, the
child is unable to utilize others in the development of secure attachment and to regulate overwhelming anxiety
and fear. The result is that real or imagined abandonment triggers a state of terror, similar to what children
experience when separated from their mothers.

TABLE 16.1
Neurodevelopmental Effects of Various Forms of Early Trauma

Neurobiological Finding Stressor Developmental Deficits


RAD, BPD, PTSD Memory
Decreased hippocampal gray matter volume1
Maltreatment Learning
Poverty Reality testing
Preterm birth Emotional regulation
RAD Visual acuity and detail
Decreased visual cortex gray matter volume2
Physical abuse Social information processing
Domestic violence
RAD, BPD Executive functioning
Decreased frontal gray matter volume3
Corporal punishment Affect regulation
Maltreatment Problem-solving
Preterm birth Learning
Gray matter abnormalities in amygdala volume4 RAD, BPD Affect regulation
Neglect Poor approach/avoidance choices
Institutionalization Enhanced or diminished fear activation
Poverty Academic success
Decreased cortical gray matter5
Physical abuse Problem solving
Maltreatment Judgment
Neglect
Institutionalization
Maltreatment Neural network integration
White matter abnormalities6
Neglect Academic success
Preterm birth
BPD Bonding and attachment
Decreased gray matter volume in cingulate cortex7
Corporal punishment Affect regulation
Sensory processing
BPD, PTSD, RAD Learning Immunological Activity
Cortisol dysregulation8
Family instability Neuroanatomical growth + integration
Adoptees

RAD—reactive attachment disorder; BPD—borderline personality disorder;

166
PTSD—posttraumatic stress disorder

Structural similarities exist in the neurobiological impacts of early stress across diagnostic categories and
different groups of victims of early stressful experiences. Table 16.1 contains a sample of the current findings
that connect early stress neurodevelopmental abnormalities and their consequences.
Based on these overlapping results, we will be focusing on individuals with complex posttraumatic stress
disorder (C-PTSD), reactive attachment disorder (RAD) and BPD as well as others who experience early
abuse and neglect. The individuals in all three diagnostic groups experienced some combination of attachment
trauma, neglect, excessive corporal punishment, domestic violence, or maltreatment. A hallmark of these
diagnoses is emotional dysregulation, difficulty with establishing and maintaining relationships, and a
fragmented sense of self. This trinity of deficits appears to be grounded in the interwoven neural architecture
and early development of affect regulation, attachment, and the experience of self.

Complex PTSD
The most satisfying thing in life is to have been able to give a large part of one’s self to others.
—Pierre Teilhard de Chardin

C-PTSD or developmental trauma disorder occurs as a result of early, prolonged, and inescapable stress from
some combination of emotional abuse, physical maltreatment, neglect, and lack of nurturing care. Although
not yet in our diagnostic manual, it is a cluster of symptomatology with clinical utility. It is differentiated from
typical PTSD because of its early origins, extensive effects on all areas of development, and persistence
throughout life (Herman, 1992; Navalta et al., 2004). C-PTSD correlates with a range of neurological
abnormalities, enduring personality traits, and coping strategies that decrease positive adaptation and increase
vulnerability to future trauma (Green, 1981; Gurvits et al., 2000). In other words, what doesn’t kill us makes
us weaker.
Although extreme or chronic stress can damage the brain and body at any point in life, it is particularly
toxic to the very young. Research has consistently shown that the more severe, longer, and earlier the exposure
to neglect and/or trauma, the greater the negative consequences. We have also learned that the impact of the
stressor cannot be judged through an adult’s eyes, but must be considered in light of the developmental stage
and disposition of the child. For example, a two-week parental vacation when a child is three will be
devastating to some children but hardly a blip on the emotional radar for others. The reasons for this have to
do with a wide variety of genetic, epigenetic, and situational variables. Although domestic violence, divorce, or
corporal punishment will have different impacts depending on the child, their role as potentially life-altering
stressors should not be minimized.
The personalities of children with C-PTSD develop in the shadow of trauma and they appear to never
make the developmental leap from amygdala to cortical executive functioning. Their lives are lived in a
survival-based, amygdala-centric manner, and they are less connected to the consensual realities of the group
mind. They are, in essence, constantly vulnerable to takeover by a more primitive version of the modern
human brain—yet are tasked to live among others whose brains are organized in a different way. Day-to-day
life is stressful, threatening, and difficult to navigate.
Don’t simply check off the boxes in an intake report “all developmental milestones reported to have been
met on time.” Find out about the social and emotional environment in which they developed—the mother’s
postpartum depression, the grandmother’s violent death, the father’s unemployment. Factors like these,
critical to psychological development, if missed, can lead to years of confusion in therapy. In secure
attachment, the child is able to use the parent as a safe haven and to avoid experiencing sympathetic nervous
system activation in response to stress. This is why the quality of early relationships serves as a powerful
modulating factor in the experience of early stress.
Dissociation allows the traumatized individual to escape the trauma via a number of biological and
psychological processes. Derealization and depersonalization reactions allow the victim to avoid confronting
the reality of his or her situation or watch it from an imaginal distance as a detached and safe observer.
Compulsive defenses related to eating and gambling, as well as somatization disorders (in which emotions are
converted into physical symptoms) all reflect complex adaptations to early deprivation and trauma. While
these strategies allow us to survive childhood, they may later come to handicap our abilities to adapt to a
nontraumatizing world.

167
Reactive Attachment Disorder
Loneliness comes with life.
—Whitney Houston

Political and economic conditions in 1980s Romania resulted in tens of thousands of orphans being
warehoused with the disabled and mentally ill. They received little attention, supervision, or medical care,
while being subject to neglect, physical and sexual abuse, and sedative medication to control their behavior
(Wilson, 2003; Zeanah et al., 2004). The exposure of these practices led to the adoption of a significant
number of these orphans by families from around the world.
These Romanian orphans were generally small for their age, had difficulties in concentration, attention,
and attachment, and suffered with symptoms usually associated with psychosis, autism, and PTSD. The term
institutional autism, sometimes used to describe them, was based on their poor language abilities, deficient
memories, and self-stimulatory behavior (Federici, 1998; Minnis et al., 2013). These early experiences had
also clearly impacted the development of cortical-amygdala circuitry that organized both emotional regulation
and attachment. The behavioral, emotional, and cognitive after effects of their experiences became the initial
impetus for the diagnosis of RAD (Kay & Green, 2012).
A characteristic feature of RAD is the victim’s ways of relating to others. Some of the children failed to
initiate or respond to social interactions at all while others would indiscriminately approach strangers—a
behavior also noted in some children raised in foster care (Pears, Bruce, Fisher, & Kim, 2009). An early
animal model for RAD was created by Harry Harlow, who raised young rhesus monkeys in isolation. When
placed with a troop of other monkeys for the first time at six months, they were terrified, curled into the fetal
position, stayed in a corner, or did whatever they could to stay away from those around them. Because
cortical-amygdala circuitry hadn’t developed in these isolate monkeys, the other monkeys were likely
experienced as incomprehensible and frightening creatures from which they quickly retreated.
The amygdala’s role as an organ of appraisal—what’s good and bad, what we should approach and avoid—
makes the amygdala a key structure of the social brain. The amygdala has been known to serve as a “social
brake” in primates that keeps them from approaching strangers or familiars of higher rank who might attack
them. When they experience amygdala damage, primates will be indiscriminate in their approach, not only to
inappropriate others, but to natural predators like snakes (Amaral, 2002). Early stress and deprivation
powerfully impact the development and connectivity of the amygdala and basic survival skills it serves
(Chisholm, 1998; Lyons-Ruth et al., 2009).
Research with institutionalized children has demonstrated reduced amygdala discrimination between
mothers and strangers (as well as indiscriminate friendliness), assuring us that the social role of the amygdala
has been conserved during primate evolution (Olsavsky et al., 2013). Other studies have demonstrated
differences in amygdala volume when compared to matched children who had not experienced serious early
stressors (Hanson et al., 2015; Mehta et al., 2009; Tottenham et al., 2010, 2011). Given that attachment
behavior has been shaped by evolution to optimize survival via safe approach-avoidance choices, the
indiscriminate maturation and connectivity of their approach to strangers seen in the children is a clear signal
that something has gone terribly wrong in the amygdala of these children.
The heartening news from follow-up studies with these Romanian orphans is that many showed
considerable “catch-up” when placed in healthy and nurturing families. This catch-up demonstrates the power
of plasticity, the generosity of adoptive families, and the drive to survive and thrive (Rutter et al., 1997, 2007).
We must remember, however, that early emotional neglect correlates with later illnesses and that these
children may remain vulnerable later in life for both physical and psychological difficulties (Fantuzzo et al.,
2005; Felitti et al., 1998; McLaughlin et al., 2015; Sánchez et al., 2005; Wilson et al., 2012).

Borderline Personality Disorder


We are all so much together, but we are all dying of loneliness.
—Albert Schweitzer

Although there are a number of theories as to its cause, the etiology of BPD most likely stems from some
combination of early deficits in emotional regulation and problematic attachment relationships. This idea is
supported by the frequent occurrence of early abuse, trauma, and the presence of dissociative symptoms in
these individuals. Their reported histories and symptoms suggest that early attachment was experienced as

168
traumatic, threatening, and possibly life threatening. Clients with this diagnosis are characterized by:

1. Hypersensitivity to real or imagined abandonment


2. Disturbances of self-identity
3. Intense and unstable relationships
4. Alternating idealization and devaluation of themselves and others
5. Compulsive, risky, and sometimes self-destructive behaviors

Victims of BPD provide us with a window into the intense and chaotic experience of an emotionally
dysregulated infancy. High levels of amygdala arousal made worse by underdeveloped and hypoactive frontal
and hippocampal circuits leave victims at the whim of unpredictable and overwhelming emotions. The
resultant chronic anxiety causes them to create problematic and chaotic relationships that not only increase
their stress but also result in an endless series of abandonments.
They experience criticism, shame, and abandonment from all directions. If confronted with even a hint of
criticism or rejection, they can become emotionally overwhelmed and catapulted into attacking others or
harming themselves. Friends and family become targets of rage, are accused of sadistically causing pain, and
are bewildered by sudden shifts of mood and unpredictable behavior.
As with those suffering the impact of C-PTSD and RAD, what we witness in the lives of those with BPD
appears to be the result of the disruption of the development and integration of social brain and other neural
systems that control emotional regulation, executive functioning, interpersonal experience, and a coherent
sense of self. It is most likely that these neurodevelopmental abnormalities are the result of an interaction
between innate genetic variables and life experiences (Distel et al., 2011). There is evidence in these
individuals of altered patterns of brain maturation even during childhood (Houston, Ceballos, Hesselbrock, &
Bauer, 2005).
Traditional neuropsychological testing demonstrates deficits in executive function, attention, memory, and
a variety of cognitive processes similar to those of medical patients who sustain frontal and temporal lobe
damage (Coolidge, Segal, Stewart, & Eliot, 2000; Dinn et al., 2004; Paris et al., 1999; Posner et al., 2002;
Swirsky-Sacchetti et al., 1993; van Reekum et al., 1993). Scan studies have further shown abnormalities in the
size, activation patterns, and neurochemical levels in several brain regions (Cowdry, Pickar, & Davies, 1985;
Johnson et al., 2003; Lange et al., 2005). These patients have reduced gray matter in their hippocampi,
amygdala, left orbital medial and right anterior cingulate cortices, and a variety of other neuroanatomical
abnormalities (Bremner et al., 2000; Chanen et al., 2008; Johnson et al., 2003; Takahashi et al., 2009).
As we might expect, the ways in which their brains differ from others are found in networks involved in
affect regulation and the experience of self and others (Bazanis et al., 2002; Johnson et al., 2003). These
findings may reflect the experience-dependent nature of the development of our executive networks and how
powerful the impact of early stress can be to derail their optimal development. An alternative hypothesis is
that individuals who become borderline may have a general bias resulting in neurodevelopmental
abnormalities that makes normal affect regulation and attachment impossible.
At rest, the brains of individuals with borderline symptoms demonstrate lower metabolic activity in
prefrontal, cingulate, and hippocampal regions (Díaz-Marsá et al., 2011; Juengling et al., 2003; Soloff et al.,
2003). Interestingly, higher levels of neurotoxins have also been found in the frontal lobes of these patients
(Tebartz van Elst et al., 2001). Although these individuals are under-aroused on measures of heart rate, skin
conductance, and pain sensitivity (Bohus et al., 2000; de la Fuente et al., 1997; Goyer et al., 1994; Herpertz,
Kunert, Schwenger, & Sass, 1999), they show greater activation in the amygdala, prefrontal cortex, temporal
and occipital lobes, and the fusiform gyrus when shown slides of emotionally adverse situations (Herpertz et
al., 2001; Johnson et al., 2003; Juengling et al., 2003).
Of special note are deficits of visual processing and visual memory, which may account for the high
comorbidity of BPD and body dysmorphic disorder (Beblo et al., 2006; Harris, Dinn, & Marcinkiewicz,
2002; Semiz et al., 2008). Clients with borderline symptoms demonstrate reduced metabolism in the right
occipital lobe when evaluating emotional expressions (Merkl et al., 2010). This may relate to why they tend to
misinterpret or exaggerate neutral and negative expressions and are more likely to see disgust and surprise in
the faces of others while underrecognizing fear (Unoka, Fogd, Füzy, & Csukly, 2011). Amygdala activation
and its connections to primary visual areas may negatively distort the perception of faces and activate higher
levels of anxiety and fear when one is confronted with negative stimuli (Koenigsberg et al., 2009).
BPD and PTSD patients have greater left amygdala activation in response to facial expressions, causing
them to experience neutral expressions as negative or threatening (Donegan et al., 2003; Umiltà et al., 2013).
Thus, there is a tendency to both negatively distort neutral information and take it personally, which likely

169
drives their antagonism, suspiciousness, and assaultiveness (King-Casas et al., 2008; Minzenberg, Poole, &
Vinogradov, 2006). They are particularly hypervigilant for and sensitive to overt and subliminal negative social
cues (Sieswerda, Arntz, Mertens, & Vertommen, 2006) and show larger startle responses to unpleasant words
(Hazlett et al., 2007), making cooperation and social problem solving exceedingly difficult (Dixon-Gordon,
Chapman, Lovasz, & Walters, 2011; Maurex et al., 2010).
Borderlines are highly sensitive to any stress (especially rejection), both their own and the stress or social
exclusion they see happening to others (Buchheim et al., 2008; Minzenberg et al., 2008; Ruocco et al., 2010).
Individuals with BPD traits have also been found to be more vulnerable to developing PTSD in adulthood in
situations of domestic violence (Kuijpers et al., 2011). Memories of abandonment are associated with
increased bilateral dlpfc activation and decreases in right anterior cingulate cortex activation, suggesting that
they tend to move away from self-awareness and empathy and are more likely to react to stress by taking
action (Schmahl et al., 2003). Individuals with these symptoms will go far out of their way to punish those
they feel have wronged them and are especially likely to devote themselves to causes like animal rescue where
they can fight injustice.
Activation of the amygdala is at the heart of emotional and traumatic memory, inhibiting the neural
networks (frontal-hippocampal) that could contextualize and attenuate them (Cahill & McGaugh, 1998;
McGaugh, 1990). When in therapy, emotional interpretations can trigger a catastrophic reaction akin to
separation and abandonment. The most salient indication of being with a borderline individual is feeling
attacked, inadequate, and in emotional danger. This is likely exactly what they feel, and they are very adept at
creating that feeling in others. These reactions and our reaction to them may be our best window into the
chaotic emotional world of their early childhood.
Despite their vulnerability to the pain of others, they exhibit deficits in both the emotional and cognitive
aspects of empathy with abnormal functioning in the right insula and left superior temporal sulcus during
empathy tasks (Dziobek et al., 2011). Being empathic requires the ability to put your own perspective and
agenda aside, a task that requires both emotional regulation and executive functioning that they find
exceedingly difficult. Like people with depression, borderline clients appear to have deficits in switching their
state of mind from self to other-relevance, which may be why they are unable to convert their high level of
emotional attunement with others into empathy for them (Schnell et al., 2007).
Their brains are put on high alert for danger, and misjudge and distort incoming information, while
simultaneously decreasing inhibition, reality testing, and emotional control. When clients with borderline
symptoms experience negative feelings, they are overwhelmed and unable to use the conscious cortical
processing needed to test the appropriateness of their reactions or to solve whatever problems they are facing.
They lose perspective, the ability to remember ever feeling good, or the idea that they may ever feel good
again. Overwhelming fear and lack of perspective combine to create the experience that their very life is at
risk. This vulnerability to overwhelming affect is shared by many victims of early stress, trauma, and neglect.

Self-Fragmentation
Friendship with one’s self is all important, because without it one cannot be friends with anyone else in the world.
—Eleanor Roosevelt

Our subjective sense of self depends on a combination of somatic, psychological, and social processes that
create a feeling of continuity and cohesion to our experience through time. Together, an array of conscious
and unconscious processes combine to knit together a sense of self. While our breathing and heartbeat create a
background rhythm, or internal organs bind us to our bodies. As our relationships remind us of our social
identities, our personal and cultural narratives ground us with blueprints for action while providing us a role in
group history.
An instinct for self-preservation and the urge to protect ourselves from physical and social harm may also
aid in the construction of a sense of self. The act of self-care and self-preservation serves as a referent to our
sense of having a self to protect. The universality of self-preservation helps us to appreciate the power of self-
harm and suicide. Something fundamental to our biological and social programming has to be undermined for
self-harm to occur.
Adults who engage in repeated self-harm almost always describe childhoods that include abuse, neglect,
cruel teasing, and shaming at the hands of caretakers (Mazza & Reynolds, 1998; Pfeffer et al., 1997; Zoroglu
et al., 2003). Suicide attempts, especially those with BPD, are almost always related to shame, abandonment,
and loss (Brodsky et al., 2006). Repeated self-harm and suicide attempts also appear to be reinforced by the

170
rapid attention of health care professionals, family, and friends (Schwartz, 1979). It is logical therefore to
consider that the pain of abandonment and ostracism has the power to overwhelm our primitive instinct for
physical self-preservation.
Endorphins are naturally occurring morphine-like substances that (along with oxytocin and serotonin)
provide us with the sense of well-being associated with secure attachment. There may also be an important
connection between abandonment and self-harm that is mediated by endorphins. In case of injury or
prolonged physical exertion, the body releases endorphins to provide analgesia for pain to allow us to continue
to fight or flee (Villalba & Harrington, 2003). The release of endorphins in response to fear also helps us to
inhibit anxiety and enhance coping with stress (Fanselow, 1986; Kirmayer & Carroll, 1987). Self-injurious
(nonsuicidal) patients have been found to have both problematic attachment histories and lower levels of
endorphins (New & Stanley, 2010; Prossin et al., 2010; Stanley et al., 2010). This implies that our available
levels of endorphins may be related to the quality and nature of early experiences.
The analgesic effects of these endorphins may account for the reports of reduced anxiety and sense of calm
after self-injury. This may be a way to activate endorphins and achieve the biological feelings associated with
the safety of secure attachment. This theory is supported by the fact that self-harm decreases or stops
completely when patients are given a drug that blocks the effects of endorphins (Pitman et al., 1990; van der
Kolk, 1988). Thus, the drive to self-harm and suicidal gestures triggered by abandonment may be reinforced
by the biochemistry of secure attachment.
Repeated self-harm and suicidal gestures are common symptoms of BPD. A fundamental emotional
reality for these individuals is a physical sense of self-disgust and feelings of being defective and unworthy of
love. This persistent emotional truth is held tightly in spite of being accomplished, exceptionally
conscientious, and an all-around good citizen. Yet, despite the lack of objective validation for feelings of self-
disgust, to become self-aware is to feel repulsive and repulsed by one’s self. Having a feeling of disgust when
becoming self-aware would certainly make reflection a painful experience. It is easy to imagine that these
feelings would undermine self-esteem and secure attachment and trigger thoughts of self-destruction.
Disgust is an extremely primitive emotion designed to make us reflexively retreat from potential danger
such as rotting food or decaying flesh. The possibility exists that the early experiences of traumatized and
neglected individuals could lead them to associate the experience of self with disgust. Caretakers who either
push them away or look at them with disgust may be internalized as “inner objects.” Interestingly, one study
found that women with BPD or PTSD had higher levels of “disgust-sensitivity” and were “disgust prone”
when discussing their self-image (Rüsch et al., 2011). Another found that when women with BPD recalled
memories of childhood, they tended to show disgusted facial expressions (Buchheim et al., 2007).
Disgust may become part of one’s self-image through a neurobiological pathway. Organized like a map of
the body, the insula and the anterior cingulate integrate primitive bodily states with the experience and
expression of behavior, cognition, and emotion from love to disgust (Bartels & Zeki, 2000; Calder et al.,
2003; Carr et al., 2003; Phan et al., 2002). The insula is also central to our sense of our body in space and our
ability to distinguish between self and others (Bechara & Naqvi, 2004; Critchley et al., 2004; Farrer & Frith,
2002; Gundel et al., 2004). Both the insula and anterior cingulate become activated when people are asked to
recall behavior for which they felt ashamed (Shin et al., 2000). To show just how closely bodily states are tied
to cognition and reasoning, giving subjects something sweet or bitter just before asking them to make a moral
judgment will influence their decision in a positive or negative direction, respectively (Eskine, Kacinik, &
Prinz, 2011).
Given its functions and early development, the insula and anterior cingulate may be responsible for the
basic associations between a conscious sense of one’s bodily states and emotions. For example, in the context
of secure attachment, the insula may associate feelings of love and physical well-being with self-awareness. On
the other hand, experience of neglect, abuse, or seeing disgust in the eyes of caretakers may become
neurologically linked to our emergent self-identity. This could serve as a neurobiological substrate for both
self-disgust and core shame, making self-awareness something to be avoided at all costs.
The ability to attune to the feelings of others is central to both intimate relationships and our ability to
link to the group mind. The numbing, social withdrawal, and “self-imposed” isolation so common in those
suffering with PTSD reflect the disruption of fundamental social abilities (Galovski & Lyons, 2004;
Nietlisbach & Maercker, 2009; Ray & Vanstone, 2009; Riggs, Byrne, Weathers, & Litz, 1998).

Social Fragmentation
What loneliness is more lonely than distrust?

171
—George Eliot

A central feature of many victims of early trauma and stress is an inability to attain and sustain interpersonal
relationships (McFarlane & Bookless, 2010). There appears to be a complex set of neurodevelopmental
deficits that results in a disconnection from the group mind. This phenomenon is most well documented in
borderline clients but also true for many individuals suffering with RAD, PTSD, and C-PTSD. The
fragmentation of self—shame, self-disgust, self-injurious behavior, and suicide attempts—is interwoven with
problems with intimacy and attachment. A central question, however, is whether the fragmentation of social
relationships is a simple by-product of other symptoms or a separate consequence of trauma with its own
neurological mechanisms.
In describing the brain’s executive functioning, we discussed two different cortical systems: the first was
the frontal-parietal network that attends to monitoring and performing tasks in the outside world, and the
second is the default-mode network (DMN) that becomes activated when our attention is turned to
considering the self and relating to others. These systems are generally anticorrelational, meaning that when
one is active, the other is inhibited. You may also remember that there is a salience network centered in the
insula and anterior cingulate cortices that signals the frontal-parietal system to become active in situations of
novelty and danger, which also serves to inhibit the DMN (Bruce et al., 2013; Thome et al., 2014; Zhang et
al., 2015). Given these newer neurobiological insights, can we make better sense of the fragmentation of self
and social relationships we witness in those who experience early stress and extreme trauma?
It is clear that the structures and neurochemistry of the brain are impacted by trauma in ways that bias it
toward sustained arousal, hypervigilance, and chronic anxiety. This means that the brain’s primitive executive
network centered in the amygdala serves to activate the salience system and inhibit the DMN (Palaniyappan,
2012; Patel et al., 2012; Rabinak et al., 2011; Seeley et al., 2007; Shin et al., 2008; Sridharan et al., 2008).
From an early age, victims of trauma and stress will be more stressed both at rest and when challenged, have
overactive salience networks, more active frontal-parietal networks, and inhibited DMNs (Lieb et al., 2004;
Lyons-Ruth et al., 2011).
If the DMN is chronically inhibited during early development or after severe trauma during adulthood, the
victim’s ability to develop a coherent sense of self, regulate affect, be introspective, or have empathy for others
will be significantly impaired. When stress and trauma occur from early childhood, a sense of having a safe
inner world to withdraw to for comfort may never be achieved. Looking at the many functions of the DMN—
from having a sense of self, to reading social cues, to imagining a future—it is certainly possible that chronic
inhibition of this important network can drastically change the experience of being human. This theory is
supported by findings that those with PTSD and BPD exhibit deficits in empathy, self-referential processing,
autobiographical memory, and theory of mind, all of which involve the DMN (Frewen et al., 2011;
Nietlisbach & Maercker, 2009).
At rest, individuals suffering with BPD showed hypometabolism in the frontal lobes and hypermetabolism
in the motor cortex and anterior cingulate (Salavert et al., 2011). They also show decreased cortical and
fronto-limbic connectivity, and deficits in cortical inhibition (Barnow et al., 2009; Cullen et al., 2011; Wolf et
al., 2011; Zhou et al., 2012). Deficits in metabolism are also seen in the precuneus parietal lobes and posterior
cingulate regions, components of the DMN that contribute to the inner experience of self and
autobiographical memory (Lange et al., 2005). All in all, this pattern of functional activation makes victims
less able to stop, reflect, and evaluate their situation while more likely to act, react, and be guided by their
primitive fear and other emotions.
It has been demonstrated that those suffering with BPD have difficulty reflecting on their own thoughts,
especially when they are under stress. That is, they lose the ability to think about their own thinking, making
metacognition and self-monitoring across emotional states highly problematic. Decompensation in the face of
interpretations, like the reaction to real or imagined abandonment, reflects a rapid shift from frontal to
subcortical (amygdala) dominance, manifesting in an emotional storm and resulting in functional regression.
When the amygdala hijacks executive control of the brain during states of emotional arousal, the reflective self
disappears.
Together, these data suggest that even in the absence of stimuli, victims of early attachment trauma may
find it difficult or impossible to find a safe place within themselves. In fact, trying to relax, meditate, or taking
a yoga class may overwhelm them with internal thoughts and painful feelings that are too difficult to contain.
The combination of anxiety and distorted social information drives them to obsess about potential
abandonment (Sharp et al., 2011). Therefore, while meditation and self-reflection may be a positive means of
coping for many, these individuals need to distract themselves from self-awareness in order to feel calm.
The central focus in the treatment of victims of early trauma is to modulate arousal in ways that are well
matched to the individual. Having a therapist and treatment context that serve as “amygdala whisperers” is the

172
first step. When working with clients with borderline symptoms, structure and limit setting combined with
flexibility and patience is key. The therapist must provide an external executive scaffolding within which the
client can build brain networks of memory, self-organization, and affect regulation that were impossible earlier
in life. In this way, the therapist serves as an external neural circuit to aid in the integration of networks left
underdeveloped and dissociated during development.
All in all, it appears to me that these new scientific findings suggest that the fragmentation of self and
others should be considered as a separate symptom cluster across the diagnoses discussed in this chapter. Even
though the inhibition of the DMN is secondary to high levels of arousal, I think the distortion of self-
experience and the ability to attune to and connect with others does not simply dissolve when anxiety
diminishes. The consequences of prolonged periods of DMN inhibition, especially when experienced early in
life, have permanent effects on the DMN network and a wide range of functions that require separate and
specific interventions.

Summary

We need to expand our notion of trauma from the catastrophic events of adulthood to the everyday
interactions during childhood that we depend on for our survival. Most of our learning and adaptation is not
traumatic but rather subtle and mostly unconscious. The interactions between parent and child, the politics of
the schoolyard, and experiences of small victories and defeats all contribute to shaping who we become. We
need to keep in mind that, as primates, attachment equals survival and abandonment equals death. This may
help us appreciate the power of parental abuse and abandonment to shape children for the rest of their lives.
Early stress in the form of neglect, abandonment, and abuse has a profound effect on many aspects of
brain development. Those experience-dependent networks that allow us to connect with others and develop a
positive sense of self-identity appear to be particularly impacted. Across the diagnostic categories of RAD, C-
PTSD, and BPD, we witness the lifelong impact on the victim’s ability to regulate emotions, attach to others,
and function in the world in a consistent and constructive manner.
With the recent discovery of the DMN, we now have a model for a neural network that may be central to
the organization of self and other. This provides us with the opportunity to explore the construction and
regulatory functions of subjective experience in a way previously unavailable to us. The DMN and the new
technologies available to us may make it possible to integrate meditative and yogic traditions and techniques
into the mainstream of psychotherapeutic treatment.

173
Chapter 17
The Impact of Trauma:
Biochemical Dysregulation and
Neural Network Dissociation
The beauty of the world has two edges, one of laughter, one of anguish, cutting the heart asunder.
—Virginia Woolf

For each of us, there is a point at which fear crosses the line into trauma, resulting in a fragmentation of
sensory, emotional, and cognitive processing. The long-term psychological and neurobiological consequences
of trauma lie on a broad continuum of severity. As a general rule, the earlier, more severe, and more prolonged
the trauma, the more negative and far reaching its effects (De Bellis, Baum, et al., 1999; De Bellis, Keshavan,
et al., 1999). Over time, unresolved trauma can result in symptoms of PTSD.

The Symptoms of Posttraumatic Stress Disorder


The best way out is always through.
—Robert Frost

Traumatic experiences result in a variety of well-understood physiological and psychological reactions to


threat, which trigger the expression of a number of predictable symptoms. These symptoms tend to gradually
diminish after the resolution of the traumatic situation as we gather support from others and repeatedly share
the experience. These healing conditions allow us to gradually regain neurobiological homeostasis and
psychological integration.
Constructing narratives in an emotionally supportive environment supports the psychological and
neurobiological integration required to avoid dissociative reactions. Talking through the trauma with
supportive others creates the neurobiological conditions needed to reestablish neural coherence. Narratives
drive the integration of cognition, affect, sensation, and behaviors, which can remain dissociated, especially
when early trauma, such as childhood sexual abuse, is never discussed. The suffering of Holocaust survivors
and combat veterans is often exacerbated by the political and social dynamics that encourage individuals to
remain silent about their horrifying experiences.
When trauma is severe or chronic, individuals may develop PTSD, which is caused by the dysregulation of
the neurobiological processes responsible for appraising and responding to threats. When this system becomes
dysregulated, the body reacts as if the past trauma is continuing to occur. The three main symptom clusters of
PTSD—hyperarousal, intrusion, and avoidance—reflect de-integration between neural networks that control
cognition, sensation, affect, and behavior.
Hyperarousal reflects a stress-induced dysregulation of the amygdala and autonomic nervous system that
result in an exaggerated startle reflex, agitation, anxiety, and irritability. Chronic hyperarousal leads an
individual to experience the world as a more dangerous and hostile place. Constant agitation and wariness
make us less desirable companions, and these symptoms often cut us off from the healing effects of
relationships.
Intrusion symptoms occur when traumatic experiences break into conscious awareness and are experienced
as if they are happening in the present moment. There is no sense of distance from the trauma in time or
place because the cortico-hippocampal networks have not been able to integrate and contextualize the
somatic, sensory, and emotional memories within autobiographical memory. Intrusions may manifest as
flashbacks, resulting in a veteran hitting the ground in response to a car backfiring or a rape victim having a
panic attack while making love to her husband. These are activations of subcortical systems cued by stimuli

174
reminiscent of the trauma. You may remember from Chapter 5 on memory and Chapter 7 on fear that the
amygdala both controls this activation and tends to generalize from the initial stimuli to a wide variety of cues.
Overgeneralization causes individuals to be triggered by an ever-increasing number of minor or
nonthreatening stimuli.
Avoidance is the attempt to defend against perceived danger by limiting contact with the world,
withdrawing from others, and narrowing the range of consciously experienced thoughts and feelings.
Avoidance can take the form of denial and repression and, in more extreme instances, dissociation and
amnesia. The power of avoidance was highlighted by the research of Williams (1994), who found that 38% of
adult women who had suffered documented sexual abuse as children had no memory of the event. Compulsive
activities can also aid in the avoidance of negative affect, as can alcohol and drug abuse, both common in
victims of trauma. Avoidance enables short-term anxiety reduction while maintaining the lack of neural
network integration that perpetuates the illness.
When experienced in combination, these symptoms result in a cycle of activating and numbing, reflecting
the body’s memory for, and continued victimization by, the trauma (van der Kolk, 1994). Instead of serving to
mobilize the body to deal with new external threats, traumatic memories trigger continuous frightening
emotional responses. Someone suffering from PTSD becomes caught in a loop of unconscious self-
traumatization and fight/flight reactions. When these symptoms are experienced on a chronic basis, they can
devastate every aspect of the person’s life, from physical well-being to the quality of interpersonal
relationships, to the individual’s experience of the world.
We have all heard the sayings “What doesn’t kill you makes you stronger” and “Time heals all wounds.”
These bits of common wisdom conjure up pictures of difficult and traumatic experiences that, once overcome,
result in greater levels of physical and emotional well-being. Although trials and tribulations can certainly
build character, they can also permanently compromise biological, neurological, and psychological
functioning. Trauma produces widespread homeostatic dysregulation that interfere with all realms of personal,
interpersonal, and physical functioning (Perry et al., 1995; Winning et al., 2015).
Support for the idea that trauma has a pervasive impact comes from research showing that cumulative
lifetime trauma increases the likelihood of developing PTSD (Yehuda et al., 1995). A history of previous
assaults also increases the chances of developing PTSD following rape (Resnick, Yehuda, Pitman, & Foy,
1995). Likewise, childhood abuse victims are more likely to develop PTSD after adult combat exposure
(Bremner, Southwick, et al., 1993). It has also been found that severe stress reactions during combat make
subsequent negative reactions to mild or moderate stressors more probable (Solomon, 1990). All evidence for
the fact that what doesn’t kill you makes you more vulnerable to future stressors.

The Neurochemistry of PTSD


Gulf War Syndrome is not one cause, not one illness. It is many causes, many illnesses.
—Christopher Shays

States of acute stress result in predictable patterns of biochemical changes, including increases in
norepinephrine (NE), dopamine, endorphins, and glucocorticoids and a decrease in serotonin. These changes
are part of the body’s mobilization in the face of threat. When stress is prolonged or when it becomes chronic,
changes continue to occur in the baseline production, availability, and homeostatic regulation of these
neurochemicals, altering long-term behavioral and psychological functioning. Each of these substances has its
own role in the stress response and contributes in different ways to the long-term impact of PTSD.
As we have seen, increased levels of NE prepare us for fight-or-flight readiness and reinforce the biological
encoding of traumatic memory. Chronically high levels of NE result in increased arousal, anxiety, irritability,
and a heightened or unmodulated startle response (Butler et al., 1990; Ornitz & Pynoos, 1989). Besides being
stronger, the startle response is also more resistant to habituation in response to subsequent milder and novel
stressors (Nisenbaum, Zigmond, Sved, & Abercrombie, 1991; Petty et al., 1994; van der Kolk, 1994). Being
consistently startled increases the victim’s experience of the world as an unsettling and dangerous place,
creating an escalating feedback loop between physiological and psychological processes. In fact, drugs that
block the impact of NE are now being used experimentally to determine if they may help individuals with
PTSD decrease their physiological response to reminders of their trauma (Brunet et al., 2008). NE levels have
also been tied to the vividness of perceptual experiences, a possible link with enhanced perceptual experience
during trauma (Todd et al., 2015).
Research with rats has demonstrated that exposure to inescapable shock sensitizes their hippocampi to

175
subsequent releases of NE under stress (Petty et al., 1994). This suggests that prolonged stress and trauma
may cause us to react more strongly to subsequent, milder stressors. This may help to explain the coping
difficulties seen in PTSD when individuals are confronted with mild to moderate stressors (Petty et al., 1994).
Think back to Sheldon, who still suffered from anxiety 60 years after his childhood experiences during World
War II.
High levels of dopamine correlate with hypervigilance, paranoia, and perceptual distortions when under
stress. Symptoms of social withdrawal and the avoidance of new or unfamiliar situations (neophobia) are
shaped by these biochemical changes. Elevated levels of endogenous opioids, which serve as analgesics to
relieve pain in fight-or-flight situations, can have a profoundly negative impact on cognition, memory, and
reality testing. Higher opioid levels also support emotional blunting, dissociation, depersonalization, and
derealization, all of which provide a sense of distance from the traumatized body (Shilony & Grossman,
1993). However, when they become harmful or maladaptive defenses, they disrupt our ability to stay engaged
in daily life.
As we have seen, high levels of glucocorticoids have a catabolic effect on the nervous system and are
thought to be responsible for decreased hippocampal volume and related memory deficits (Bonne et al., 2001;
Bremner, Scott, et al., 1993; Nelson & Carver, 1998; Watanabe, Gould, & McEwen, 1992). Chronically high
levels have negative effects on brain structures and the immune system, resulting in higher rates of learning
disabilities and physical illness, which enhances a person’s experience of being fragile and vulnerable. The
hippocampi of patients with childhood abuse and PTSD have been shown to be 12% smaller than those of
comparison subjects (Bremner et al., 1997). Another study showed that right hippocampi were 8% smaller in
patients with combat-related PTSD (Bremner et al., 1995). Glucocorticoids sacrifice long-term conservation
and homeostasis for short-term survival. In addition, lower levels of serotonin have been found in traumatized
humans and animals who were subjected to inescapable shock (Anisman, Zaharia, Meaney, & Merali, 1998;
Usdin, Kvetnansky, & Kopin, 1976). Chronically low levels of serotonin are correlated with higher levels of
irritability, depression, suicide, arousal, and violence (Canli & Lesch, 2007; Coccaro, Siever, Klar, & Maurer,
1989).
These biochemical and neuroanatomical changes are paralleled by such symptomatology as emotional
dyscontrol, social withdrawal, and lower levels of adaptive functioning. Together, these and other negative
effects of trauma result in compromised functioning across multiple domains. The impact of trauma depends
on a complex interaction of the physical and psychological stages of development during which it occurs, the
length and degree of the trauma, and the presence of vulnerabilities or past traumas. The impact of chronic
trauma becomes woven into the structure of a person’s personality and is hidden behind other symptoms,
making it difficult to identify, diagnose, and treat.

Expanding the Definition of Trauma


To the child . . . traumas are not experienced as events in life, but as life defining.
—Christopher Bollas

The standard diagnostic manual defines trauma as exposure to serious injury, sexual violence, and actual or
perceived threat of death; an individual may be exposed to trauma through experience, directly witnessing the
event, or by repeated exposure to traumatic information (American Psychiatric Association, 2013). However,
this definition is limited, and it does not address a wide range of traumas or how uniquely each individual will
respond to events that are perceived as traumatic. For a young child, trauma may be experienced in the form
of separation from parents, looking into the eyes of a depressed mother, or living in a highly stressful
household (Cogill, Caplan, Alexandra, Robson, & Kumar, 1986). For an adolescent, trauma may come in the
form of incessant teasing by peers or caring for an alcoholic parent. For an adult, chronic loneliness or the loss
of a pet may be traumatic.
There is increasing evidence that stress is possible even before birth; an unborn child can be affected by a
mother’s stress due to their shared biological environment. Studies suggest that maternal stress is associated
with lower birth weight, increased irritability, hyperactivity, and learning disabilities in children (Gunnar,
1992, 1998; Zuckerman, Bauchner, Parker, & Cabral, 1990). Rats born to stressed mothers show more
clinging behavior toward their mother and decreased locomotion and environmental exploration (Schneider,
1992). Prenatal stress may also result in permanent alterations in dopamine activity and cerebral lateralization,
making offspring more susceptible to anxiety and impairing their functioning as adults (Field et al., 1988).
This can be seen in children of Holocaust survivors who have an increased prevalence of PTSD despite similar

176
rates of exposure to traumatic events when compared to children of non-Holocaust survivors, suggesting a
transferred vulnerability from their traumatized parents (Yehuda, 1999).
Maternal depression may serve as a highly stressful or traumatic experience for infants and children.
Tiffany Field and her colleagues found that infants of depressed mothers show neurophysiological and
behavioral signs of depression and stress, including greater right frontal lobe activation, higher levels of NE
and cortisol, lower vagal tone, and higher heart rates (Field & Diego, 2008b; Field, Diego, & Hernandez-
Reif, 2006). Like their depressed mothers, these infants engage in fewer interactive behaviors (e.g., orienting
toward and gazing at others), which are vital for healthy development. Infants of depressed mothers also
behave in this way with other adults, making it more difficult for them to successfully interact with
nondepressed others (Field et al., 1988).
Another study found that depressed mothers were angry at their infants more often, disengaged from them
more frequently, were more likely to poke them, and spent less time in matching emotional states (Field,
Healy, Goldstein, & Guthertz, 1990). These results indicate that infants are modeling their mother’s
behavior, resonating with their depressed moods, and reacting to the negative behaviors directed toward them.
Although we would not consider these infants traumatized in the traditional sense, the loss of maternal
presence, engagement, and vitality may all be experienced as life threatening during the earliest years of life as
infants depend on these interactions for survival. Fortunately, it has been shown that interventions with
depressed mothers and their infants have positive results. For example, remission of maternal depression and
teaching mothers to massage their infants on a regular basis improves the infants’ symptoms and the mothers’
mood (Field, 1997).
The effects of early and severe trauma are extremely widespread, devastating, and difficult to treat. Because
of the importance of safety and bonding in the early construction of the brain, childhood trauma compromises
core neural networks. It stands to reason that the most devastating types of trauma are those that occur at the
hands of caretakers. Physical and sexual abuse by parents not only traumatizes children, but it also deprives
them of healing interactions that would mitigate the effects of trauma. They also lack a safe haven because
they cannot rely on their parents for protection, causing further damage. The wide range of effects involved in
the adaptation to early unresolved trauma can result in C-PTSD.
Dissociation allows the traumatized individual to escape the trauma via a number of biological and
psychological processes. Increased levels of endogenous opioids create a sense of well-being and a decrease in
explicit processing of the traumatic situation. Derealization and depersonalization reactions allow the victim
to avoid the reality of his or her situation or to watch it as a detached observer. These processes can create an
experience of leaving the body, traveling to other worlds, or immersing oneself in fantasy. In fact, many
victims of violence and sexual abuse report watching themselves being attacked from a distance. Hyperarousal
and dissociation in childhood create an inner biopsychological environment primed to establish boundaries
between different emotional states and experiences. If it is too painful to experience the world from inside
one’s body, the mind is capable of organizing the experience of self-identity outside the body.
Early traumatic experiences create the tendency to disconnect various tracks of neural processing, creating
a bias toward unintegrated information across the multiple domains of conscious awareness: sensation, affect,
and behavior. General dissociative defenses that result in an aberrant organization of networks of memory,
fear, and the social brain contribute to deficits of affect regulation, attachment, and executive functioning (van
der Kolk et al., 1996). The malformation of these interdependent systems results in many disorders that
spring from extreme early stress. Compulsive disorders related to eating or gambling, somatization disorders
in which emotions are converted into physical symptoms, and borderline personality disorders all reflect
complex adaptations to early trauma (Saxe et al., 1994; van der Kolk et al., 1996).

I Am Not Crazy!
Demoralize the enemy from within by surprise, terror, sabotage, assassination.
—Adolf Hitler

Jesse was referred to me by her neurologist after months of extensive medical and neurodiagnostic testing. Her
team of doctors could find no physical cause for the debilitating pain she experienced in her head and upper
body. She had tried several alternative forms of treatment, such as chiropractics and acupuncture, without
relief. Jesse came to see me at the insistence of her husband, and she was not the least bit happy about it. She
sat down with her arms crossed, her jaw set, glared at me, and said, “I am not crazy!”
For many years, life had been going well for Jesse. She had a solid marriage and a healthy and happy four-

177
year-old daughter. She enjoyed her work as an executive in a small technology firm, liked her colleagues, and
was a valued member of the team. About a year before coming to see me, she started experiencing pain in her
head, hands, and back, and she began a fruitless search for a cure. The pain became the center of her world as
her interest and ability to be an executive, friend, wife, and mother diminished. By the time she came for
therapy, she was spending most of her days taking medication and withdrawing to her room whenever she
could find an excuse. There was no longer any fun or relaxation in her life, and her husband had become
increasingly concerned.
Given her resistance to therapy and her fear of being seen as crazy, developing a therapeutic relationship
with Jesse was a slow process. She reluctantly began to share about her very troubled childhood, but at the
same time expressed confidence that she had overcome her traumatic past based on her success at work and in
her marriage. Unfortunately, a common occurrence in her childhood was to be locked in her room by her
father as a prelude to him beating her mother. She would lie in bed, frozen by both of their screams, her
mother’s cries for help, and the long ominous silences that followed. As she told me of her mother’s physical
abuse at the hands of her father, she remained confident that there was no connection between her present
physical pain and the emotional pain of her youth.
During her mother’s beatings, Jesse would eventually begin to pound on the door and yell for her mother.
As she grew older, however, she gave up her outward protests and, instead, lay in bed crying and clutching her
face and head. This clutching eventually turned into driving her nails into her head and shoulders, drawing
blood, and scarring herself. She showed me her scars with a mixture of shame and pride. As she described
these experiences, I began to suspect that her pain symptoms were physical manifestations of her implicit
memories of her traumatic past. The stresses in her present life, including the fact that her own daughter was
reaching the age she had been when she first became aware of her mother’s beatings, could serve as triggers to
activate these memories. From a psychological perspective, her physiological suffering could be seen as a form
of loyalty and continued connection to her mother.
I decided not to share these interpretations because of Jesse’s resistance to the possible psychological
origins of her pain. Instead, I continued to encourage her to talk about her childhood in as much detail as she
could tolerate. She went on to tell me about her teenage years when she nursed her mother through the final
months of a prolonged battle with cancer. Throughout my work with Jesse, I avoided any talk of her physical
pain and continued to refocus on her sharing childhood experiences with me.
In the process of repeatedly sharing stories from her childhood, Jessie’s memories became increasingly
detailed. Her emotions also became more available, and they better matched the situations she described. Jesse
expressed her rage at her father for his violent behavior and realized that she was also angry at her mother for
not leaving him when she was young. As she worked through these memories and put them into the
perspective of her current life, Jesse became open to understanding her pain as an expression of her traumatic
emotional experiences, and she gradually came to an understanding that instead of remaining loyal to her
mother through physical suffering, she could best honor the memory of her mother by being a good parent to
her own daughter.
As therapy progressed, we both began to notice that the intensity of her pain and the time she spent
focusing on it gradually diminished. Toward the end of our last session, she thanked me for helping her,
although she didn’t really understand how it happened. Jesse winked at me and said, “You are a tricky fellow.”

Traumatic Memory
Memories are nothing but the lash with which yesterday flogs tomorrow.
—Philip Moeller

It has long been recognized that high levels of stress impair our ability to learn new information (Yerkes &
Dodson, 1908). This is because the biochemical and hormonal changes triggered by stress impair protein
synthesis and other neuroplastic processes required for memory encoding. Trauma can also impair integration
across the domains of memory and is capable of dissociating the usually integrated tracks of sensation,
emotion, behavior, and conscious awareness.
When NE is administered to rats after a stressful event, low doses enhance memory retention whereas
high doses impair it (Introini-Collison & McGaugh, 1987). This finding parallels Yerkes and Dodson’s
theory that moderate levels of arousal enhance memory whereas high levels impair it. In a study by Cahill,
Prins, Weber, and McGaugh (1994), subjects were read both emotionally evocative and neutral stories and
shown related slides. Half of the subjects were given propranolol (a drug that blocks the effects of NE), and

178
the others were not. Results demonstrated that subjects who received propranolol had significantly impaired
memory for emotion-arousing stories, but not for neutral stories.
Activation of the amygdala and related physiological and biological changes are at the heart of modulating
traumatic memory (Cahill & McGaugh, 1998). The release of NE during the stress response serves to
heighten the activation of the amygdala, thus reinforcing and intensifying traumatic memories (McGaugh,
1990). Individuals with PTSD have had their amygdaloid memory systems imprinted with trauma at such an
extreme level that their memories are resistant to cortical integration, updating, or inhibition (van der Kolk et
al., 1996). This results in decreased attention to and processing of external stimuli, giving the traumatic
memories more power over conscious experience (Lanius et al., 2001). When we think of trauma
overwhelming our defenses, we can also think in terms of an intense activation of subcortical networks serving
to inhibit the participation of the cortex and hippocampi in the memory processes.
Traumatic experience can disrupt the storage (encoding) of information and the integration of the various
systems of attention and memory (Vasterling, Brailey, Constans, & Sutker, 1998; Yehuda et al., 1995; Zeitlin
& McNally, 1991). Memory encoding for conscious explicit memory can be disrupted when the hippocampus
is blocked or damaged by glucocorticoids or is inhibited by heightened amygdala activation. This could lead to
a lack of conscious memory for traumatic and highly emotional events (Adamec, 1991; Schacter, 1986; Squire
& Zola-Morgan, 1991). Memory integration can be impaired by disruption of the cortico-hippocampal tracks
dedicated to the integration of new memories into existing memory networks. Remember that these systems
also provide contextualization in time and space and integration of sensory, affective, and behavioral memory
with conscious awareness.
Although we may have very accurate physiological and emotional memories for a traumatic event, the
factual information may be quite inaccurate given the inhibition of cortico-hippocampal involvement during
the trauma. Given the tendency of the left hemisphere interpreter to confabulate a story in the absence of
accurate information, this may represent the underlying mechanisms responsible for what has been referred to
as false memory syndrome (Paz-Alonso & Goodman, 2008).

Traumatic Flashbacks and Speechless Terror


Memories are contrary things; if you quit chasing them and turn your back, they often return on their own.
—Stephen King

Flashbacks are frightening experiences commonly reported by traumatized individuals (Brewin, 2015). They
are described as full-body experiences of traumatic events, which include the physiological arousal, sensory
stimulation, and emotional impact of the traumatic experience. In a sense, a person experiencing a flashback is
transported back to the event, and as far as the brain is concerned, the trauma is happening right now.
Flashbacks are so intense that they overwhelm the saliency of the present situation and send the victim into an
all-too-familiar nightmare. Brain scans taken during flashbacks reveal increased activation in sensory, somatic,
and motor regions while those involved in episodic memory show decreased activity (Whalley et al., 2013).
The power of traumatic flashbacks was driven home for me one day in a therapy session with a
professional football player who was nearly twice my size. When recalling his early abuse, he began to cry
softly as he spoke of one particularly painful experience from his childhood. He described in agonizing detail
his small body growing limp after repeated blows from his father’s fists. Suddenly, he was standing over me,
breathing heavily, his arms down at his sides. Despite my alarm, I managed to calmly ask him what he was
feeling. While looking into my eyes, he said in a child’s voice, “Please don’t hurt me anymore.” His fear of me,
despite the difference in our sizes, was a stark demonstration of the all-encompassing nature of flashbacks and
their ability to override the objective reality of the moment.
Traumatic flashbacks are memories of a different nature than those of nontraumatic events. To begin with,
they are stored in more primitive circuits with less cortical and left-hemisphere involvement. Because of this,
these memories are strongly somatic, sensory, and emotional, as well as inherently nonverbal (Krystal,
Bremner, Southwick, & Charney, 1998). The lack of cortico-hippocampal involvement results in an absence
of the localization of the memory in time, so when the memory is triggered, it is experienced as occurring in
the present. Flashbacks are also repetitive and stereotypical, often seeming to proceed at the pace at which the
events originally occurred. This suggests that although the cortex may condense and abbreviate memories in
narrative and symbolic form, subcortical networks store memories in more stimulus-response chains of
sensations, behaviors, and emotions. In a sense, they become procedural memories, similar to learning to play
a piece of music note by note or a complex dance routine step by step.

179
In flashbacks, the amygdala-mediated fear networks biased toward the right hemisphere and subcortical
systems become dominant. The amygdala’s dense connectivity with the visual system most likely accounts for
the presence of visual hallucinations as part of flashbacks. Bereaved individuals often report seeing their loved
ones sitting in their favorite chair or walking across the room in a familiar way. Those who have been attacked
will sometimes think they see their attacker out of the corner of their eye.
Rauch and colleagues (1996) took eight patients suffering from PTSD and exposed them to two
audiotapes: one was emotionally neutral and the other was a script of a traumatic experience. While they were
listening to these tapes, patients’ heart rates and regional cerebral blood flow (RCBF) were measured by PET
scans. RCBF was greater during traumatic audiotapes in right-sided structures including the amygdala,
orbitofrontal cortex, insula, anterior, and medial temporal lobe, and the anterior cingulate cortex. These are
the areas thought to be involved with the experience of intense emotion.
An extremely interesting and potentially important clinical finding was a decrease in RCBF in and around
Broca’s area, an area of the left inferior frontal cortex that controls speech. These findings suggest an active
inhibition of language centers during trauma. Based on these results, speechless terror—often reported by
victims of trauma—may have neurobiological correlates consistent with what we know about brain
architecture and brain-behavior relationships. This inhibitory effect on Broca’s area may impair the encoding
of conscious semantic memory for traumatic events. It will then naturally interfere with the development of
narratives that serve to process the experience and lead to neural network integration and psychological
healing. Activating Broca’s area and related left cortical networks of explicit memory may be essential in
psychotherapy with patients suffering from PTSD and other anxiety-based disorders.

Activating Broca’s Area During a Flashback


Hope will never be silent.
—Harvey Milk

Jan, seeing me for a one-time consultation, reported that she had suffered severe physical and sexual abuse
from early childhood into her late teens. Before our first meeting, she told me that her flashbacks were
increasing in frequency to three or four a day. Although her therapist had encouraged her to experience them
and express her emotions as much as possible, Jan felt like she was getting worse instead of better. Expressing
her feelings seemed to only trigger more frequent and intense flashbacks. She reported becoming less and less
functional, which made her decide that she needed a different approach to therapy.
Jan arrived at my office with a stack of diaries and the Wall Street Journal under her arm. It was hard to
believe that this was the same person I had spoken to over the phone. My first thought was that dissociation is
an amazing defense. Jan was a well-dressed woman in her mid-40s who was obviously bright and insightful.
The childhood experiences she recounted in my office were horrendous, and I marveled at her very survival.
Her intelligence and sheer will to live were remarkable. It seemed obvious, however, that her repeated
reexperiencing of these memories was not helping. The nature of these memories was not changing over time,
nor were the emotions evoked by her memories diminishing. In this case, each flashback seemed to traumatize
her anew.
She began by talking about her work and then described the psychotherapy and other forms of treatment
that she had engaged in. Approximately 10 minutes into the session, as Jan was discussing the family members
who had abused her, she began to have a flashback. She reported pain in various parts of her body and
contorted as if what she was describing was happening to her at that very moment. She began to gag as the
memory of the sexual abuse from decades earlier was evoked. She was reexperiencing these painful episodes,
not only visually in her mind, but as somatic memories throughout her body.
As she curled into the fetal position on the couch and gasped for breath, my mind raced, trying to think of
a way to help. Remembering the research done by Rauch and his colleagues, I decided to try to activate
Broca’s area. I began speaking to Jan in a firm, but gentle voice, loud enough to reach her in the midst of her
traumatic reenactment, but not so loud as to frighten her and add to her trauma. I wondered if it mattered
which ear I spoke into and which ear had a more direct connection to the left hemisphere language centers. I
moved closer to her, careful not to get too close, and repeated over and over, “This is a memory; it isn’t
happening now. You are remembering something that happened to you many years ago. It was a terrible
experience, but it is over. It is a memory. It is not happening now.”
As I repeated these and similar statements, I was concerned that Jan would be unable to breathe or that my
presence might cause more fear. The words of one of my supervisors flashed through my mind: “Whatever

180
you do, don’t panic.” I was also encouraged by the fact that Jan had survived this many times before. After 10
minutes, which to me seemed like 10 hours, she calmed down and returned to the present. Jan reported that
she heard me speaking as if I were far away, but focused on my voice and words as best she could. It was as if I
was there in the past with her, calling to her from a safe future where she would be away from all these people
who hurt her.
At the end of the session, she thanked me and left; I didn’t hear from her for a number of months. When
she called one afternoon, she reported that since her visit with me, the nature of these flashbacks had changed.
She said that she had wanted to wait before she called me because she didn’t expect that the change would
last. Jan described that since our session, the flashbacks were less physically intense and less frequent. On a
few occasions she had even been able to stop one that was coming on by thinking of my words during the
session: “This is just a memory. You are safe now. No one can hurt you.”
Perhaps most interesting was that during her flashbacks, she was now able to remember that she was not a
child, that she was not to blame, and that it was those who were hurting her who were bad. Although her
other therapists had told her this in the past, only recently could she process these thoughts during her
flashbacks. I told her that I felt these were signs that the experiences were beginning to be connected to her
conscious, adult self and that now she was able to fight and care for herself even in the face of her past. I
encouraged her to keep talking throughout the flashback experiences and bring with her as much
assertiveness, anger, and power as she could muster. After a few minutes, we ended our conversation, and I sat
back struck at how our knowledge of neuroscience could be applied to psychotherapy.
It is impossible for me to know with any certainty whether what I did with Jan during our one meeting
had anything to do with the changes she experienced during her flashbacks. If it did, perhaps the active
ingredient was the simultaneous activation of the verbal areas of the left hemisphere along with the emotional
centers of the right hemisphere and limbic structures that stored her flashbacks. Being simultaneously aware
of inner and outer worlds may support a higher level of cortical functioning and increased network
integration. In other words, this process results in a memory configuration that is no longer only implicit, but
instead becomes integrated with the contextualizing properties of explicit systems of memory in the cerebral
cortex (Siegel, 1995).
Speechless terror, which has been recognized as part of posttraumatic reactions since ancient times, now
has a neural correlate consistent with what is known about brain functions. Why does Broca’s area become
inhibited during trauma? Why would evolution select silence in times of crisis? Perhaps when one is
threatened, it is better to either run, fight, or remain silent in hopes of eluding detection by a predator. In
other words, evolution may have shaped the brain to “Shut up and do something!” when in danger. The
freezing reaction of animals, being still and quiet when they sense a predator, allows them to be less visible
(because a still and silent target is more difficult to detect). Spoken language is sound, which primitive fear
circuitry is able to silence. Perhaps those early prehumans who hung around for conversation and negotiation
with predators did not fare well enough to pass down as many genes as did those who either kept quiet,
fought, or ran away.

Addiction to Stress and Self-Harm


Every form of addiction is bad, no matter whether the narcotic be alcohol or morphine or idealism.
—C. G. Jung

Another clinical phenomenon with a possible underlying biochemical mechanism is what appears to be an
addiction to stress experienced by some patients with PTSD. While anxious and ill at ease in normal daily life,
they report feeling calm and competent in risky or life-threatening situations. A so-called normal life leaves
traumatized persons a blank screen onto which their dysregulated psyches can project fearful experiences,
keeping them in a state of fearful vigilance (Fish-Murry, Koby, & van der Kolk, 1987). This may motivate the
creation of stress in order to stimulate the production of endogenous opioids that would lead to an increased
sense of calm and well-being. Paradoxically, trauma, in many cases, can lead to a sense of competency, control,
and safety.
Those addicted to stress can become so physically worn down that they will present in therapy with
exhaustion, depression, and a variety of medical conditions. It is as if they have a drug addiction, except that it
is completely unconscious, and they are their own pharmacy. Initial work with these patients should ideally
focus on helping them to reduce stress and learning to tolerate and understand the anxiety triggered by the
absence of stress. This can usually be accomplished through a combination of stress-reduction techniques,

181
medication, and psychotherapy.
The addiction to stress has a related but more severe variant: self-harm. Adults who engage in repeated
self-harm commonly describe childhoods that included abuse, neglect, or a deep sense of shame. This
correlation has led many theorists to explore the psychodynamic significance of self-harm as an ongoing
negative attachment to destructive parents. Along these lines, suicide has been described as the final act of
compliance with the parents’ unconscious wish for the death of the child (Green, 1978). In addition, there
appears to be a strong association between self-harm and attachment disorders because self-injurious
behaviors are a consequence of feelings of abandonment and loss.
Endogenous opioids may also play a role in some instances of self-harm and suicide (van der Kolk, 1988).
This opioid system, originally used to cope with pain, was adapted by later-evolving networks of attachment
to reinforce the positive effects of bonding (Pitman et al., 1990). Research has demonstrated that the
frequency of self-harm decreases when patients are given a drug to block the effects of endogenous opioids
(Pitman et al., 1990; van der Kolk, 1988). Abstracting from the animal model, this would suggest that the
endorphins released during self-injury reverse the feelings of distress activated by abandonment panic. The
analgesic effects of these morphine-like substances may account for the reports that describe a sense of calm
and relief after individuals cut, burn, or hurt themselves. Thus, self-injurious behaviors may be reinforced by
both psychological factors and endogenous opioids.
Repeated suicide attempts are also reinforced by the rapid attention of health care professionals, family,
and friends, which may boost endorphin levels triggered by the arrival of loved ones. When woven into coping
strategies as a means of affect regulation, this attention-getting behavior results in a kind of characterological
suicidality (Schwartz, 1979). In other words, suicidal ideation and behavior become a method of summoning
connection. This behavior parallels the distress calls of primates whose endorphin levels drop in the absence of
their mothers. The reappearance of the mother results in an increase of endorphin levels, and crying ceases.
Characterological suicidality can come to serve a similar biochemical regulatory purpose if this system was
inadequately formed during childhood. Although there are many sound psychological explanations for the
relationship of childhood abuse to self-harm and suicidality, a pharmacological intervention designed to block
the impact of endogenous endorphins may also prove helpful.

Neural Network Integration


The patient discovers his true self little-by-little through experiencing his own feelings and needs, because the analyst is able to accept and respect
them.
—Alice Miller

Unresolved trauma disrupts integrated neural processing so that conscious awareness is split from emotional
and physiological experiences. Some believe that dissociative symptoms immediately following a traumatic
event are predictive of the later development of PTSD (Koopman et al., 1994; McFarlane & Yehuda, 1996),
while others emphasize the importance of posttrauma dissociation as most predictive (Briere, Scott, &
Weathers, 2005). Neurochemical changes and a lack of integration of right- and left-hemisphere functions
may also impede interpersonal bonding and bodily regulation (Henry, Satz, & Saslow, 1984). Children
victimized by psychological, physical, and sexual abuse have been shown to have a significantly greater
probability of brainwave abnormalities in the left frontal and temporal regions (Ito et al., 1993), and
brainwave dyscoherence may put individuals at higher risk for developing all forms of psychiatric disorders
(Teicher et al., 1997).
The biochemical changes that occur secondary to trauma enhance primitive (subcortical) stimulus-
response pairing of conditioned responses related to sensation, emotion, and behavior. These changes
undermine cortical systems dedicated to the integration of learning across systems of memory into a coherent
and conscious narrative (Siegel, 1996). As we understand more about the neurobiological processes underlying
PTSD, we will better learn how to treat and possibly prevent this debilitating yet curable mental illness.
Therapies of all kinds, especially those within the cognitive schools, have proven successful in the
reintegration of neural processing subsequent to trauma. Systematic desensitization, exposure, and response
prevention can all enhance these integrative processes.

Summary

182
The brain’s reaction to trauma provides us with a window into the functions and effects of neural network
dissociation. From the physiological symptoms of adult PTSD to the characterological adaptations of long-
term adjustment to early trauma, we see the brain, body, and psyche attempting to survive in the face of
overwhelming dysregulation. The array of adaptations to stress and trauma are at the core of the work of the
psychotherapist. The safe emergency of psychotherapy activates dissociated neural networks and attempts to
reintegrate them in the service of decreased arousal and improved functioning. From the first moments of life,
stress shapes our brains in ways that lead us to remember experiences most important for survival.

183
Chapter 18
The Self in Exile: Narcissism
and Pathological Caretaking
Love is a union with somebody outside oneself, under the condition of retaining the separateness and integrity
of one’s self.
—Erich Fromm

Each of us is born twice: first from our mother’s body over a few hours and then again from our parents’
psyche over a lifetime. As we have discussed, the organization of the social brain is initially sculpted via
parent-child interactions. These interactions shape the infrastructure of our moment-to-moment experience
of other people and of the world. As the child’s brain continues to form, self-awareness and self-identity
gradually coalesce. Consciousness and identity are complex functions that appear to emerge from the synergy
of multiple, primarily unconscious neural networks. Pathological states highlight the fact that the self is a
fragile construction with considerable flexibility concerning our location, experience, and organization.
Victims of rape and torture frequently report out-of-body experiences during their ordeal. A young woman
described to me, in great detail, how she stood behind a closet door, watching herself being raped from across
the room from where it was actually happening. Another client, who was brutally attacked while getting into
his car after work, told me that he watched himself from across the street as he was repeatedly stabbed. The
perception of the self is also vulnerable to alteration and distortion. Anorexic clients, with their bones
protruding through their skin and their health in serious jeopardy, insist they look fat. Patients with multiple
personalities are perhaps the most complex example of the plasticity of self as they generate many
subpersonalities associated with different behaviors and emotional states.
Narcissism, a common form of self-disturbance, is often related to a reversal of the mirroring process
during childhood. Narcissistic children’s social brains and sense of self are not shaped by their own emerging
emotions and sensibilities; rather, they are determined by their parents’ need for nurturance, attunement, and
affect regulation. What emerges in the narcissist is what Winnicott called the false self, a pseudoadult
embedded in the networks of the left-hemisphere interpreter, which filters out emotional input from the right
hemisphere and the body. In this chapter, we explore the reversal of the mirroring process and the adult
conditions of pathological caretaking and codependency that emerge from early dysfunctional attachment
experiences (Bachar et al., 2008).

Silent Hammers
I cannot give you the formula for success, but I can give you the formula for failure—which is . . . try to please everybody.
—Herbert Swope

Jerry, a successful screenwriter, came to therapy complaining of depression and exhaustion. His long work
hours and last-minute deadlines kept him in a state of chronic stress. His minimal personal life centered on
his girlfriend, Cara, whom he described as “high maintenance,” likening their relationship to having a second
job. Jerry experienced his life as a relentless struggle to please Cara, his boss, and everyone he knew. Despite
his depression, Jerry felt guilty about coming to therapy and expressed fear about wasting my time; his identity
was as a helper, not the one who received help. “After all,” he said to me, “you could be spending your time
seeing someone who really needs you.”
After a few sessions, it became clear to me that Jerry had spent the first half of his 39 years taking care of
his immature and self-centered parents. All of his subsequent relationships reflected a similar pattern.
Although he described his romantic relationships in positive terms, he also reported feeling deprived of
attention and nurturance. He seemed to be attempting to please others in order to gain the love and attention

184
that he had always longed for from his parents. His efforts would invariably end in sadness, resentment, and
withdrawal. Although he was exhausted from trying so hard and failing so completely, Jerry still maintained
the hope that his efforts would someday pay off.
Working with Jerry was both fascinating and frustrating. He questioned nearly everything that I said,
revealing his distrust of anyone’s ability to help him. He had learned early on that people who should be
taking care of him were not up to the job. On the other hand, his flair for the dramatic resulted in
entertaining stories in almost every session. He had an uncanny ability to intuit my interests, and I would
often lapse into being the audience for his one-man show. I soon realized that Jerry experienced me as another
person to entertain and take care of while he waited to receive care in return. At the same time, he resisted
every attempt I made to help him.
Jerry was hesitant to discuss his childhood, saying that he remembered little of his life before he left for
college. Our sessions consisted primarily of stories of his interactions with unappreciative others and his
attempts to enlist my understanding and support for his side of the story. He was both comforted by my
compassion and annoyed with my continued suggestions that he pay attention to his inner emotional world,
especially when engaged with others. The intensity of his defenses reflected his emotional vulnerability, and I
realized that I needed to be careful not to move too quickly. On the other hand, if I went too slowly, he might
become resentful and terminate therapy in the same pattern that he had ended so many other relationships.
At some point, I suggested that he write a screenplay about himself. Jerry agreed and soon created a story
about a little man named Hal who sat at a control panel in his head. Hal was a sort of ship’s captain at the
helm of the U.S.S. Jerry. When Jerry was alone, Hal would monitor the people in Jerry’s life. A wall full of TV
screens kept track of where they were, what they were doing, and whether they were thinking of Jerry. When
he would come into contact with another ship, like his girlfriend Cara, Hal would select a holographic image
of Jerry specifically for that person. Hal had a screen to monitor Jerry from the perspective of the other ship,
to make sure that the hologram had the desired effect. Jerry told me that Hal’s purpose was best described by
the show business adage: “Give the people what they want.”
I was impressed by the story of Hal and felt that Jerry was probably telling me that his sense of self was
organized around his theory of the minds of others. Nothing about Hal focused on Jerry’s own thoughts,
feelings, or needs. Jerry only knew he needed to understand the needs of others; for him, this was love; this
was life. I realized that Hal reflected the early sculpting of Jerry’s social brain and how he survived childhood.
The story of Hal was a symbolic representation of the core emotional drama of Jerry’s early life. Despite the
obvious nature of this story, Jerry refused to entertain interpretations. He seemed unable to experience the
world from his own point of view. In Jerry’s conscious experience, he was Hal.
A few sessions later, Jerry used another metaphor to illustrate his inner world. He described himself as a
house on a movie set with a perfectly detailed facade, but no finished interior; there was no true place to live.
If you looked behind the set, there were only exposed 2 × 4s and bare floors. He said, “A director knows where
to place the camera to maintain the illusion of a real building.” In a moment of receptivity and trust, he told
me that his goal for therapy was to finish building out the interior space so that he could finally stop worrying
about camera angles. Jerry was tired of feeling like a ghost in a world full of real people. I could feel his inner
world shifting as his emptiness surfaced into consciousness.
I told Jerry, “As we grow, the story of who we are is written in collaboration with those around us,
especially our parents. They watch us, listen to us, and try to help us put into words what we struggle to
express. This helps us to write our story. When parents didn’t get this help as children or when they are
dealing with some problem, they may look to their children to take care of and parent them instead.”
I think this is what may have happened with Jerry’s parents. He learned to be sensitive to them and attend
to their needs when they should have been helping him to write his story. In a sense, Jerry’s story was written
by default; in serving others, he was hoping to find himself, and after all these years he was still searching. I
suggested to Jerry that his choice of a writing career may have been an expression of his desire to write his
story. Although serving others has many good and honorable aspects, it was time to write a new story in
which his own needs would be balanced with the needs of others.
Jerry asked, “Is this why, when people ask me how I am, the first thing I think about is how others in my
life are doing?” I nodded as I watched the muscles in his face relax. A new understanding of a familiar
behavior was emerging, and his inner world was being reorganized. After a period of silence, he said softly, “I
even do it with you, don’t I? I pay you to take care of me, but I end up entertaining you and protecting you
from my needs and negative feelings.” The session was over, and he walked silently out of the office. I
wondered if he would be able to withstand the stress of these insights enough to return for our next session.
He did come back and uncharacteristically touched my shoulder as he passed me on the way to his seat.
“You’re gonna think you’re pretty smart when I tell you this one.” I noticed his voice had changed; it was no

185
longer the voice of an entertainer. Jerry was sharing this experience from a different place within himself. On
the previous Friday, he had gone out with Cara, and they didn’t get back to her apartment until 3:00 AM.
The combination of alcohol and exhaustion made him fall asleep as soon as his head hit the pillow. After what
seemed like only a few minutes, he was jarred awake by the banging of hammers outside the bedroom
window. He picked up his throbbing head and saw on the clock that it was not even 7:00 in the morning.
Jerry said, “I felt like my head was going to explode!”
Jerry’s face became increasingly tense as he described the murderous fantasies passing through his mind.
He imagined going outside and punching out the entire team of construction workers. Just as he was about to
jump up and run outside, Cara looked up at him with an expression of exhausted rage and said, “I’m going to
go out there and kill those guys.” Apparently, they had been waking her up that way for most of the week.
Falling back on the couch in my office and letting out a sigh, Jerry began to describe how her words
triggered something familiar within him. “When I heard Cara’s anger, I became a different person. My
exhaustion vanished. I felt energized and alive! I completely forgot about the construction workers. My total
focus instantaneously shifted to making Cara feel better.” He described jumping up as if well rested and said,
“A little breakfast will make you feel better.” He became oblivious to the sound of the hammers and bounded
into the kitchen as Cara pulled the blankets back over her head.
Ten minutes later, Jerry was at the stove brewing coffee, frying eggs, and about to call out “Breakfast is
served,” when he again noticed the hammers. He said, “I was amazed that I hadn’t heard them since I had
gotten out of bed.” He realized that they hadn’t stopped, but from the moment he saw the anger on Cara’s
face, he became completely involved in activities to cope with her feelings, utterly ignoring his own. He
realized that attending to Cara’s distress resulted in him being unaware of his own.
As he stood, spatula in hand, these feelings triggered a long-forgotten memory. He remembered coming
home from school to find his mother crying, head down in her arms on the kitchen table. Jerry remembered
his terror as he went over and stood beside her. She didn’t respond when he asked her what was wrong; she
just continued sobbing. He didn’t know what to do or whom to call for help. His father had left them months
before, and his mother had grown more silent and withdrawn with each passing week. He stood motionless,
not knowing if she was even aware of his presence.
Jerry tried to engage her in conversation, telling her about his day at school. He even told some
inappropriate jokes in the hope of making her angry enough to chase after him. She remained silent, crushed
by the weight of her sadness. Jerry recalled growing increasingly desperate and afraid as time passed. He stood
motionless as the afternoon light turned to dusk. He eventually came up with the idea to cook for her and
went over to the stove to fry her some eggs, the only thing he knew how to make. After collecting what he
needed, he pulled a chair over to the stove, stood on it, and started to cook. This seemed to get her attention;
she came over, and they cooked together in silence.
During our first few sessions, Jerry had mentioned his parents’ divorce, his father’s drinking, and his
mother’s depression. He characterized his home as “a vault of silence” with very little interaction or shared
activities. The family took no meals or vacations together; most of their energy was used up in day-to-day
survival. He spoke of these things in passing, with little emotion and an insistence that his childhood had no
connection to his adult difficulties. He said he had coped with his family situation by immersing himself in
books and writing stories. This memory provided evidence for what I had suspected; Jerry had grown up with
too much stress and too little parenting. He had lost himself in everyone else’s stories, but had not received
the emotional support and mirroring he needed to write his own.
Jerry came away from the experience of the hammers and the memories they triggered with a new
perspective. He could see, from this and other childhood memories that began to emerge, how he was
constantly frightened by the emotional instability and distance of his parents. It was also clear that he, like
Hal, was constantly monitoring the feelings and needs of others. I found it especially fascinating how his own
anger, frustration, and exhaustion disappeared in the face of Cara’s negative feelings. He later told me that he
figured this was why he felt so uncomfortable and vulnerable when he wasn’t in a relationship. Never having
learned how to regulate his own feelings, it was safer to stay focused on the minds of others.
The power of these insights was unsettling for Jerry, and it took many weeks for him to get back on track.
As he regained his equilibrium, we began to apply this new knowledge to more aspects of his life. Thinking
back on his romantic partners, he realized that they were never as good at anticipating or attending to his
needs as he was to theirs. This imbalance led him to feel unloved and uncared for in relationship after
relationship. At one point, he stood up abruptly and shouted, “I blamed them for not taking care of me even
though I wouldn’t let them.” Jerry came to understand that feeling like an empty shell was connected to not
knowing his own feelings. He said, “My feelings were never important when I was a kid. It’s my feelings that
are going to furnish those empty rooms inside of me, and I have to have my own feelings to be whole.”

186
These experiences were a turning point in Jerry’s treatment. We created a common language and used it to
explore Jerry’s inner world. Hal gradually replaced some of the old monitors with new ones designed to track
Jerry’s own feelings. Eventually, Hal and his monitors became unnecessary. At times, Jerry feared he was
being too selfish by considering his own needs. He did lose some friends who relied on his constant and
unilateral attention. Cara and Jerry grew closer, however, and she admitted that she liked Jerry better when he
didn’t try to make her happy all the time. Jerry slowly recovered from his pathological caretaking and
graduated to become a caring person.
Jerry’s difficulties highlight a number of principles related to the development and organization of the
brain examined in previous chapters. During childhood, Jerry’s brain adapted to a demanding and
nonnurturing emotional environment. Early in life, his survival depended on being highly aware of the
feelings and needs of his parents. As a result, the neural systems of his social brain became hypersensitive to
his parents’ facial expressions, body language, and behaviors, helping Jerry to monitor his parents’ emotions
and behave in ways that regulated their feelings and bonded them to him. Systems normally used to help
children come to experience themselves were usurped to monitor other people in his life. This was how Hal
came to be. Jerry’s mirroring skills continued to be utilized in adulthood with his friends, employers, and
therapist through emotional caretaking and entertaining stories.
Because Jerry didn’t have help processing and integrating his own emotions during childhood, they
remained chaotic, frightening, and overwhelming well into adulthood. Caretaking was reinforced not only by
those he attracted, but by the avoidance of his own disorganized emotions by attending to the feelings of
others. Caretaking evolved into a form of affect regulation, as well as a way of connecting to others via a false
self (Hal). This was demonstrated by the instantaneous inhibition of his feelings of anger at the construction
workers when he realized Cara was awake and upset. From early in development, hidden layers of neural
processing organized the inhibition of his own feelings and directed his social brain to focus on the internal
state of others. This neural network organization was in place prior to Jerry’s development of self-conscious
awareness, making this way of experiencing the world completely unconscious and an a priori assumption of
his life.
Jerry’s evolving sense of self was shaped via these processes, through the eyes, minds, and hearts of others.
Left-hemisphere processing networks, which inhibit affect and participate in the creation of stories about the
self, allowed Jerry to become a functional adult with a successful writing career. His ability to describe himself
symbolically, first as a monitoring robot and then as a director planning camera angles, reflected his
subconscious awareness of his self-organization. Without assistance in connecting these insights to the
organization of his conscious self, they remained correct, but useless bits of information contained within
dissociated neural networks.
Interpretations—such as the one Jerry made connecting his cooking eggs for Cara and taking care of his
mother—triggered emotions of sadness and loss. Making his defenses conscious appeared to activate the
emotional networks he had been inhibiting for most of his life. The higher association areas involved in the
organization of conscious awareness appear capable of the plasticity required for qualitative changes of
experience. Implicit memories of early childhood—stored within networks of the social brain—were
experienced emotionally and expressed in a variety of ways including Jerry’s distrust of the ability of others to
help him. These were echoes of his disappointment in his parents’ inability to take care of him when he was
young.
Jerry’s insecure attachment resulted in a complex array of emotional and behavioral adaptations. His lack
of conscious recall for much of his childhood was one example. Another was his unconscious expectation that
the expression of his needs would be met by more emotional pain. All of Jerry’s struggle to attain love and
caring reflected his brain’s adaptation to a childhood full of impingements and the absence of an integrated
and integrating other. Jerry’s parents did not assist him in creating a narrative identity grounded in his own
experience. In therapy, I helped Jerry to refocus his considerable creative skills to write his own story.

Interpretations and Neural Plasticity


There’s a world of difference between truth and facts. Facts can obscure the truth.
—Maya Angelou

You may remember that interpretations are sometimes referred to as the therapist’s scalpel. In making an
interpretation, the therapist points out an unconscious aspect of the patient’s experience, such as a defense he
or she is using to avoid negative feelings. For a client who employs humor to avoid feelings of abandonment

187
after a divorce, it may entail reminding him that he is experiencing many symptoms of depression or that his
eyes are moist. For another who is enraged at a minor slight by a coworker, an interpretation might consist of
connecting her present feelings to emotional memories of abuse from a previous relationship. In both of these
instances, the therapist addresses what appears to be a disconnection between different tracks of cognitive,
emotional, sensory, and behavioral processing.
When an interpretation is accurate and delivered in an appropriate and well-timed manner, a number of
things occur. The client generally becomes quiet, and there may be a change in facial expressions, posture, and
tone of voice. The client will frequently begin to fully experience the emotions against which he or she was
defending. There is a shift from fluent reflexive language to speaking in a slower, more self-reflective manner.
Some clients report becoming confused or disoriented, whereas others describe physiological symptoms of
panic or grief. Clients with borderline personality disorder may demonstrate extreme reactions to
interpretations, including losing emotional control and functional decompensation. They may become
extremely emotional, bolt from the consulting room, or engage in self-injurious behavior. Patients like these
appear unable to cope with the emotions released when their defenses are made conscious.
What might be happening in the brain during and after an accurate and well-timed interpretation? Each
interpretation that hits home is like the death of a small aspect of the false self. I suspect that it begins with
seeing past the products of the left-hemisphere interpreter, which disinhibits the activation of subcortical
circuits containing negative memories. In the process of working through these new feelings, the client
gradually regains emotional equilibrium, allowing for plasticity and new learning in prefrontal regions. The
concurrent availability of negative subcortical memories and the enhanced ability of the cortex to create new
connections allows neural networks containing various components of a particular memory to become
integrated. Like breaking and resetting a bone that has healed badly, memories become unstable and can be
reformed in a more positive way. This process allows painful implicit memories to be accessed by cortical
networks for contextualization in time and space and to be regulated and inhibited where no longer necessary.
Bringing a defense to consciousness activates both the cortical networks that organize the defense and the
subcortical networks that contain the negative memories and associated affect. This disinhibition results in the
emotional and physiological arousal seen in therapy as the amygdala becomes reactivated and alerts the body
to the old danger. This may also be the mechanism for regression through the reactivation of old sensory-
motor-affective memories stored in normally inhibited amygdaloid systems. There is most likely a shift in
hemispheric bias from left to right, correlated with the breakthrough of negative emotions. This left-to-right
shift may account for the cessation of reflexive social language of the left-hemisphere interpreter and a shift to
greater self-awareness.
Interpretations need to undergo a process called working through, meaning they need to be stated,
restated, and applied to multiple situations and circumstances, similar to relapse prevention in cognitive
therapies. This process serves to connect new learning to multiple memory networks, and may need to reach a
certain critical mass of connections throughout the brain to become reflexive. Working through reflects the
expansion and stabilization of new associative matrices of memory. It is also reflected in the construction of a
new narrative containing altered aspects of behaviors, feelings, and self-identity that serve as a way to retain
and reinforce new learning.

Pathological Caretaking
By the time a man is 35, he knows that the images of the right man, the tough man, the true man which he received in high school do not work in
life.
—Robert Bly

Jerry’s pathological caretaking is one possible expression of a disturbance of self referred to as narcissism.
Narcissism is characterized by a two-sided existence: one reflecting an inflated sense of self-importance, the
other mired in emptiness and despair. The origin of this formation of the self occurs when a child looking for
love and attunement instead discovers the mother’s own predicament (Miller, 1981). The child, robbed of the
possibility of self-discovery, compensates by caring for the parent under a real or imagined threat of
abandonment.
Bright and sensitive children attune to and regulate the parents’ emotions and come to reflect what the
parents want from them. These children will usually appear mature beyond their years and find comfort in
their ability to regulate the feelings of the people around them. Because of their power to regulate the affect of
one or both parents, these children are filled with a sense of inflated self-importance. Miller and Winnicott

188
call pathological caretaking a particular manifestation of the false self, described by Jerry as both a theatrical
facade and an inner robot monitoring others.
The other side of narcissism reflects aspects of the child’s emotional world that have found no mirroring.
This true self, or the part that is unique to the individual and searching for expression, is left undeveloped and
eventually forgotten. This emotional core, or inner child, secretly and silently awaits parenting in each new
relationship while dutifully taking care of others. This aspect of the emotional self shows the emptiness, loss,
and shame of being abandoned, and it demonstrates how survival was contingent on taking care of others.
This is the source of depression, the sense of being a fraud, and the lack of an emotional connection with life
associated with core shame.
The development of the social brain and the subsequent formation of the sense of self become dedicated to
the prediction of, and attunement to, the moods and needs of parents and others. This ability serves to ward
off abandonment anxiety while truncating the development of an understanding, expression, and regulation of
one’s own feelings. Such children grow into self-awareness experiencing others’ emotions as their own, and an
overwhelming sense of responsibility, or even compulsion, to regulate the emotions of those around them. In
Jerry’s case, cooking breakfast for the upset women in his life helped him to avoid his own poorly understood
and dysregulated emotional world.
Thus, caretaking of others serves as a substitute for self-soothing abilities and inner emotional
organization. Pathological caretakers come to therapy primarily because they are depressed and exhausted by
their inability to create a boundary between themselves and the needs of others. Although being with others is
hard work, being alone is even more difficult because they need to regulate others in order to avoid their
internal world. For these people, a battering or abusive relationship is far less frightening than solitude.
Caretakers are difficult clients because they have learned during early attachment relationships that help is not
forthcoming when they are in distress. Like Jerry, they have come to believe that it is best to put their inner
needs out of mind and keep “giving the people what they want.”

Alice Miller: Archaeologist of Childhood Experience


I was not out to paint beautiful pictures. . . . I wanted only to help the truth burst forth.
—Alice Miller

The central importance of parental relationships in shaping the social brain is nowhere better articulated than
in a series of elegantly simple works by the Viennese psychotherapist Alice Miller. Her work with what she
called gifted children targeted adults, like Jerry, who were raised by parents whose emotional needs were
greater than their ability to attune to their children. Taking a stand against her analytic colleagues, Miller
reshaped the therapeutic role into one of being an advocate for the child within her adult patients. Reaching
back through the years to reconnect with long-forgotten, repressed childhood experiences, she reinterpreted
much of her clients’ adult behaviors as a reflection of their early adaptational attachment histories. In
observing the nonverbal reenactment of implicit memory, Miller formed hypotheses concerning what her
clients had been exposed to, how their young brains had adapted, and what it would take to unearth the true
self that they had to abandon in order to survive.
Miller’s archaeological view of exploring childhood included the awareness that memory from different
developmental stages also reflects different modes of experience as the brain develops. In her role as an
advocate, Miller saw therapy as a process in which therapists help clients unearth their history, not from an
adult point of view, but as it was stored as a child. Memories in these implicit subcortical networks do not
change with time, but remain in their initial form, as they were experienced by the level of brain development
and maturity at a young age. Miller formulated this view from her clinical experience rather than from a
knowledge of the multiple systems of explicit and implicit memory.
She used the term double amnesia to describe the process by which these children have had to first forget
certain parts of themselves (e.g., feelings, thoughts, and fantasies) that could not be accepted or tolerated by
their family. The second layer of forgetting is to forget that these feelings have been forgotten. These two
layers of forgetting ensured that children would not slip back into wanting or needing the parents they do not
have. Given our knowledge of the multiple systems of memory and their dissociability, Miller’s double
amnesia is most likely grounded in both the disconnection between systems of implicit and explicit memory
and constructing a self-narrative that excludes reference to personal needs.
The lack of assistance in the construction of a self-narrative, combined with the heightened anxiety and
vigilance necessary for survival with narcissistic caretakers, also leads to a disruption in the consolidation of

189
autobiographical memory. Miller’s work is a reconstruction of the past based on available conscious memory
woven together with nonverbal expressions from implicit systems of memory. Such patients’ current
experiences are examined for emotional truths, then traced back through a hypothesized trajectory. The
patients’ considerable empathic ability can be utilized to their advantage by asking how they think some other
child might feel in a situation similar to their own. This method leads to a narrative with emotional
congruence with conscious experience even if its veridical truth is untestable. Later in life, the gifted child’s
lack of rebellion becomes the problem. Unable to attend to their own emotional memories and unable to
construct the story of their lives, these children constantly search for someone who needs nurturance.
For Miller, gifted children are exquisitely sensitive to the cues of parents, and they have the innate ability
to mold themselves to their parents’ conscious and unconscious messages. These are the children who are
often called codependent and are overrepresented in the service professions such as doctors, nurses, social
workers, and therapists. In essence, their jobs are an attempt to parlay their defenses into a career. Jerry—with
his sensitivity, intellect, and wit—fit well in this category.
Although the gifted children described by Miller may be quite functional, they often feel empty and
devoid of vitality. This is because their vitality and true self were not adaptive and were inhibited and banished
from awareness. This creates a vulnerability, not only to disturbances in personality, but also to the
unconscious transmission of mirror reversal to the next generation. Parents who have not been adequately
parented themselves often look to their children for the nurturance and care they were unable to receive years
before. Miller stated, “What these mothers had once failed to find in their own mothers they were able to find
in their children, someone at their disposal who can be used as an echo, who can be controlled, is completely
centered on them, will never desert them, and offers full attention and admiration” (1981, p. 35).
Children’s instinct to bond with their parents drives them to do so regardless of the terms and conditions.
When such children look into their mother’s eyes and find no reflection but, rather, “the mother’s own
predicament,” they will mold themselves (if able) to their mothers’ psychic needs. Compulsive compliance—
initially adaptive in response to narcissistic or abusive caretakers—becomes maladaptive in the development of
the self and relationships with others (Crittenden & DiLalla, 1988).
Because he or she was completely helpless in childhood to resist the coercion of the parents’ unconscious,
Miller felt that the child within the adult patient always needs an advocate. Our brains are designed to adapt
to the environmental contingencies presented to us. Remember that the child’s first reality is the parents’
unconscious, transferred via right-hemisphere-to-right-hemisphere attunement well before self-awareness and
the emergence of a separate identity (Enlow et al., 2013). Because it is implanted in early implicit memory, it
is never experienced as anything other than the self. Miller was quick to describe the tragedy experienced by
parents who may be well aware of the pain from their own childhood and who may have vowed never to make
their children feel as they did. The intergenerational transmission continues because it is reflexive and
unconscious, and because each generation, at some level, protects the image of their parents and guards
against the pain of their own unfulfilled emotional needs.
Although patients do not generally have explicit memory for early relationships with their parents, Miller
posits that these learning experiences are implicitly recorded in how clients think of and treat themselves. The
strictness and negativity in the patients’ self-image and superego will expose their parents’ negative or
indifferent attitudes toward them from years before (Miller, 1983). These implicit emotional and behavioral
memories—in the form of attitudes, anxieties, and self-statements—contribute to the continued repression of
real emotions and needs.
Because children equate punishment with guilt, they can confuse being neglected or abused as evidence of
their innate badness. A teacher told Dr. Miller that after seeing a film about the Holocaust, several children in
her class said, “But the Jews must have been guilty or they wouldn’t have been punished like that” (Miller,
1983, p. 158). This assumption of guilt on the part of children both protects the parent and serves as the
developmental core of shame and a negative self-image. Because this self-image is organized and stored by
implicit systems of affective memory, the child’s later-developing identity forms around this unconscious,
negative core. Caretaking others and compulsive perfectionism reflect the ongoing attempt to compensate for
a feeling of unworthiness and the anticipation of more disappointment and abandonment.

René Magritte and His Mother


We must not fear daylight just because it almost always illuminates a miserable world.
—René Magritte

190
A chilling example of the reversal of the mirroring process is demonstrated in a painting by the surrealist René
Magritte, titled The Spirit of Geometry (see Figure 18.1). It depicts a mother holding a child, but with a
startling twist: their heads and faces have been exchanged. Magritte was the eldest of three boys in a middle-
class household in turn-of-the-century Belgium. His mother suffered from depression throughout his
childhood, and she made multiple suicide attempts. She was, in fact, locked in her room each night for her
own protection. As a young boy, René was locked in with her to keep her company. One cold February
morning, she managed to slip out of her room and drowned in the Sambre River. René was obviously not
successful in making his mother’s life worth living.

FIGURE 18.1
The Spirit of Geometry, René Magritte
Magritte’s image of a mother and her son may reflect an implicit emotional memory of his relationship with his mother

(© 2002 C. Herscovici, Brussels/Artists Rights Society (ARS) New York).

191
Based on his adult life, it is fair to assume that René was a bright and sensitive child and that he suffered
from a lack of positive maternal attention for most or all of his childhood. To lose his mother to suicide at the
age of 14 served as an additional blow to his sense of safety as a child. This painting suggests that the young
Magritte looked into his mother’s eyes and found fatigue, depression, and emptiness. In her eyes he read,
“You be the mommy—I’ll be the baby,” and he complied. In retrospect, a biographer suggested that Magritte
himself seemed at peace only when he was “tormented by problems” (Gablik, 1985).
Much of the body of Magritte’s surrealist work presents us with the message that the world is not what it
appears to be. His respectable middle-class family had a dark and painful secret at its core, one that became a
central aspect of the young boy’s experience. Although as an adult Magritte repeatedly stated that his early
experiences had no bearing on his artistic work, it is difficult to imagine that the loss, betrayal, and
abandonment experienced in his childhood did not reverberate in his many works, warning us not to be fooled

192
by the assumptions on which we depend.

Summary

The separation between the true and false selves reflects the brain’s ability to develop dissociated tracks of
experience. Early trauma and stress in nonhealing environments can even result in the formation of multiple
separate personalities, now referred to as dissociative identity disorder (DID). It is logical to assume that these
different experiential states are encoded within different patterns of neural network activation. The existence
of pathological caretaking, DID, and other disorders of the self demonstrate its fragility and its inclination to
adapt to whatever social realities are presented to it.
The purpose of both the brain and the self is survival. For each of us, the organization of our own self—
including our personality, defenses, coping styles, and the like—reflects the conditions to which we have had
to adapt. Put another way, all aspects of the self are forms of implicit memory stored in neural networks that
organize emotion, sensation, and behavior. These networks are sculpted in reaction to real or imagined threats
as the brain strives to predict and control its physical and social environment.

193
Part VI
The Reorganization of Experience

194
Chapter 19
The Evolutionary Necessity
of Psychotherapy
The useless, the odd, the peculiar, the incongruous—are the signs of history.
—S. J. Gould

The human brain is an amazing organ, capable of continual growth and lifelong adaptation to an ever-
changing array of challenges. Our understanding of how the brain accomplishes this mandate increases with
each new theoretical development and technological advance. At the same time, we are uncovering some of
natural selection’s more problematic choices. If necessity is the mother of invention, then evolution itself has
created the necessity for psychotherapy by shaping a brain that is vulnerable to a wide array of difficulties.
Over the last century, psychotherapists have demonstrated that many of the brain’s shortcomings can be
counterbalanced by the application of skillfully applied techniques in the context of a caring relationship.
Thus, in our ability to link, attune, and regulate each other’s brains, evolution has also provided us a way to
heal one another. Because we know that relationships are capable of building and rebuilding neural structures,
psychotherapy can now be understood as a neurobiological intervention with a deep evolutionary history. In
psychotherapy, we are tapping the same principles and processes available in every relationship that allow us to
connect to and heal another brain.
In this chapter, we focus on eight aspects of brain functioning that bring many people to psychotherapy:

1. The suppression of language and predictive capacity under stress


2. Divergent hemispheric processing
3. The bias toward early learning
4. The tenacity of fear
5. The damaging effects of stress hormones
6. The speed and amount of unconscious processing
7. The primacy of projection
8. Unconscious self-deception

This will serve as a review of multiple aspects of this book by drawing on many of the neuroscientific findings
previously discussed. We will then apply some of these principles to a case of PTSD.

The Suppression of Language and Predictive Capacity Under Stress


When a man’s knowledge is not in order, the more of it he has, the greater will be his confusion.
—Herbert Spencer

When animals hear a strange and potentially threatening sound, they freeze in their tracks, become silent, and
scan the environment for danger. The logic behind this primitive reflex is quite clear—being still and silent
make us less likely to be seen or heard by predators as we prepare to flight or flee. Research suggests that
during these states of high arousal, Broca’s area, responsible for speech production and other behaviors,
becomes inhibited.
While this response may be a positive adaptation in animals without language, it is a high price for
humans to pay for being frightened. Putting feelings into words and constructing narratives about our
experiences are integral to emotional regulation, associating neural networks of emotion and cognition, and
organizing coherent autobiographical memories. Most importantly, a lack of language can separate us from

195
the healing effects of positive connections with others. The loss of the ability to construct narratives is
especially problematic in situations where individuals are forced into silence by their abusers, or after enduring
the unspeakable horrors of torture, war, or the death of loved ones.
Losing the ability to verbalize feelings also interferes with building descending inhibitory cortical networks
down to the amygdala. You may have had a traumatic experience like an auto accident or robbery where you
felt compelled to tell the story over and over again in the following days and weeks. In time, the pressure to
tell others diminishes as the emotions connected to the traumatic events slowly dissipate. I suspect that telling
the story builds the circuitry, involved in amygdala inhibition and the dissipation of fear.
In addition to its role in language, Broca’s area also contributes to neural networks involved in sequencing,
prediction, anticipation, and action (Nishitani et al., 2005). Thus, along with loss of language, traumatized
individuals may also experience difficulties in the automatic navigation of day-to-day life that the rest of us
handle without a thought. This may be one of the reasons why traumatized individuals seem to experience
more than their share of accidents, bad relationships, and misfortune. The combined loss of words and
decreased predictive abilities enhance the long-term impact of the trauma by increasing the probability of
ongoing stress and revictimization.
The talking cure stimulates language networks and encourages the creation of adaptive narratives about
traumatic experience. The therapist’s caring presence, availability, and skill promote a moderate state of
arousal, which supports the neuroplastic processes necessary for building descending inhibitory fibers to limbic
and brain stem centers. Putting feelings into words also stimulates Broca’s area and supports the balance of
right- and left-hemisphere processing. Therapy activates Broca’s area, disinhibits language, restores predictive
abilities, and supports neuroplastic processes involved in adaptive learning.

Divergent Hemispheric Processing


The mind is the last to know.
—Michael Gazzaniga

Over the course of primate and human evolution, the left- and right-cerebral cortices have become specialized
to serve the formation of the conscious, linguistic self in the left and the somatic, emotional self in the right. It
is likely that as the hemispheres differentiated, it became increasingly necessary for one to take the lead in
executive control of conscious processing. There is considerable evidence that the left hemisphere took on this
function, along with specializing in maintaining moderate states of arousal and social connectedness. The
right hemisphere constantly provides information to the left, but while we are awake, the left hemisphere may
or may not allow this input into consciousness.
The disruption of a proper integration and balance between left and right hemisphere input can result in
dominance of one hemisphere or the other. As we have seen, too much inhibition of the right hemisphere by
the left can result in alexithymia, while too little inhibition can result in excessive emotionality, magical
thinking, or auditory hallucinations. Trauma increases the likelihood that the balance and coordination
between left and right will be disrupted. The intrusive symptoms of PTSD are clear examples of right-
hemisphere infiltration of the left. As we have also seen, the left and right prefrontal cortices are biased toward
positive and negative emotions, and a disturbance of the homeostatic balance of the two can result in extremes
of depression and mania.
When trauma occurs in early development, the hemispheres grow to be less coordinated and integrated,
resulting in problems in affective regulation and positive social awareness. People with histories of childhood
abuse and neglect have been shown to have a smaller corpus callosum and are more likely to suffer from
symptoms of PTSD (De Bellis et al., 1999; Teicher et al., 2004). Their brains have fewer connecting fibers
available to integrate right and left processes, and their development is characterized by decreased lateral
integration.
Therapists intuitively seek to balance the expression of affect and cognition. We encourage overintellectual
clients to be aware of and explore their feelings. On the other hand, we provide clients who are overwhelmed
by anxiety, fear, or depression with tools to use the cognitive capabilities of their left hemispheres to help
regulate these emotions. The common imbalance between affect and cognition in many of our clients rests, in
part, in the struggle between the hemispheres to integrate and learn a common language. The narratives we
create in therapy strive to be inclusive of the conscious and unconscious input from both hemispheres.

196
The Bias Toward Early Learning
In the practical use of our intellect, forgetting is as important as remembering.
—William James

At birth, the more primitive structures of our brains responsible for social and emotional processing are highly
developed, while the cortex develops slowly through the first decades of life. Much of our most important
emotional and interpersonal learning occurs during our first few years when our primitive brains are in
control. We mature into self-awareness having been programmed by early experience with sensory and
emotional assumptions that we unconsciously accept as truth. As a result, a great deal of extremely important
learning takes place before the development of autobiographical memory or self-awareness (Casey, Galvan, &
Hare, 2005). For most of us, the early interactions that shape our brains remain forever inaccessible to
conscious memory, reflection, or modification. This artifact of evolution, expressed in the sequential nature of
our neural and psychological development, turns the accidents of birth and the ups and downs of childhood
into the core of our identities and the causes for which we suffer and die.
Early experiences shape structures in ways that have a lifelong impact on three of our most vital spheres of
learning: attachment, emotional regulation, and self-esteem. These three areas establish our abilities to
connect with others, cope with stress, and feel we are wothwhile and loveable. Given how little control we
have over our early experience and that anyone can have children, an incredible amount of human brain
building is left to chance. In addition, we have seen that early experience shapes genetic expression, which
impacts our ability to learn, regulate our emotions, and nurture our future children.
It is obvious that our dependency on early caretakers can influence us in perfectly terrible ways. We see this
in abused and neglected children, who often enter adolescence and adulthood with explosive anger, eating
disorders, drug and alcohol problems, and other forms of destructive behavior. They may also have identity
disturbances and a poor self-image, exacerbated by angry feelings and antisocial behaviors. Like a veteran with
PTSD, these children have brains shaped to survive the combat of their day-to-day lives, but ill equipped to
navigate peace.
In psychotherapy, we have tools that allow us to explore early experiences with the possibility of coming to
understand our symptoms as forms of sensory, motor, and emotional memory. Projection, transference, self-
esteem, and internal self-talk are all expressions of early implicit memories from which we can get a potential
view of early unremembered interactions. Making the unconscious conscious is, in part, coming to an
awareness and understanding of the continuing impact of early experience. Once these forms of memory can
be consciously thought about and placed in a coherent narrative, we can reintegrate dissociated neural
networks of affect, cognition, abstract thinking, and bodily awareness. This process opens the door to
decreasing shame, increasing self-compassion, and the possibility of healing.

The Tenacity of Fear


It is the perpetual dread of fear, the fear of fear, that shapes the face of a brave man.
—Georges Bernanos

At birth, the amygdala is fully developed and functions as the executive center of our brains. Its first job is to
figure out who is safe and who is dangerous. Although the amygdala begrudgingly comes to share executive
control with the prefrontal cortex as our brains develop, it remains capable of hijacking the brain in states of
distress and fear. The amygdala’s job is to remember any and all threats and to generalize these experiences to
other signs of danger. In other words, the amygdala never forgets. Its tendency to generalize is why a panic
attack outside the home can lead to agoraphobia or why getting scratched by a cat can morph into a fear of all
furry animals. In contrast, the hippocampus is constantly remodeled in response to new information and can
easily differentiate one furry animal from another. Evolution has shaped our brains to err on the side of
caution and to be afraid whenever it might be remotely useful.
Fear makes our thinking and behavior more rigid. We become afraid of taking risks and learning new
things, which results in a tendency for those who are sick to remain sick. One recognized symptom of trauma
is neophobia, the fear of anything new. Once our brains have been shaped by fear to perceive, think, and act in
stereotyped ways, we tend to remain in rigid patterns that are reinforced by our very survival. Our internal
logic is self-perpetuating, making it difficult for us to find answers that are different from the ones we already
know. Although our best chance of learning rests in getting input from other people, fear may lead us to keep

197
them at arm’s length. When people are hurt or afraid, caring relationships are not easily entered into nor easy
to benefit from. Openness and trust are fragile creatures, even with the people we love most.
The training of the therapist and the therapeutic context are designed to enhance support and trust, and
provide consistent emotional availability. Within the consulting room, therapists attempt to be amygdala
whisperers and work to reactivate networks of new learning in the hippocampus and prefrontal cortex.
Warmth, empathic caring, and positive regard can create a state of mind that enhances neuroplastic processes
and increases the likelihood of positive change. It now appears that therapists help people get over their fears,
not by erasing traumatic memories but by building new connections to inhibit these amygdala-based
memories from triggering autonomic arousal.

The Damaging Effects of Stress Hormones


Man is an over-complicated organism. If he is doomed to extinction, he will die out for want of simplicity.
—Ezra Pound

For our more primitive cousins with smaller cortices and simpler environments, danger is frequently
encountered and quickly resolved. They either win or lose the fight, escape a predator, or become dinner. But
when you add a huge cortex capable of building freeways and information superhighways, the brain has to
adapt to a world of constant challenge where there is never any clear resolution. A large cortex also adds a
memory for the future with endless possibilities for anticipatory anxiety. Beyond all this, we now have a vast
capacity for creating frightening fantasies that our primitive brains are unable to distinguish from reality.
As we have already discussed, stressful situations trigger the release of the catabolic stress hormone
cortisol. Cortisol’s motto is “live for today, for tomorrow we may die.” As one of the glucocorticoids, one of its
job is to break down complex carbohydrates into usable energy for our muscles. Another job of cortisol is to
inhibit protein synthesis to conserve energy. Because neural growth and our immune system both depend on
protein synthesis, prolonged high levels of stress hormones impair our ability both to learn and to remain
healthy.
The conservation of our primitive stress system, well adapted for a simple life and a small cortex, can leave
our modern human brains bathed in chronically high levels of stress hormones. This leads to compromises of
brain maintenance, poor learning, and physical decline. Because chronic stress inhibits learning, successful
psychotherapy depends on our ability to downregulate stress in our clients. From specific stress reduction
techniques to the soothing effects of a supportive relationship, stress modulation and success in psychotherapy
go hand in hand. Thus, stress reduction skills should not be limited only to a few specific diagnoses, because
stress reduction is always necessary for successful psychotherapy.

The Speed and Amount of Unconscious Processing


Man is ready to die for an idea provided that idea is not quite clear to him.
—Paul Eldridge

In order to survive, animals have to be tough or fast; being a tortoise or a hare is an equally viable survival
strategy. While our elaborate cortices separate the human brain from theirs, farther down, all three are pretty
similar. Our expanded cortex does allow us vast response flexibility over our more primitive cousins. Of
course, thinking through options takes time and in some circumstances, a speedy reflex is far more adaptive.
Because of this, we have retained many primitive reflexes and subcortical executive control of certain functions
in the service of survival. Our harelike brains allow us to make rapid decisions and have useful knee-jerk
responses.
While it takes approximately 500–600 milliseconds for an experience to register in conscious awareness,
the amygdala can react to a potential threat in less than 50 milliseconds. This means that by the time we have
become consciously aware of a threatening situation, it has already been processed many times by our more
primitive neural networks, activating memories based on instinct and implicit memories organized by past
learning. This unconscious backdrop shapes the perception of what is being consciously attended to by
constructing our experiences of the present moment (Nomura et al., 2003; Wiens, 2006). When we become
aware of the outcome of this unconscious process, we experience it as if it is happening in the present, and as
if we are acting of our own free will and conscious deliberation. The illusion of free will has obvious survival

198
advantages, foremost of which is the ability to be confident and assertive in complex situations. The downside
of this strategy is when we become so sure of our personal beliefs that we are unable to consider alternatives.
Ninety percent of the input to the cerebral cortex comes from internal neural processing. This makes sense
for rapid appraisal and action based on past learning but also results in cognitive distortions that can keep us
frightened, withdrawn, and confused. Think of the veteran years after combat, who ducks when he hears a car
backfire or runs for cover as a news helicopter flies overhead. A person who experienced early abandonment
may, as an adult, be perfectly capable of starting new relationships. At a certain point, however, intimacy may
trigger implicit memories of insecure or disorganized attachment, leading him to become frightened and flee
from a potentially healthy relationship (Koukkou & Lehmann, 2006; Rholes, Paetzold, & Kohn, 2016). The
impulse to run, driven by implicit memories embedded in primitive brain circuitry, can be overpowering and
inescapable. The underlying reasons for these irrational behaviors based on implicit memories are described by
Christopher Bollas (1987) as the “unthought known.”
Openness to questioning one’s assumptions, especially when they are incorrect and self-defeating, is a key
predictor of positive outcome in psychotherapy. Once clients begin to understand that what they assumed to
be reality is actually a product of their brain and mind, they either flee or become fascinated. We attempt to
get them to question their thoughts, beliefs, and assumptions and to “act in,” that is, to come to sessions and
talk about their impulses with the hope of integrating inhibitory cortical input with primitive memories,
emotions, and urges. Attachment schemas, transference, and self-esteem are all examples of implicit memories
that shape and distort conscious awareness. These very patterns of unconscious processing led Freud to
develop a therapeutic context that supported an exploration of the unconscious. Psychotherapy encourages
being skeptical of the perceived realities of our brains, a sound strategy that we share with research scientists
and Buddhist monks.

The Primacy of Projection


It is easy to spot an informed man—his opinions are just like your own.
—Miguel De Unamuno

Human brains possess complex social networks that become activated as we observe and interact with those
around us (Cozolino, 2012). Hours after birth, we begin to focus on and imitate the facial expressions of our
caretakers. Mirror neurons begin to link observations and actions, allowing us to (1) learn from others by
watching them, (2) anticipate and predict actions, and (3) activate social brain networks responsible for
emotional resonance and empathy. We also possess neural circuits that automatically analyze the actions and
gestures of others, generating a theory of their mind—what they know, what their motivations may be, and
what they might do next. As with mirror neurons, having an automatic theory of what is on the mind of
another allows us to predict behavior to ensure our safety while supporting group cohesion and the spread of
culture.
These sophisticated social neural systems embody the millions of years of natural selection that have
shaped our brain’s ability to read the emotions, thoughts, and intentions of others. All of this attention to
others clearly shows us the importance of social information processing to our very survival. As a result, we are
quick to think that we know others because mind reading is instantaneous and obligatory. In biblical terms, it
is a reflexive habit of the social brain to attend to the mote in our brother’s eye and not to the beam in our
own. Until recently, evolution has not seen fit to invest much neural circuitry into self-awareness. Projection is
automatic and lessens anxiety while self-awareness requires effort and generates anxiety—which do you think
is going to be the norm?
Given that we use our internal expressions as implicit models for how we understand others, it could be
that what Freud called projection is actually a simple by-product of how our brains interweave our automatic
theories of others’ minds with our understandings of ourselves. This may be why we often discover our own
truths in what we think and feel about others, and since our individual identities emerge from dyads, perhaps
the separation of self and other (i.e., boundaries) is always a dicey distinction, one that many cultures do not
even bother with. Because self-reflection is never free from personal biases, self-analysis is generally not
successful. This is why the most naive observer can see many things about us that we cannot see in ourselves.
In close relationships, we provide each other with another set of eyes that can reflect our lives to us in ways
that are impossible to perceive on our own.
In therapy, we teach our clients to ask themselves if their thoughts and feelings about others can teach
them something about themselves. While in couples therapy, we encourage our clients to stop mind reading

199
and learn to actually ask their partners what is on their minds. We engage in a parallel process whenever we
explore how much of our reaction to a client is countertransferential. In our training as therapists, we learn to
question our judgment and assumptions in light of our own personal histories. We also learn to use mirror
neurons and theory of mind to enhance our attunement with our clients and explore their inner worlds.
Taking back our projections and working with transference and countertransference in therapy allows us to
use our thoughts about others as potential sources of self-insight.

Unconscious Self-Deception
Men use thoughts to justify their wrongdoings, and speech only to conceal their thoughts.
—Voltaire

Based both on neural architecture and everyday experience, self-insight does not appear to have exerted a
strong pressure on natural selection. In fact, it may have been selected against because real self-knowledge can
create hesitation, doubt, and demoralization. Defenses that distort our reality can help us by decreasing
anxiety, shame, and depression. At the same time, they can enhance social coherence by putting a positive
spin on the behavior of those closest to us. A wise judge once observed that everyone he sent to prison had “a
mother with an innocent child.” Self-deception not only decreases anxiety, it also increases the likelihood of
successfully deceiving others. If we believe our own distortions and untruths, we are less likely to give away our
real thoughts and intentions via nonverbal signs and behaviors. Reaction formation, or behaviors and feelings
that are opposite of our true desires, is also quite effective. Freud’s defense mechanisms and the attributional
biases of social psychology document a vast array of these distortions.
Secure attachment and ego strength are correlated with our ability to hear feedback, accept our own
limitations, and use less reality-distorting defenses—humor instead of repression and sublimation instead of
denial. In psychodynamic psychotherapy, we provide our clients not only with interpretations, clarifications,
and reflections, but also with an alternative perspective (our own) that they can utilize to help discover
themselves. This is why our own personal therapy is so important to our clients. In today’s large social groups,
self-awareness has become increasingly important for the survival of our species, perhaps as important as self-
deception is to the survival of the individual. Distorting reality to reduce one’s anxiety may allow the rich to
get richer and the righteous to feel justified in their beliefs, but in the long term there are consequences to the
environment and society that have negative impacts on us all.
I suspect that you have recognized many of your clients’ problems in these discussions and perhaps some of
your own. While the ways in which the brain and mind have evolved supported our survival, they have
simultaneously created a wide variety of threats to our emotional and physical well-being. These legacies of
evolution make both clients and therapists vulnerable to dissociation and distress and lead us to a common
search for solutions. Let’s take a look at how this can play out in life and in therapy.

Patrick
Man is born broken. He lives by mending.
—Eugene O’Neill

Patrick, a man in his mid-30s, came to therapy with the complaint that his life was “unraveling.” Talkative
and obviously bright, Patrick was the kind of person who could quickly launch into animated conversation. I
felt immediately drawn into his stories and found our sessions interesting and enjoyable. I soon learned that he
had grown up on the Connecticut coast, that his father was a fisherman with a “drinking problem,” and that
both his parents had died a few years earlier. A recent physical altercation during a relationship breakup
triggered him to take stock of his life. “That’s just not me,” he said. “Something’s very wrong.” Over our first
few sessions, two things stood out to me; he lived under a great deal of emotional pressure, and alcohol was
pervasive in his life.
Patrick’s initial interest was to understand what had happened in his last relationship and during the
breakup. Although he had the sense that some broader issues lay beneath the surface, he was unable to
articulate what they might be. “That’s your job,” he told me with a smile. It was clear from our discussions
that alcohol had been a part of Patrick’s life from the beginning. His father’s alcoholism had many negative
effects on him and everyone in his family. Both Patrick and his ex-girlfriend drank heavily, and alcohol was

200
always a backdrop to their disputes. It became very clear that both Patrick and his father used alcohol to cope
with anxiety, sadness, and loss.
What turned out to be my first important intervention was a matter-of-fact question about whether he felt
he had a drinking problem. Given his father’s addiction and all he had told me about his lifelong relationship
with drinking, I expected to hear a “Duh!” and some sympathy for being so slow witted. Instead, Patrick was
surprised and taken aback by my question. It seemed that he had never given any serious thought to being an
alcoholic himself. At first, I was surprised by his surprise and impressed by his brain’s ability to
compartmentalize different aspects of awareness. The power and persuasiveness of his early learning to
normalize alcohol and deny emotions created a dissociation enabling him to be unaware of the proverbial
elephant in the room.
Instead of being defensive, he began discussing his relationship with alcohol like an investigator looking
for clues to a crime. He spontaneously shared that many difficulties in his relationships seemed to occur while
he was drinking. Patrick could see that he used alcohol to deaden his emotions and that it caused him
difficulties in school, at work, and in all of his relationships. As he listened to the unfolding of his own
narrative, he gradually became aware of the fact that he shared his father’s addiction. By the expressions on his
face, it was clear that he was learning about himself by listening to his story—a story he had lived, but never
thought about. Patrick was beginning the process of using language to integrate networks of emotion and
thought, and reevaluating past experiences from a perspective of expanded self-awareness.
Once the reality of his own alcoholism was firmly established in conscious thought, he asked me what I
thought he should do. Again expecting a “Duh!” I said, “Well, you could stop drinking, join Alcoholics
Anonymous, and try working the steps.” And again, Patrick reacted as if this were a novel idea that he might
try out. Based on past experience, my expectation was that we would be working through his resistance for
quite a while before he would give AA a try. But, true to his word, he stopped drinking, got involved in AA,
found a sponsor, and worked the steps. (Note to beginning therapists—this sort of thing almost never
happens.)
The openness, enthusiasm, and rapidity with which Patrick gained awareness, explored, and tackled the
issue made me wonder what his life would have been like with even a modicum of insightful parental
guidance. His obvious intelligence and drive could have been applied to all kinds of challenges had he only
been taught a language for his feelings, learned to be self-aware, or applied his problem-solving skills to his
own life. His early chronic stress, combined with the family silence and shame, led him to inhibit language
and conscious awareness.
Like many children of alcoholics, Patrick was great at giving advice to others while remaining a complete
mystery to himself. His father’s drinking, financial difficulties in the family, and problems with his siblings
always took precedence over his personal needs. Almost no time was spent during his childhood on helping
him learn how to cope with negative emotions—a learned and not innate ability. Like so many, he grew up
alone, surrounded by friends and family, with only the ability to act out his inner emotional turmoil. His
chaotic inner emotional world as well as his coping strategies and defensive styles were established early in life
and continued into adulthood. The fear of both his internal chaos and of confronting it kept him drinking and
avoiding his emotions for decades.
Once sober, the emotions and defenses behind his addiction began to emerge. With the help of AA he
learned about the typical struggles and pitfalls of alcoholism. These meetings eventually led him to explore the
principles of Adult Children of Alcoholics, which proved to be of even more importance to him than the
impact of his own drinking. Over the following months, his impulsivity and impatience, pathological
caretaking, and inability to navigate negative emotions became the focus of therapy. Situation by situation, we
deconstructed his assumptions and searched for new ways of thinking, feeling, and behaving. In essence, we
were doing the work that should occur in childhood during countless interactions with parents. The tenacity
of fear is impressive, causing many of us to spend a lifetime avoiding the pain and confusion of childhood.
During our sessions, Patrick mentioned his work in the New York financial district during 2001. For some
reason, I had never thought to ask him how the terrorist attacks of 9/11 impacted him and, similarly, he never
brought it up. When I asked if and how he was affected, he offhandedly mentioned that he had lost many
friends but didn’t think he had suffered any ill effects. His tone of voice sounded hauntingly similar to when
he described his early family life, so I asked him to tell me about his experience of that horrible day. In
response he told me the following story:

I was living in New Jersey at the time and commuted to Wall Street via ferry across the Hudson. I overslept that morning, so instead of
being at work at 8:30, I was still crossing the river at 9. Halfway across, some of us noticed smoke rising from downtown. People began
gathering at the front of the ship to see what looked like a fire near the Trade Center. Minutes later, a southbound plane tracking the

201
Hudson passed low overhead. We all watched as it passed over the Statue of Liberty, banked hard left, and headed back toward
Manhattan.
As the second plane hit we were close enough to the docks to see people running through the park. The concussion shock from the
explosion made everyone move forward and back, like blades of grass in the wind. When the plane hit the second tower, we knew we
were under attack. I was enraged and anxious to reach the dock to see what I could do to help. I remembered my father talking about
Pearl Harbor and how he and his friends stood in line at the recruiting office the next morning.
As we approached the dock, the crowds rushed forward to escape the carnage and chaos behind them. But just before we reached the
dock, the ferry slowed and reversed direction to avoid being swamped by the massive crowd. As we headed back to New Jersey, I began
thinking about the people I knew who worked in the Trade Center and imagined what it would be like to be in those buildings as they
burned. I remember feeling numb as I stared at the huge plume of smoke rising above the city. When we got back to Jersey we all stood
in the terminal watching the television as the buildings collapsed, first one and then the other. Some of the women cried and even some
of the men; we were all in shock, leaning against one another for support.
The phones were down and there was nothing I could do but imagine where my friends were and if they had survived. Once I arrived
back home, I couldn’t take my eyes off the television. From time to time I would see someone I knew, dazed and covered with dust,
staggering past the camera. I eventually noticed that it had grown dark and that I was hungry. I was able to connect with my girlfriend
later that night, and the next day we drove south in a sort of waking coma. Eventually we found a hotel room and got drunk for a few
days.

Both Patrick’s reaction to the impact of 9/11 and his own alcoholism were processed in a disassociated
manner. This defense paralleled how he dealt with his experiences and feelings from childhood. Early in life
he learned to stay away from his emotions through denial, hypomanic activity, and combative sports.
Drinking, working, and socializing had all become compulsive ways of avoiding or inhibiting his feelings. His
intellect allowed him to succeed at work and rationalize his unhealthy behaviors. But upon awakening each
day, he discovered the same sad, frightening, and lonely world.
The many negative effects of having an alcoholic parent are well known. The fear and uncertainty created
by the alternating presence and absence of the alcoholic combined with an increased probability of extreme
emotional reactions can have a devastating impact on all aspects of development. These children often learn
not to be needy as they develop a false self to take care of the parent and maintain the artifice of a happy
home. Unfortunately, they never receive the help they need to articulate their experiences and develop a
comforting inner world. Their early terror persists in networks of implicit memory, but they lack mechanisms
for expression or social connection that could heal them. Thus, they re-create the chaos from which they are
so desperately trying to escape by acting out their emotions in relationships with traumatized partners.
At first it was difficult for Patrick to focus on his own feelings and personal issues. His initial attention
kept returning to the problems, shortcomings, and limitations of his girlfriend. It was only after he became
less agitated and more self-reflective that he could see that he and his girlfriend shared many of the same
problems. Like most of us, he found it easier to focus on her problems than his own. His gradual
improvement required that he reorient his focus onto his own behaviors and feelings. Next Patrick had to
develop a language with which to describe and share his inner world. And finally, over and over again, he had
to experience his life from the inside out and learn to connect with others as peers rather than as a caretaker or
a raging alcoholic.

Summary

The sophistication of the human brain reflects millions of years of evolutionary adaptation that conserved and
modified old structures while new structures emerged and expanded. Countless interactive networks and
design compromises created fertile ground for the disruption of smooth integration of neural systems. The
very complexity of the development and functioning of the brain is also what makes it such a fragile structure.
Assuming that the trillions of components arrive in their proper places and work according to their genetic
templates, there are a host of other challenges to integrated psychological functioning.
The discontinuity of conscious and unconscious processes, multiple memory systems, differences between
the hemispheres, hidden processing layers, and multiple executive structures are all potential sources of
dissociation and dysregulation. Disruption in the coordination and homeostatic balance of these neural
systems is the neurobiological substrate of psychological distress and mental illness.

202
Evolution is driven by the physical survival of the species and thus, much of the brain’s functioning is
centered around automatic fight-or-flight mechanisms as opposed to conscious and compassionate decision
making. Because of this, the conscious and unconscious management of fear and anxiety is a core component
of our personalities, attachment relationships, and identities. The considerable degree of postnatal brain
development and the disproportionate emphasis on early childhood experiences in the sculpting of the brain
add to our vulnerability to psychological distress.
Psychotherapists are trained to use their social brains as a tool to connect to and modify the brains of their
clients. Through interpersonal neurobiological processes, therapists serve as an external regulatory circuit to
help reestablish the optimal flow of energy and information. This is done for the circuits within and among
the cerebral hemispheres, and through all levels of the nervous system from the primitive lower regions to the
most recently evolved components of the neocortex. It is through this increased integration that more optimal
mental processing is established and symptoms are replaced with functional behavior.

203
Chapter 20
Stimulating Neural Plasticity
It is not the strongest of the species that survive, not the most intelligent, but the one most responsive to
change.
—Charles Darwin

Therapists often lament the fact that parents don’t do a better job during their children’s early development
when relationships have such a powerful impact on their brains. But if we approach the brain from the
perspective of ongoing plasticity, what are we capable of in our consulting rooms? How plastic is the brain?
Can plasticity be enhanced, and how much of an effect can the therapeutic relationship have on the brain?
These questions are central to psychotherapy because we rely on the brain’s ability to change well after
traditional sensitive periods have passed.
Psychotherapy came into being and survived for a century in the absence of a brain-based model of change.
The old view of the brain as a static entity necessitated that the development of psychotherapy be independent
from the biological sciences. Fortunately, current neuroscientific research discovered a brain capable of
constantly adapting to new challenges. In fact, therapists have learned to utilize attachment, emotional
attunement, and narratives as tools to modify the brain.
Research on imprinting of parental figures in geese (Lorenz, 1991) and the importance of periods of visual
exposure in the development of the occipital lobes in cats (Hubel & Wiesel, 1962) have been standard fare in
undergraduate education and popular science for half a century. Unfortunately, the popularity of these studies
left the impression that the timing of all brain growth is predetermined by template genetics that indelibly
etch early experiences into our neural architecture (Rutter & Rutter, 1993).
It turns out that extracting the concepts of imprinting and critical periods from ethology and applying
them to human development can be quite misleading (Michel & Moore, 1995; Roth & Sweatt, 2011). We
now know that genetic expression is controlled by experiences throughout life, that changes in the
environment, both good and bad, continue to have positive and negative effects on us and that our brains are
more plastic. Learning in humans is far more complex and flexible than the early findings about imprinting
and critical periods in less complex animals led scientists to believe (Hensch, 2004; Keverne, 2014).
Recall that sensitive periods are times of exuberant growth in neural networks, corresponding with the
rapid development of the skills and abilities to which they are dedicated (Chugani, 1998; Fischer, 1987).
Although there is no doubt that these impactful periods exist, the indelibility of learning remains in question.
As neuroscience finds more examples of neurogenesis, neuroplasticity, and ongoing epigenetic brain building,
there is an increasing recognition that we retain neural plasticity throughout life (Bornstein, 1989).
The research on neural plasticity has historically focused on the brain’s ability to adapt after early brain
injury (Goldman, 1971; Goldman & Galkin, 1978; Henry et al., 1984). Today, plasticity is understood to be a
basic principle of healthy brains at any age. Rather than lacking plasticity, the adult brain is now seen as
having an increased tendency toward neural stabilization while retaining the ability for new learning. There
also appears to be a shift in how information is processed in a manner more appropriate to different stages of
life, which is another aspect of its plastic reorganization (Cozolino, 2008; Stiles, 2000). This vital shift in
perspective has led to a rapid expansion of interest in and exploration of the neurobiological mechanisms of
lifelong learning and adaptation (Rosenzweig, 2001).

Use-Dependent Plasticity
Education is not a preparation for life; education is life itself.
—John Dewey

204
You will remember that changes in synaptic strength in response to an inner or outer stimulus are believed to
be the basis for learning. The process of long-term potentiation (LTP) prolongs excitation of cell assemblies
that are synchronized and interconnected in their firing patterns (Hebb, 1949). This is only a small piece of a
vastly complex set of mechanisms and interactions shaping the connection, timing, and organization of the
firing between the billions of individual neurons woven into neural networks.
Plasticity refers to the ability of neurons to change the way they relate to one another as we adapt to the
changing demands of life (Buonomano & Merzenich, 1998). This can occur in the modulation of signal
transmission across synapses, changes in the organization of local neural circuits, and in the relationship
between different functional neural networks (Trojan & Pokorny, 1999). It has been demonstrated that
portions of the cortex involved in sensory and motor functions reorganize in response to changing uses, after
injury, and during skill learning (Braun et al., 2000; Elbert et al., 1994; Karni et al., 1995). Violinists have
larger cortical representations in areas dedicated to the fingers of the left hand than do non–string players
(Elbert et al., 1995); Braille readers demonstrate similar patterns of cortical plasticity in sensory regions (Sterr
et al., 1998a, 1998b); and cab drivers have larger hippocampi for increased storage and retrieval of visual-
spatial information (Maguire et al., 2006).
These and other studies have discovered use-dependent plasticity in both cortical and subcortical regions.
Because of their early maturation and organization, sensory-motor areas have been thought to have the earliest
sensitive periods, and the most permanent and least flexible neural organization. However, the extensive
plasticity discovered in these regions suggests that executive and association areas of the cortex, characterized
by their adaptation to change, should demonstrate even more synaptogenesis, altered synaptic connections,
and potentially neurogenesis (Beatty, 2001; Dalla, Bangasser, Edgecomb, & Shors, 2007; Gould, Reeves,
Graziano, et al., 1999; Hodge & Boakye, 2001; Mateer & Kerns, 2000). In fact, research has found a
continual increase in white matter volume in the frontal and temporal lobes of males well into the fifth decade
of life (Bartzokis et al., 2001).
The activation and organization of the cortex appears capable of continual change, with expansion and
contraction of cortical representation alternating with varying amounts of stimulation and deprivation (Polley,
Chen-Bee, & Frostig, 1999). In other words, the brain is capable of far more and faster functional
reorganization than previously thought (Ramachandran, Rogers-Ramachandran, & Stewart, 1992). Our
ability to learn new skills and information throughout life is clear evidence for the existence of ongoing neural
plasticity. The study of the speed, degree, and nature of neural plasticity is a vast new scientific frontier, and
the potential to enhance plasticity has profound implications for neurosurgery, education, and psychotherapy
(Classen, Liepert, Wise, Hallett, & Cohen, 1998; Johansson, 2000).

Enhancing Plasticity
Life is growth. If we stop growing, technically and spiritually, we are as good as dead.
—Morihei Ueshiba

As we learn more about the biological mechanisms of sensitive periods, the possibility of controlling them
begins to emerge (Moriceau & Sullivan, 2004). What if neuroscientists could learn how to reinstate sensitive
periods in adults during psychotherapy? Huang and his colleagues (1999) found that the sensitive period of
the visual cortex was accelerated in certain genetically altered mice. It turns out that a genetic variation in
these mice results in the earlier secretion of brain-derived neurotrophic factor (BDNF), a neural growth
hormone. Could BDNF reestablish sensitive periods and be used to stimulate more exuberant learning at any
time during life?
Kang and Schuman (1995) found that BDNF and NT-3 (another neurotrophic factor) enhanced LTP
activity when introduced to the hippocampus of an adult rat. In a related study, it was found that a strain of
mice with higher levels of N-methyl-D-aspartate (NMDA) receptors had enhanced performance in learning
and memory tasks (Tang et al., 1999). NMDA, a neurotransmitter involved in the formation of connections
among neurons, is necessary for cortical reorganization and has the ability to specify certain transcriptional
processes related to neural plasticity (Jablonska et al., 1999; Rao & Finkbeiner, 2007; Wanisch, Tang,
Mederer, & Wotjak, 2005). NMDA receptors have also been shown to be necessary for the initiation of LTP
in monkeys (Myers et al., 2000).
This work has led to the suggestion that the use of D-cycloserine, which enhances the activation of
NMDA receptors, may boost the impact of psychotherapy. One study showed an improved effect of exposure
therapy with phobic individuals after using D-cycloserine (Ressler et al., 2004). The effects of cholinergic

205
stimulation suggest it also plays a role in neural plasticity by activating neural growth hormones (Cowan &
Kandel, 2001; Zhu & Waite, 1998). Future research with these biological subcomponents of learning and
memory may lead to pharmacological interventions that could enhance the brain’s ability to learn while taking
piano lessons, preparing for final exams, or during certain critical phases of psychotherapy (Davis, Myers,
Chhatwal, & Ressler, 2006).
Just a few years ago, the conventional wisdom in neuroscience was that we were born with all the neurons
we would ever have. More recent research has found an increasing number of areas within the brain where
new neurons are generated. Stem cells, capable of indefinitely renewing themselves, have been found in the
olfactory bulb and a portion of the hippocampus called the dentate gyrus (Jacobs et al., 2000). When the
biological processes that stimulate stem cells are understood, neurosurgeons may be able to trigger the growth
of new tissue in damaged areas (Hodge & Boakye, 2001).
The underlying biochemistry of neuronal growth is being investigated and may someday be utilized to
enhance and support plasticity and treat neurodegenerative disorders such as Alzheimer’s and Parkinson’s
diseases (Akaneya, Tsumoto, Kinoshita, & Hatanaka, 1997; Barde, 1989; Carswell, 1993). The acceleration
or reestablishment of sensitive periods, the enhancement of learning and memory via biochemical
adjustments, and the cultivation of new cells are all relevant to the potential therapeutic use of neural
plasticity. While future biological interventions are causes for hope, we are already capable of enhancing
plasticity through our day-to-day behaviors and interactions.

Enriched Environments and Stimulating Lives


Intellectual growth should commence at birth and cease only at death.
—Albert Einstein

It has been known for decades that enriched and stimulating early environments can have a positive long-term
impact on both neural architecture and neurochemistry. Research has demonstrated that when rats are raised
in complex and challenging environments, they show advances in many aspects of brain building (see Table
20.1). An enriching environment enables mammals to build larger, more complex, and more resilient brains.
The effects of the environment in stimulating brain growth are so robust that it occurs even in situations of
malnutrition. If rats are malnourished but placed in enriched environments, they will have heavier brains than
well-fed but less-stimulated rats. These findings exist despite the fact that the bodies of malnourished rats
weigh significantly less (Bhide & Bedi, 1982). Although nutritional and environmental deprivation often go
hand in hand, some deficits may be reversed in adulthood. Placing adult rats in enriched environments
enhances synaptic plasticity and ameliorates the effects of earlier nervous system damage and genetically based
learning deficits (Altman, Wallace, Anderson, & Das, 1968; Kolb & Gibb, 1991; Maccari et al., 1995;
Morley-Fletcher, Rea, Maccari, & Laviola, 2003; Schrott et al., 1992; Schrott, 1997).

TABLE 20.1
The Impact of Enriched Environments in Experimental Animals

Increases In
Weight and thickness of cortex1 Levels of neurotransmitters7

Weight and thickness of hippocampi2 Level of vascular activity8

Length of neuronal dendrites3 Level of metabolism9

Synapses among neurons4 Amount of gene expression10

Activity of glial cells5 Levels of nerve growth factor11


Levels of neural growth hormones6

Although controlled studies of nutritional and environmental deprivation are not possible with humans,
some naturally occurring situations offer similar insights into the power of environmental enrichment. A study
of Korean children adopted by families in the United States found that environmental enrichment
counteracted their early malnutrition and deprivation. On measures of height and weight, these children
eventually surpassed Korean averages, while their IQ scores reached or exceeded averages for American
children (Winick, Katchadurian, & Harris, 1975). In a different study of older adults, a positive correlation

206
was found between the length of dendrites in Wernicke’s area and the subjects’ level of education (Jacobs et
al., 1993). It has also been suggested that an enriched environment may improve poststroke recovery in
humans (Ulrich, 1984).
The theory of cognitive reserve, which arose from these observations, has triggered research comparing the
brains and cognitive functioning in later life of people with varying levels of challenge and stimulation. The
cognitive reserve hypothesis suggests that stimulating lives build more neural material, and the more you
build, the more you can afford to lose and still function in a competent manner later in life (Richards &
Deary, 2005; Stern, Alexander, Prohovnik, & Mayeux, 1992). A number of these studies support the idea that
those who have had more education and challenging occupations tend to have brains that age better and resist
the onset and progression of dementia.
It is believed that the cognitive declines associated with normal aging are related to the gradual
degeneration of dendrites, neurons, and the biochemical mechanisms that support plasticity and neural health
(Jacobs, Driscoll, & Schall, 1997; Morrison & Hof, 2003). People with more cognitive reserve typically have
had better diets, higher-quality educations, and more intellectually challenging jobs than those with lower
reserve (Stern et al., 2005; Whalley, Deary, Appleton, & Starr, 2004). Factors like larger brain size, early
learning, and greater occupational attainment are associated with greater cognitive reserve and seem to
mitigate the effects of Alzheimer’s disease, traumatic injury, and the general impact of brain aging (Compton,
Bachman, Brand, & Avet, 2000; Kessler et al., 2003; Scarmeas et al., 2004; Schmand et al., 1997; Staff,
Murray, Deary, & Whalley, 2004; Stern et al., 1995).
Skills like verbal fluency, controlled processing, and abstract thinking, demanded by high-complexity
occupations, appear to contribute most to cognitive reserve (Ardila, Ostrosky-Solis, Rosselli, & Gomez, 2000;
Le Carret et al., 2003). So, while being a college professor does not protect against cognitive decline, it may
slow down some of its manifestations (Christensen, Henderson, Griffiths, & Levings, 1997). About 25% of
older individuals showing no symptoms of Alzheimer’s disease while alive show significant Alzheimer’s-
related brain pathology upon autopsy (Ince, 2001; Katzman et al., 1989). Those with more education had a
significantly greater amount of plaques and tangles, yet they functioned as well as others with less advanced
disease (Alexander et al., 1997). This suggests that individuals with more education can sustain a greater
amount of neural damage, and still maintain the same level of cognitive functioning as those with less
education.
Studies have also found that expected age-related intellectual decline can be halted or reversed in many
older adults by increasing environmental and social stimulation (Schaie & Willis, 1986). The most probable
explanation would be that these experiences correlate with biological processes that enhance plasticity,
creating more elaborate, complex, and flexible brains. Given that psychotherapy is an enriched environment
for social-emotional learning, we can assume that the challenges that we provide our clients build more
complex and resilient brains. Research has yet to explore the possibilities of enhanced longevity and brain
health in those who engage in psychotherapy.

Moderate States of Arousal


I am always doing that which I cannot do, in order that I may learn how to do it.
—Pablo Picasso

Something we do in all forms of psychotherapy that enhances neural plasticity is to focus on modulating stress
in our clients. Regulating subjective units of distress in systematic desensitization, balancing confrontation and
empathic attunement in psychoanalysis, or crafting the safe emergency of Gestalt therapy all reflect an
appreciation of the delicate balance between challenge and support. We intuitively understand that people
need to be motivated and aroused to learn, while simultaneously free from the autonomic activation of a fight-
flight response that shuts down cortical plasticity. Knowing how to find and keep clients in learning’s sweet
spot is a vital element of the art of psychotherapy.
The nontechnical story goes something like this: When all is well and we are in a state of calm, there is no
reason to learn anything new. When our needs for food, companionship, and safety are satisfied, the brain has
done its job and there is little reason to invest energy in learning. At the other extreme, states of high arousal
and danger are no time for new cortical learning but a call for immediate action. A state of mind somewhere
between the two appears optimal for new learning and complex problem solving (Anderson, 1976). An
attitude best described as interest, enthusiasm, or curiosity stimulates positive arousal. In secure attachment,
the child is able to use the parent as a safe haven and avoid experiencing autonomic activation in response to

207
stress. A similar process is likely to take place in a secure therapeutic alliance, allowing clients to verbalize their
experience rather than defending themselves or fleeing from the situation.
The formalization of this notion in experimental psychology, described in a classic paper by Robert Yerkes
and John Dodson (1908), came to be known as the inverted-U learning curve. They found that mice learned
to avoid a moderate shock faster than one of high or low intensity. They charted their findings on a graph
with arousal on the x-axis and learning (performance) on the y-axis (see Figure 20.1). Over the years the same
phenomenon was found across species in a variety of learning tasks (Broadhurst, 1957; Stennet, 1957). While
this research took place before we knew anything about the neurochemistry of learning, it makes sense that
this inverted-U pattern will be reflected in the underlying neurobiological processes of learning (Baldi &
Bucherelli, 2005).
We now know that learning depends on building new dendritic structures, and on the ability of these
structures to interconnect. Building dendrites is dependent on the synthesis of proteins guided by genetic
transcription, while neuronal firing shapes the architecture of these dendrites to encode new learning. Not
surprisingly, the hormones secreted by the HPA axis in response to stress (cortisol, norepinephrine, and
endorphins), as well as other learning mechanisms (NMDA receptors, BDNF secretion, etc.), modulate
learning on the same inverted U-shaped curve (Hardingham & Bading, 2003; Parsons, Stöffler, & Danysz,
2007). At moderate states of arousal, amygdala activation opens windows of modulated stimulation where
top-down plasticity is facilitated (Popescu, Saghyan, & Paré, 2007). At moderate levels of arousal, these and
many more systems enhance learning, while at high and low levels of arousal, learning is inhibited.

FIGURE 20.1
The Inverted-U Learning Curve

Hippocampal neurons require low levels of cortisol for structural maintenance, while higher levels of
cortisol inhibit their neuroplastic properties and can even kill them (Gould, Woolley, & McEwen, 1990).
Cortisol impacts learning and plasticity by regulating the protein synthesis required for dendritic growth and
patterns of neural connectivity such as LTP, long-term depression, and primed burst potentiation (a low-
threshold version of LTP) in the same pattern (Diamond, Bennett, Fleshner, & Rose, 1992; Domes,
Rothfischer, Reichwald, & Hartzinger, 2005; Lupien & McEwen, 1997; Roozendaal, 2000). High levels of
stress also trigger endorphin release, which impedes both protein synthesis and the consolidation of explicit
memory (Introini-Collison & McGaugh, 1987). See Table 20.2 for a sample of findings supporting this
pattern of the biochemistry of arousal and its effects on learning.
Overall, the various neural systems dedicated to learning and arousal are tightly interwoven. They work

208
together to activate plasticity and learning when it is needed to adapt to challenge, and turn it off in the
absence of challenge or when the body needs to be mobilized for immediate survival. The inverted-U learning
curve reflects both the underlying neurobiological processes and the many overt manifestations of learning
seen in the laboratory, classroom, and consulting office. By understanding these principles, we can use them to
optimize neural plasticity in the service of positive changes in the brain.

TABLE 20.2
The Inverted-U Curve of Learning and Arousal

mRNA expression1
Cortisol levels
Verbal memory2
Social memory3
Spatial memory4
Hip prime burst potentials5
Long-term potentiation6

Norepinephrine levels
Olfactory plasticity7

Endorphin levels
Protein synthesis and memory consolidation8

Attachment Plasticity
Organic matter, especially nervous tissue, seems endowed with a very extraordinary degree of plasticity.
—William James

Attachment in infancy is usually conceptualized as relationship specific, while attachment in adulthood is


thought of as a general aspect of character. Early attachment patterns may become generalized and self-
perpetuating because of their impact on our neurobiology, our ability to regulate our emotions, and the
expectations we have of ourselves, others, and the world. The sensory, emotional, and behavioral systems
influenced by early attachment experiences can shape our brains in ways that make the past a model for
creating the future.
We have learned a great deal from attachment research: how our primitive bonding instincts regulate
anxiety, how categories of attachment reflect a parent’s behavior and a child’s reaction to stress, and how
attachment schemas are carried into adulthood, affecting our choice of partners, the nature of our
relationships, and the way we parent our own children. Our enthusiasm about the power of attachment
categories to explain emotional development often fails to acknowledge the considerable fluctuation of
attachment schemas. In our search for a straightforward explanation of intimate human relationships, it is easy
to overlook its variability over time and across relationships.
In an attempt to support the predictive power of attachment schemas, fluctuation and change are
attributed to measurement problems or uncontrolled external variables such as positive and negative life events
or changing relationships. However, we seldom consider that an attachment schema should change because
we think of it as hard-wired. As we saw in the animal research, maternal attention to rat pups was mediated
by the environment in ways that made the pups a better fit to the environment to which the mother was
responding. Perhaps, because our brains are so much more complex, it takes greater effort and time for us to
change, but the role of attachment may be essentially the same in humans.
In human subjects, we see that a decrease in security or the maintenance of insecurity during adolescence is
more likely to occur in the presence of psychological, familial, or environmental stressors. Adolescents who see
their mothers as supportive are more likely to gain security, while maternal depression correlates with a shift
from secure to insecure attachment (Allen, McElhaney, Kuperminc, & Jodl, 2004; Hamilton, 2000;
Weinfield, Sroufe, & Egeland, 2000). Overall, secure attachment appears more resistant to change, while
negative life events tend to maintain insecure attachment (Hamilton, 2000; Kirkpatrick & Davis, 1994;
Thompson, 1982). These findings certainly parallel the animal literature.
The major implication for psychotherapy from all of these findings is that insecure attachment is subject to
change as a result of positive social input (Pilowsky et al., 2008). As shown in Table 20.3, the consistency of

209
attachment ratings, primarily secure versus insecure, is 24–64% depending on the study. This may be bad
news for those interested in attachment as a stable trait, but good news for those of us in the business of
change.
As a psychotherapist who is very interested in positive change, the messiness of attachment is welcome. I
want attachment schemas to be a malleable form of implicit memory, so relationships with clients can alter
them in salubrious ways. This allows psychotherapy to be a guided relationship that regulates of anxiety and
fear for the purpose of creating secure attachment schemas (Amini et al., 1996; Cappas, Andres-Hyman, &
Davidson, 2005; Corrigan, 2004; Siegel, 1999).

TABLE 20.3
Attachment Plasticity

Time Span Percentage of Subjects with Same Attachment Source


Classification
6 months during infancy 62 (N=100) Vaughn et al., 1979
7 months during infancy 53 (N=43) Thompson, Lamb, &
Estes, 1982
2 years during infancy 60 (N=189) Egeland & Farber, 1984
3.5 years during childhood 24 (N=223) Vondra et al., 2001
Childhood to adolescence 63 (N=30) Hamilton, 2000
Childhood to adolescence 42 (N=84) Lewis, Feiring, &
Rosenthal, 2000
Childhood to adolescence 39 (N=57) Weinfield et al., 2000
Childhood to adulthood 64 (N=50) Waters et al., 2000
17 weeks during adulthood 55 (N=33) Lawson et al., 2006
20 sessions of adult therapy 34 (N=29) Travis et al., 2001

These studies reflect an array of ages, target populations, and methods of


assessing attachment.

Although there is evidence of organized attachment schemas by our first birthday, they do not appear to be
set in neural stone. These naturally occurring changes and the fact that we attach and reattach with many
people throughout our lives suggests that the underlying neural systems maintain their plasticity. If you doubt
this, ask grandparents whether they feel attached to their grandchildren. In support of the neuroplasticity of
attachment networks, research suggests that adults can create secure attachment for their children despite
negative experiences in their own childhood.
Thus, the powerful shaping experiences of childhood can be modified through personal relationships,
psychotherapy, and increased self-awareness. The ability to consciously process stressful and traumatic life
events appears to correlate with more secure attachment, flexible affect regulation, and an increased availability
of narrative memory. The integration of neural circuitry across cognitive, behavioral, sensory, and emotional
domains is the likely neuroanatomical substrate of this earned autonomy. A healing relationship with a secure
partner or with a good-enough therapist, in which past pain can be processed and resolved, supports neural
integration and the attainment of a secure attachment schema (Simpson & Overall, 2014).

The Power of a Healing Relationship


Tell me who admires and loves you, and I will tell you who you are.
—Antione de Saint-Exupery

Because our brains are social organs and interwoven with the brains of those around us, relationships have the
power to heal us. Psychotherapy is successful in large part because of the therapist’s empathic abilities and the
client’s belief in the therapist’s humanity and skill. Through the instillation of optimism and hope, a therapist
(or any healer) uses powerful mechanisms of mind to reshape the brain. The mind’s impact on the brain has
been studied across a variety of fields from social neuroscience to psychoneuroimmunology, and is expressed in
such phenomena as expectancy effects and the power of placebos and suggestion. These are all examples of the
mind leading the brain via social relationships.
The term placebo, Latin for “I shall please,” reflects the ancient idea that the positive response to an
inactive treatment results from the patient’s desire to live up to the doctor’s expectations. More generally, the
placebo effect is the expectation for illness outcome based on what the patient is led to believe by the doctor.

210
Moerman and Jonas (2002) suggested renaming it “the meaning response,” reflecting of the fact that the
placebo effect is mediated via the meaning assigned within the mind of the patient.
Until modern times, shamans and witch doctors understood and utilized suggestion of all kinds in their
role as tribal healers (Frank, 1963). Similarly, the expectancy effect is likely at play in the fact that ill Chinese
Americans born during inauspicious years have significantly shorter life spans, with the size of the effect
related to how strongly they adhere to traditional cultural beliefs (Phillips, Ruth, & Wagner, 1993). Placebo
effects are seen as a nuisance in Western scientific medicine, and the patients affected by them are
pathologized as weak minded, impressionable, or malingering.
Modern technological medicine has led to a decline in the use of the meaning response as the role of the
doctor has shifted from healer to technician. This is especially ironic given that placebo control groups are a
staple of the research upon which their clinical decision making should be based. Don’t the time, money, and
effort expended to ensure this standard of research support the very power of patient expectancy to impact
symptom expression? Perhaps doctors have become too impressed with technology and have forgotten the
power of their humanity.
In psychotherapy outcome research, expectancy and placebo effects are relegated to the wastebasket
category of nonspecific effects. I would argue that these effects are in fact specific—that is, we know exactly
what social and psychological factors lead to healing. Carl Rogers outlined them well during the 1960s as
warmth, acceptance, caring, and unconditional positive regard. Epidemiologists call them social support,
network connectivity, and spiritual beliefs.
Different activation patterns in the brain during successful placebo response reflect the multiple neural
pathways involved in mind-brain regulation. In neuroscience terms, the placebo effect is an example of top-
down cortical modulation of mood, emotion, and immune activity (Beauregard, 2007; Ocshner et al., 2004).
The placebo effect relies heavily upon the prefrontal lobes to integrate social experiences with positive affect
and an optimistic state of mind. In the same way that a mother can shape a child’s brain by stroking his hair
and telling him that things will get better, a doctor can influence a client’s immune system by presenting an
optimistic prognosis and projecting confidence about the proposed treatment.
The placebo effect is a social phenomenon that likely activates the same reward systems (dopamine-
serotonin-endorphin) triggered by loving touch and the anticipation of positive connections and feelings
(Esch & Stefano, 2005; Fricchione & Stefano, 2005). The amygdala, which regulates our experience of fear
and pain, is inhibited by positive emotions and activated by negative emotions (Neugebauer et al., 2004). We
know, for example, that placebo and opioid analgesia share a common neural pathway, which strongly
suggests we are capable of opioid self-administration (Pariente, White, Frackowiak, & Lewith, 2005).
There are many examples of the social, top-down regulation of emotion. When a woman holds her
husband’s hand in the face of threat, fear activation is attenuated. Not surprisingly, the better she feels about
the relationship, the greater the soothing effect of his hand will be (Coan, Schaefer, & Davidson, 2006).
Soothing touch is obviously powerful, but so too is a soothing facial expression or a kind word communicated
in an attuned emotional state. Some evidence for this has been found when levels of arousal (as measured by
skin conductance) were monitored simultaneously in clients and therapists during sessions. During states of
matched arousal, there were significantly more positive social and emotional interactions between client and
therapist than when they were out of sync (Marci, Ham, Moran, & Orr, 2007). Even in Parkinson’s disease, it
is believed that the anticipation of positive reward stimulates the release of dopamine from the nucleus
accumbens. Because dopamine depletion is the central cause of Parkinsonian symptoms, this release of
dopamine results in symptomatic improvement. See Table 20.4 for an outline of studies that have
demonstrated changes in neural activation related to the placebo effect.
When physicians use the placebo effect to enhance illness outcome, they are using strategies that parallel
the core principles of client-centered therapy. For example, an article in the Journal of Family Practice
suggested a “sustained partnership” with patients characterized by an interest in the whole person, knowing
the patient over time, caring, sensitivity, empathy, being viewed as reliable and trustworthy, adapting medical
goals to the needs of the patients, and encouraging their full participation (Brody, 2000). Another physician in
a piece from the British Medical Journal asked, “For a thousand years the action of the placebo has made vast
numbers of patients feel better. Have we today provided a consultation in which the placebo does not act?”
(Thomas, 1987, p. 1202). These are the aspects of traditional healing still employed by psychotherapy that
appear to have been lost in the morass of modern technological medicine.

TABLE 20.4
The Impact of Positive Expectancy (Placebo) on Brain Activation

Major Depression

211
Increased activity in prefrontal, anterior cingulate, premotor, parietal, posterior insula, and posterior cingulate1

Decreased activity in subgenual cingulate, parahippocampus, and thalamus2

Pain Disorders
Increased activity in lateral and orbital PFC, anterior cingulate, cerebellum, right fusiform gyrus, parahippocampus, and pons3
Increased activity in the right dorsolateral cortex, anterior cingulate cortex, and midbrain4
Increased bilateral orbitofrontal activity and anterior cingulate cortex contralateral to pain stimulus5
Reduced activity in anterior cingulate cortex, anterior insula, and thalamus6
Increased activity in the prefrontal cortex7
Activated endogenous opioids in dlpfc, anterior insula, and nucleus accumbens8

Parkinson’s Disease
Reduced activity in subthalamic nuclei9
Increase in striatal dopamine10

The implications of the power of the placebo effect to change the brain are profound. Yet our awe of
science leads us to underestimate the importance of the doctor’s role as healer, and patients’ abilities to
contribute to their own healing through the modulation of their endogenous biochemistry. There are many
ways in which the power of the placebo effect could be harnessed and used to increase the effectiveness of
medical treatment including increasing visits, encouraging and supportive interactions, and shaping patients’
experiences by telling them what positive outcomes they should expect (Walach & Jonas, 2004). See Table
20.5 for a list of their suggestions. Ironically, we are only required to tell clients of the many risks and
potential negative side effects of treatment, which likely create negative expectancies and self-fulfilling
prophesies.

TABLE 20.5
Leveraging the Placebo Effect in Medical Treatment

Increase frequency of contact


Determine what treatments your patient believes in
Align beliefs between patient, doctor, family, and culture
Be sure you believe in the treatment you are administering
Inform patients about what they can expect
Deliver treatment in a warm and caring way
Listen and provide empathy and understanding
Touch the patient

Findings from the medical research supporting the power of specific doctor-patient interactions.
Adapted from Walach and Jonas (2004).

All of the “givens” of classical healing and psychotherapy have been squeezed out of modern Western
medicine. Perhaps these nonspecific ingredients of a healing relationship will someday be woven back into
medical practice.

Summary

The power of psychotherapy to change the brain rests in our ability to recognize and alter unintegrated or
dysregulated neural networks. As knowledge of neural plasticity and neurogenesis increases, so will our ability
to impact and alter the brain. The possibility exists that sensitive periods can be reinstated in the context of
psychotherapy, and that stress can be utilized in a controlled manner to “edit and re-edit” emotional memories
(Post et al., 1998). Although the practical application of these principles to humans remains on the horizon,
the ability of psychotherapy to sculpt the brain is evident. It is certain that psychotherapists are already
enhancing plasticity without the help of genetic manipulation or chemical interventions.

212
Therapy is a safe emergency because of the supportive structure in which difficult emotional learning takes
place. A client’s sense of safety is enhanced by the therapist’s skill, knowledge, and confidence, which support
emotional regulation and maintaining moderate states of arousal. It is quite possible that the caring,
encouragement, and enthusiasm of the therapist also reinforce learning, neural growth, and plasticity through
the enhanced production of dopamine, serotonin, and other neurochemicals (Barad, 2000; Kilgard &
Merzenich, 1998; Kirkwood, Rozas, Kirkwood, Perez, & Bear, 1999). Successful therapeutic techniques may
be successful because of their very ability to change brain chemistry in a manner that enhances neural
plasticity.
In the recent shift to neural optimism, critical periods of neural development are no longer seen as the final
word on neural structure. The impact of enriched environments has demonstrated the brain-building capacity
of positive experiences throughout the life span. More recent research in neural plasticity (e.g., use-dependent
plasticity, neurotransmitter alteration, stem cell implantation) suggests that new experiences, and future
potential biological interventions, may be capable of providing us with many tools with which to rebuild the
brain. Psychotherapy is on the verge of an exploding new paradigm: the psychotherapist as neuroscientist.

213
Chapter 21
The Psychotherapist as Neuroscientist
In this field we are merely at the foothills of an enormous mountain range . . . unlike other areas of science, it
is still possible for an individual or small group to make important contributions.
—Eric Kandel

Psychotherapists are applied neuroscientists who create individually tailored enriched learning environments
designed to enhance brain functioning and mental health. We are skilled at teaching clients to become aware
of unconscious processing, take ownership of their projections, and risk confronting anxiety in the service of
emotional maturation (Holtforth, Grawe, Egger, & Barking, 2005). In our work, illusions, distortions, and
defenses are exposed, explored, tested, and modified in the service of enhanced adaptation. Implicit memory
—in the form of attachment schemas, transference, and superego—are made conscious and explained as
expressions of early learning experiences. We use a combination of emotional attunement, empathy, ideas,
stories, and behavioral experiments to promote neural network growth and integration.
Through all of this work, more primitive subcortical and right-hemisphere networks that store memories
of fears, phobias, and traumas are activated and made accessible for integration with regulatory cortical
circuitry. This allows for enhanced information flow between explicit and implicit memory and top-down
regulation of painful memories stored as sensations and emotions. Regardless of the client’s particular problem
or the orientation of the therapist, psychotherapy teaches different methods to help better understand and
influence our brains. As the dialogue between psychotherapy and neuroscience continues to evolve, an
increasing number of new scientific findings will be applied to theory and practice.
Key to the success of the therapeutic process is an empathic and supportive relationship, the maintenance
of moderate states of arousal, activation of both cognition and emotion, and the coconstruction of more
adaptive narratives. A safe and empathic relationship establishes both the emotional and neurobiological
contexts most conducive to neural plasticity. It also serves as a scaffold within which a client can better tolerate
the stress required for neural reorganization. We have already seen that birds are able to learn, their songs
after sensitive periods when exposed to other birds singing, but are unable to learn the same songs heard from
a tape recorder (Baptista & Petrinovich, 1986). Under certain conditions, birds require positive social
interactions and nurturance in order to learn both from other birds and human trainers (Eales, 1985;
Pepperberg, 2008). The nearly insatiable drive of children to be with their mothers, and adolescents to be in
constant contact with their peers, reflects the powerful drive for interpersonal stimulation during sensitive
periods of socialization. Emotional expression and modulation are key elements of psychotherapy because they
leverage these evolutionary drives in the service of plasticity, learning, and adaptation.
The importance of the activation of both emotion and cognition for healing is recognized by most
psychotherapists. Releasing emotions associated with painful memories, facing a feared situation, or
experimenting with new interpersonal relationships all involve some sort of stress, anxiety, or fear. Although
this way of thinking has been accepted clinically, we now have considerable evidence to support the idea that
moderate levels of arousal optimize plasticity via the production of neurotransmitters and neural growth
hormones that enhance LTP, learning, and cortical reorganization (Cowan & Kandel, 2001; Zhu & Waite,
1998).
Trauma changes us in many ways that subsequently make life more challenging. Changes in biochemistry
lead to problems with attention, concentration, and learning. Inhibition of the default mode network leads to
disturbances in interpersonal attachments and self-identity. Dissociation in reaction to trauma represents a
breakdown of neural integration, plasticity, and our ability to update memory. In therapy, we use moderate
levels of arousal to access cortical mechanisms of plasticity in controlled ways with specific goals. The safe
emergency of therapy provides both the psychological support and the biological regulation necessary for
reregulation of neural chemistry and rebuilding the brain. Much of this neural integration and reorganization
most likely takes place in the association areas of the frontal, temporal, and parietal lobes, which serve to
coordinate, regulate, and direct top-down circuits of memory and emotion.

214
The importance of the coconstruction of narratives is grounded in the coevolution of the cerebral cortex,
sociality, and language. Language communication within significant relationships has shaped the brain during
evolution and given rise to self-awareness and what we think of as the human mind. Because of this
coevolution, shared narratives embedded within emotionally meaningful relationships are capable of
resculpting the brain’s neural networks. Constructing and editing autobiographical memory can bridge
processing from previously dissociated neural networks into a more cohesive story of the self. Narratives allow
us to combine—in conscious memory—our knowledge, sensations, feelings, and behaviors in a coherent
manner that supports neural network integration.
The coconstruction of narratives with parents serves as a medium of transfer of the internal world of the
parent to the child, from generation to generation. These narratives reflect the implicit values, problem-
solving strategies, and worldviews of the parents—both adaptive and maladaptive. They also serve to define us
to ourselves and others, and guide us through our complex social world. Research in attachment has
demonstrated that the coherence and inclusiveness of narratives correlate with both attachment security and
self-reflective capacity (Main, 1993; Fonagy, Gergely, Jurist, & Target, 2002).
Within the interface of social and neural evolution, different levels of language have emerged that parallel
various modes of consciousness:

1. A reflexive social language (of the left-hemisphere interpreter) that allows us to spontaneously speak to
others in socially acceptable ways. This language serves the purpose of creating a logically cohesive and
positive presentation to others. This language evolved from grooming and hand gestures with the
primary goals of group affiliation and coordination.
2. A reflexive internal narrator that comments on our behavior driven by implicit memories of early
emotional learning. These voices are often critical and shaming and are most likely a holdover from
earlier stages of evolution as a form of social control of alphas over betas.
3. A reflective internal language directed to the outer world that allows us to have private thoughts, plan
and guide behavior, and deceive others.
4. A language of self-reflection, which arises in states of openness, low defensiveness, and safety, likely
dependent upon the activation of the default mode network.

While the first two levels of language arise spontaneously, the self-reflective forms of inner language require
higher levels of neural network integration, affect regulation, and cognitive processing. Reflexive language
keeps us in the present moment, while reflective language demonstrates our ability to escape from the present
moment, gain perspective on our thoughts and feelings, and make decisions about what we would like to
change and how to change it. Most forms of psychotherapy rely on developing and utilizing both forms of
self-reflective language. Four forms of language sharing a common lexicon can result in a great deal of inner
confusion. Many people report feeling crazy because of the simultaneous and contradictory voices they hear
within them. Psychotherapy often involves sorting out these different soundtracks in order to provide a clearer
idea of just what is going on in there.

Diagnosis and Treatment


The principal activities of brains are making changes in themselves.
—Marvin L. Minsky

Functional brain imaging has opened a window to the living human brain in the acts of motor tasks,
imagining a feared situation, being rejected by playmates, or telling a lie. An examination of areas activated
during these and many other behaviors has enhanced our understanding of basic brain-behavior relationships.
Although the application of scanning technology to psychopathology is still in its infancy, there have already
been many important and provocative findings. As scanning techniques become more precise and the
hardware more affordable, they will no doubt become incorporated into the practice of psychotherapy and
other learning contexts.
Neuroimaging has the potential to aid in diagnosis, treatment selection, and the prediction of treatment
outcome (Etkin et al., 2005; Linden, 2006). As part of an initial assessment, it could help therapists pinpoint
areas of neural activation and inhibition. Treatment planning will eventually come to include specific
psychotherapeutic and pharmacological interventions to enhance the growth and integration of affected
networks. Baseline patterns of activation will give clinicians clues to who will benefit from medication,

215
therapy, or some combination of the two. Regular scans during the course of therapy may someday be a useful
adjunct to psychological tests, as ways of fine-tuning the therapeutic process and measuring treatment success.
Associations between psychiatric symptoms and changes in the relative metabolism of different areas of
the brain are in the process of being uncovered. We have already seen lower levels of metabolism in the left
prefrontal cortices of depressed patients (Baxter et al., 1985, 1989) and increased metabolism in the right
prefrontal and limbic region of patients with PTSD (Rauch et al., 1996). The importance of the right frontal
region in PTSD is supported by clinical evidence such as the onset of PTSD after injury to the right frontal
area (Berthier, Posada, & Puentes, 2001) and a cure of PTSD symptoms after a right frontal lobe stroke
(Freeman & Kimbrell, 2001). The inhibition of Broca’s area during intense fear states is already a focus of
cognitive-behavioral therapies, and a reactivation of the language centers may become a standard measure of
success in the treatment of PTSD and other anxiety-related disorders. All of these findings support the
existence of specific circuitry involved in the recognition, reaction, and regulation of anxiety and fear in the
aftermath of traumatic experiences.
The focus of this new work, however, is somewhat different than the old localization theories that
attributed disorders of behavior to specific areas of the brain. We now understand that each region of the
brain participates in multiple neural systems with highly complex interactions and homeostatic functions. It is
actually the relationship between clinical symptoms and relative activity levels of specific neural networks that
are salient. The neurobiology of obsessive-compulsive disorder (OCD) has been of particular interest in this
regard. A neural circuit thought to mediate OCD includes the ompfc and subcortical structures called the
caudate nucleus, globus pallidus, and thalamus. This cortical-subcortical circuit, involved with the primitive
recognition of and reaction to contamination and danger, becomes locked into an activation loop in patients
with OCD (Baxter et al., 1992). It is hypothesized that the ompfc, or some other component of the OCD
circuit, activates the circuit with a worry signal, decreasing inhibition of the thalamus, which in turn excites
the ompfc and caudate nucleus (Baxter et al., 1992). The result is a feedback loop that is highly resistant to
inhibition or shutting down.

Network Homeostasis and Treatment Outcome


The constant conditions which are maintained in the body might be termed equilibria.
—Walter Cannon

Research in the neurobiological correlates of symptom change and the effects of psychotherapy on the brain is
ongoing (Barsaglini et al., 2013; Fitzgerald, Laird, Maller, & Daskalakis, 2008; Messina, Sambin, Palmieri, &
Viviani, 2013). Clinical improvement logically correlates with normalization of metabolic and neural activity
in brain regions involved with anxiety, mood, cognition and other functions. The same changes are observed
when clients demonstrate symptom relief during pharmacotherapy and psychotherapy (Abbas et al., 2014;
Apostolova et al., 2010; Nakao et al., 2005). Whether this is a result of treatment or the passage of time in any
individual is always unknown. However, studies with contrast groups (those who receive no treatment)
suggest that on the whole, treatment is better than no treatment.
Considerable evidence supports the idea that the reregulation of neural networks parallels some of the
symptomatic changes we witness over the course of psychotherapy. In general, decreases in fear and anxiety
correlate with activation reductions in bottom and right-hemisphere regions. In OCD, there is a reduction in
activation in a region dedicated to the control and inhibition of impulses, specifically the ompfc. In the
successful treatment of social and spider phobias, there is a decrease in activation in limbic and primitive
cortical areas and in right-hemisphere processing. In situations where there is a deficit in cognitive processing
such as in schizophrenia and brain injury, we see symptom reduction correlated with increased frontal
activation. Keep in mind that the correlations do not prove causal relationships; changes in brain activation
patterns could also be secondary to symptomatic changes, or both could be caused by some other unknown
factor.
Patterns of brain activation in both panic disorder and PTSD are a bit more complex. In both of these
disorders, sensory and memory networks are hijacked by the amygdala and become internal sources of fear.
While studies have not yet focused on changes in brain activation related to positive treatment outcome, we
can speculate that positive treatment response would correlate with decreased activation in the amygdala,
sensory-motor areas, and the cerebellum (Bryant et al., 2008; Pissiota et al., 2002), along with increased
ompfc activation (Phan et al., 2006; Williams et al., 2006). There would also be a decreased activation in
regions dedicated to autobiographical memory, and an increase in the processing of information from the

216
current external environment (Sakamoto et al., 2005).
Psychotherapy outcome research with a number of disorders has found changes in brain activation
paralleling symptomatic improvement. In each case, treatment seems to have reestablished a homeostatic
balance between interactive neural networks that were previously out of balance. Table 21.1 outlines the
studies that have been performed to measure the neural correlates of successful therapy across a variety of
patient populations.

TABLE 21.1
Successful Psychotherapy and Changes in Neural Activation

Diagnosis and Results


Treatment

OCD
BT vs. fluoxetine Both: Decreased metabolism in right caudate1
BT vs. fluoxetine BT: Left ompfc activation correlated with positive response
Fluoxetine: Changes in the opposite direction2

BT vs. fluoxetine Decreased activation in dlpfc, ompfc, and ACC with symptom provocation3
BT vs. controls Decreased cerebral blood flow in right caudate4

BT Decreased activation in lateral prefrontal cortex5

CBT Decreased metabolism in right caudate


CBT Normalization of thalamic activity 6

CBT and fluoxetine Increased bilateral gray matter7


Increased right parietal white matter8

Social Phobia
CBT vs. citalopram Both: decreased amygdala, hippocampal, and adjacent cortex activation
CBT: decreased periaqueductal gray activation
Citalopram: decreased thalamic activation9
CBT vs. controls Increased right frontal and occipital activation 10

Spider Phobia
CBT Decreased activation in parahippocampal gyrus and dlpfc11, right prefrontal cortex12, and in the insula and ACC13

Posttraumatic Stress Disorder


EMDR (case study) Increased activation in anterior cingulate and left frontal lobe14

Panic Disorder
CBT vs. antidepressants CBT: RH decreases in inferior temporal and frontal regions
LH increases in inferior frontal, medial temporal, and insula
Antidepressants: RH decreases in frontal and temporal lobes, LH increases in frontal and temporal lobes15

CBT Decreased activation in right hippocampus, left ACC, left cerebellum, and pons
Increased activation in medial prefrontal cortex16
PT Normalization of fronto-limbic activation patterns17

Major Depressive
Disorder
CBT vs. paroxetine CBT: decreased frontal activation/increased limbic
Paroxetine: changes in the opposite direction18

CBT vs. venlafaxine Both: decreased bilateral ompfc and left ompfc activation and increased activation in right occipital-temporal
cortex19

IPT vs. venlafaxine IPT: increased activation in right posterior cingulate and right basal ganglia
Venlafaxine: increased activation in right posterior temporal and right basal ganglia20

IPT vs. paroxetine Both: decreased activation in prefrontal cortex


Both: increased activation in inferior temporal and insula21

IPT vs. paroxetine Both: symptom reduction with decreased frontal activation
Both: positive correlation with cognitive symptoms22

217
Schizophrenia
Cognitive rehab Increased frontal activation with improved performance23
Cognitive rehab Increased activation in right inferior frontal cortex and occipital lobe24

Traumatic Brain Injury


Cognitive rehab Global activation increase in 3 of 5 patients25

The results shown in this table should be considered preliminary because of small sample sizes and variations in methodology.

BT—behavior therapy; CBT—cognitive-behavioral therapy; IPT—interpersonal therapy; PT—psychodynamic therapy; EMDR—eye


movement desensitization and reprocessing; ACC—anterior cingulate cortex. Adapted and expanded from Roffman et al. (2005).

Functional scan studies have demonstrated that improvement of OCD symptoms is correlated with
decreased activation of the ompfc and caudate nucleus (Rauch et al., 1994). Especially interesting to
psychotherapists is the fact that these changes in brain metabolism are the same whether patients are
successfully treated with psychotherapy or medication (Baxter et al., 1992; Schwartz et al., 1996). Although
psychotherapy and medication are the first choices for treatment, they are not always successful. Scan-guided
psychosurgery for patients who do not respond to any other forms of treatment can disrupt runaway feedback
by severing neural links within the OCD circuit (Biver et al., 1995; Irle, Exner, Thielen, Weniger, & Ruther,
1998; Rubino et al., 2000).
Because symptoms can have multiple underlying causes, diagnoses aided by neural network activity could
improve diagnostic accuracy. Increased specificity will naturally lead to increasingly specific psychotherapeutic
and pharmacological interventions. Tourette’s syndrome—a disorder characterized by involuntary
vocalizations and motor tics—often occurs in individuals who also suffer with OCD, enuresis, or ADHD.
This is not a coincidence because these disorders share underlying neural circuitry and neurotransmitters
(Cummings & Frankel, 1985). They all stem from problems with the inhibition of subcortical impulses by the
frontal cortical areas. Thus, structural, biochemical, and regulatory abnormalities in these interrelated top-
down networks can result in all four conditions. When this circuit is more fully understood, symptoms of
OCD, ADHD, enuresis, and Tourette’s syndrome may all become subsets of some future diagnosis referred
to by the neural networks responsible for these functions.
In anxiety and depression, some studies show that therapy achieves results through increased cortical
versus subcortical activation, while others show changes in the activation patterns within the frontal lobes
(Porto et al., 2009). And while psychotherapy and medication can both lead to symptom reduction, there is
only partial neuroanatomical overlap in how they achieve their results (Roffman et al., 2005). In other words,
the same results can be achieved by different treatment strategies and through changes in the balance among
different neural networks. This is in no way bad news for psychotherapy. Cognitive therapy by experienced
therapists is as efficacious as medication in moderate to severe depression (DeRubeis et al., 2005). For
depressed patients with a history of child abuse, psychotherapy has been shown to be more effective, with the
addition of medication showing small benefits (Nemeroff et al., 2003).
Part of the answer about the benefits of psychotherapy may be reflected in the contrast with the effects of
medication in its impact on the brain. Successful psychotherapy and pharmacotherapy for anxiety and
depression have both been shown to result in functional changes in the brain’s cortical-limbic fear circuitry.
Pharmacotherapy calms amygdala activation, resulting in a bottom-up change mechanism, while
psychotherapy increases cortical activation, helping to inhibit the amygdala via top-down systems (Quidé et
al., 2012). This shows that the strategies of talk therapy leverage the mind (and the cortex) to use its natural
top-down modulatory functions to regulate symptoms of anxiety and depression.
In one of the largest studies conducted thus far, the frontal activation of 45 OCD sufferers who did not
respond to SSRIs were studied before and after behavioral treatment (Yamanishi et al., 2008). Those who
responded to behavioral therapy (BT) showed parallel decreases in activation in lateral prefrontal regions,
possibly reflecting a decreased effort in the cortical inhibition of impulses (Saxena et al., 1999). Perhaps a
more interesting finding is that response to BT was predicted by a higher baseline level of orbital frontal
activation. These results were supported in a more recent study with OCD sufferers, where treatment
response was predicted by higher baseline activity in the ACC, a basal forebrain structure closely interwoven
with the ompfc (O’Neill et al., 2013). A similar dynamic has also been seen in patients with panic disorder
and depression who are more likely to benefit from treatment with higher pretreatment levels of activation
within the ACC and stronger ACC-amygdala connectivity (Huang et al., 2014; Lueken et al., 2013; Straub et
al., 2015). Utilizing resting-state connectivity between relevant brain regions as a way to predict treatment
response is gaining increasing interest (Crowther et al., 2015; Hahn et al., 2015).

218
While this is generally interpreted as reflecting top-down inhibition (Milad & Quirk, 2002; Yang et al.,
2014), it may also suggest that successful BT and CBT treatment for depression, OCD, panic, and other
anxiety-based disorders may rely on the interpersonal and self-reflective capacity provided by the default mode
network, of which the ompfc and ACC regions are central components. These capacities would allow patients
to better connect with their therapists, benefit from the nonspecific social and emotional aspects of therapy,
and enhance their abilities to use self-awareness and personal will to stick with treatment. These studies may
suggest that the mind, and the relationships in which it is embedded, really matters (Beauregard, 2007;
Corrigan, 2004; Siegel, 2015).

The Centrality of Stress


It’s not stress that kills us, it is our reaction to it.
—Hans Selye

Although some stress is a normal part of life, early, prolonged, or severe stress can result in significant and
long-term impairments in learning, attachment, and physiological regulation (Glaser, 2000; O’Brien, 1997;
Sapolsky, 1996). Stress plays a role in the expression and severity of most, if not all, psychiatric and medical
disorders. Therefore, assessing and targeting stress as a focus of psychotherapeutic intervention should always
be an aspect of healing relationships. Since therapists are trained to think in terms of diagnostic categories and
treatment modalities, stress often flies under our diagnostic radar. Understanding and working to regulate our
clients’ stress is central to psychotherapeutic success because of its impact on neuroplastic processes.
An emerging concept in treatment involves buffering victims of stress from neural compromise by altering
their neurochemistry. One way of accomplishing this is to block the secretion or uptake of norepinephrine and
glucocorticoids soon after a traumatic experience (Brunet et al., 2008; Liu et al., 1997; Meaney et al., 1989;
Watanabe, Gould, Daniels, Cameron, & McEwen, 1992). It has also been found that the neurotransmitter
neuropeptide-Y is found in higher concentrations in the amygdalas of individuals who respond more favorably
to high levels of stress (Morgan et al., 2000). Artificially increasing levels of neuropeptide-Y may buffer the
nervous system from some of the damaging effects of stress.
Chemical blockade or disruption of particular amygdala circuits may decrease some of the symptoms of
PTSD such as the startle and freeze responses (Goldstein, Rasmusson, Bunney, & Roth, 1996; Lee & Davis,
1997). It has even been suggested that stimulation of the amygdala could lead to the extinction of conditioned
fear (Li, Weiss, Chaung, Post, & Rogawski, 1998). Understanding the role of long-term potentiation and
other forms of plasticity in the amygdala, as well as its role in fear conditioning, may provide another avenue
for future interventions in psychosis and PTSD (LaBar, Gatenby, Gore, LeDoux, & Phelps, 1998; Rogan &
LeDoux, 1996; Rogan, Staubli, & LeDoux, 1997).
We have seen that higher levels of maternal attention in rats decrease the pups’ subsequent HPA
activation in response to stress (Liu et al., 1997). Although I doubt that encouraging human mothers to lick
their children will be of much help, human infants demonstrate the same pattern in response to maternal
massage (Field et al., 1996) and within securely attached relationships (Spangler & Grossman, 1993).
Maternal depression, separation, and deprivation are severe stressors for infants, resulting in a variety of
negative biological, emotional, and social consequences (Gunnar, 1992). Aggressive treatment of depression in
new mothers, along with teaching them how to massage and better interact with their infants, may counteract
some of the negative impact of maternal depression. Therapy focused on resolving a mother’s attachment
difficulties or past trauma prior to giving birth may also be helpful in reducing stress in their infants, children,
and adolescents (Trapolini, Ungerer, & McMahon, 2008).
An increased appreciation of the effects of maternal separation may guide us on the advisability of optional
infant-mother separation. Where separation is unavoidable in cases of illness and death, the ability to lessen
the impact of stress hormones via interpersonal and chemical interventions may prevent problems later in life.
Given the amount of exposure in our society to stressful events such as abuse, neglect, abandonment, and
community violence, the impact of severe stress on mothers and the developing brains of their children should
be a serious public health concern (Bremner & Narayan, 1998).
Research findings suggest that early stress leads to a vulnerability to depression later in life (Widom,
DuMont, & Czaja, 2007). This, in part, is mediated by deficient organization of frontal circuitry, and the
establishment of lower levels of excitatory neurotransmitters and growth hormones during sensitive
developmental periods. Early childhood experiences leading to a bias toward right-hemisphere activation may
also play a role in the long-term development of depression. As we discussed in Chapter 6, magnetic

219
stimulation of the left hemisphere of depressed patients and the right hemisphere of patients with mania has
shown promising results and may serve as a future alternative to electroconvulsive therapy (Grisaru et al.,
1998; Klein et al., 1999; Teneback et al., 1999; Pascual-Leone et al., 1996).
In line with these findings, activation of the left hemisphere through sensory stimulation results in a
higher degree of self-serving attributions and positive affect (Drake & Seligman, 1989). Relative left frontal
activation appears linked to a state of mind of “self-enhancement,” which may decrease the risk for
psychopathology and be manipulated by changes in attitude or practices such as mindfulness meditation
(Tomarken & Davidson, 1994). The more we understand the relationship between laterality and affect, the
more we may be able to incorporate techniques of selective activation of right and left hemispheres into
multimodal treatments for mood disorders and other psychiatric difficulties.
PTSD is primarily mediated and maintained by neurobiological processes outside conscious control. The
activation of Broca’s area in the face of high levels of affect appears to be an important mechanism of action in
most interventions with patients suffering from PTSD and other anxiety disorders. We know that the ompfc
modulates and inhibits amygdala activity, the very circuit we are activating when we help clients to employ
cognition to inhibit their fears.
Despite new theories connecting neural communication and psychopathology, no major form of
psychotherapy has emerged with the stated goal of neural network integration. This being said, techniques
such as the caloric test and the eye movements used in EMDR seem to involve some sort of neural
rebalancing as an active element. Previously, we discussed the phenomenon of sensory neglect, which occurs
when there is damage to the right parietal lobe (assumed to be responsible for the integration of sensory and
motor information from both sides of the brain). In the caloric test, stimulation with cold water to the left ear
results in rapid side-to-side eye movements while activating regions of the right temporal lobe (Friberg et al.,
1985). Although there has been one report of permanent remission of sensory neglect with this treatment, for
most the cure is only temporary (Rubens, 1985). The bilateral activation of attentional centers in reaction to
the caloric test results in increased integration of previously dissociated attention and information-processing
systems (Bisiach et al., 1991).
Plenty of research evidence points to the fact that the way information and memory are processed in the
brains of those suffering with PTSD is disrupted. We know that our clients tend to narrow their range of
behaviors, activities, and emotions out of fear of being unexpectedly triggered by running into something
novel, a symptom called neophobia. When nontraumatized people experience something new, the ACC puts
us on alert to pay attention and be prepared to react to and learn from something unexpected. It could be
good or bad—we don’t know—but if we are not afraid, the curiosity and exploratory urge counterbalance the
fear of the unknown. For those with PTSD, the circuits that prepare us for novelty and new learning are not
activated. Rather, regions of the brain that organize autobiographical and somatic memory fire instead. Thus,
neophobia is actually a fear of being reminded of the pain and terror from the past. In other words, the brains
of PTSD victims become shut off from both new learning and the ability to update past experiences.
Our memory and motor systems have been engaged in collaborative evolutionary processes since we were
fish. One piece of evidence for this is that when we use the large muscles of our legs to walk or run, they
secrete neural growth factors that cross the blood-brain barrier and stimulate neuroplasticity and learning.
This connection is most likely due to the fact that our muscles evolved to tell our brains to pay attention when
we are moving because something important is happening—otherwise, why would we be moving? In a similar
fashion, side-to-side eye movements likely trigger systems for memory updating because they are historically
related to foraging and looking out for predators and prey. This is probably why rapid eye movement (REM)
sleep occurs while our brains are engaging in the consolidation of new memory. We don’t need to move our
eyes to consolidate memory, it is just an artifact of the coevolution of foraging, the orienting response, eye
movements, and needing to update our map of our territory.
In the treatment of PTSD with EMDR, past traumatic events are recalled and subjected to a protocol that
involves focusing attention on ideas, self-beliefs, emotions, and bodily sensations. In addition, EMDR uses
periodic stimulation through watching the therapist’s hand going back and forth or having the legs touched
alternately (Shapiro, 1995). This bilateral and alternating (side-to-side) stimulation may serve to activate
attention centers in both temporal lobes in a manner similar to the caloric test. Alternating activation may, in
fact, enhance neural network connectivity and the integration of traumatic memories into normal information
processing.
While side-to-side eye movements are evolutionarily significant in their own right, it is likely then that
EMDR can activate the orienting reflex through multiple sensory modalities. The orienting reflex is our
automatic response to shift our focus to something that catches our attention. This primitive reflex is seen in
all animals as an adaptive response to the environment. As memory systems evolved, it is likely that they were

220
networked with the orienting reflex to signal the brain to get ready to learn new information. It is likely that
the orienting reflex counteracts the tendency in the brains of people with PTSD to go back to
autobiographical memory in the face of new information.
Stimulating the orienting reflex through eye movements or tapping the legs activates the brain’s novelty
detection centers and allows for memory reconsolidation. This allows new information to enter and old
information to be updated and modified. When EMDR works, I suspect this may be the mechanism of
action. You might say that EMDR is a way to trick the brain of individuals with PTSD to process new
experiences in a manner that allows them to live in the present.
Techniques such as EMDR may thwart or reverse the brain’s tendency toward neural network dissociation
secondary to trauma. Bilateral stimulation may enhance the reconsolidation of traumatic memories with
cortical-hippocampal circuits providing contextualization in time and place. Activation of these circuits creates
the possibility of building descending inhibitory links to subcortical sensory-affective memory circuits (Siegel,
1995). Thus, the right-left stimulation of attention may simultaneously trigger integration of affect with
cognition, sensation, and behavior throughout the brain.
Once the relationships among neural circuits are more fully understood, psychotherapists may employ
these and other noninvasive techniques to stimulate the brain in ways that enhance neural network
integration. Could activation of right-hemisphere emotional regions during therapy with alexithymic patients
aid in the integration of emotional processes with left-hemisphere linguistic circuitry? Could activation of the
left hemisphere during emotional dyscontrol in borderline patients enhance their ability to gain cognitive
perspective and emotional regulation?
For conditions involving too much emotional inhibition, new learning may be stimulated by creating
moderate levels of affect in therapy; this learning, in turn, may create a biochemical environment more
conducive to the integration of emotional circuitry into consciousness (Bishof, 1983; Chambers et al., 1999).
This may be the underlying neurobiology of Freud’s belief that the presence of affect is necessary for change.
Simultaneous activation of neural networks of emotion and cognition may result in a binding of the two in a
way that allows for the conscious awareness and integration of emotion.

Treatment Rationales and Combinations


Sometimes I lie awake at night, and ask, “Where have I gone wrong?” Then a voice says to me, “This is going to take more than one night.”
—Charles Schulz

The fundamental premise of this book is that any form of psychotherapy is successful to the degree to which it
positively impacts underlying neural network growth and integration. I expect future research to continue to
support this basic hypothesis. Furthermore, evolving technologies will provide us with increasingly accurate
ways of measuring activity within the brain and a greater understanding of exactly what it is we are measuring.
My hope is that including neural network activity in our case conceptualization may help to establish a
common language for us to select, combine, and evaluate the treatments we provide. It will, one hopes, help
us to move past debates between competing schools of thought to a more inclusive approach to psychotherapy.
A long and hard-fought debate about treatment continues between supporters of psychopharmacology and
psychotherapy despite empirical support for both approaches, individually and in combination. Brain
functioning offers us a way to look more deeply into the effects of both talk and medication in regulating the
brain and stimulating neuroplasticity. Patients who come to see me for psychotherapy are often adamant in
their refusal to consider medication. Some feel frightened or shamed if I suggest the use of drugs as an adjunct
to psychotherapy. At the same time, I know that many discount talk therapy and will only seek help from
therapists who will prescribe medication. All clients could benefit from education about brain functioning and
the potential (even synergistic) power of both interventions. On the one hand, the therapeutic alliance
supports positive expectancy, medication compliance, and psychological well-being. On the other, medication
can help to achieve a state of body and mind that allows clients to benefit from psychotherapy.
Many patients who suffer brain damage resulting from accidents participate in multimodal rehabilitation
programs that include physical, cognitive, and psychosocial interventions. The general approach to
rehabilitation after brain injury is to first assess which systems have been damaged and which have been
spared. The next step is to develop a program that plays to the patients’ strengths and attempts to compensate
for their weaknesses. Traffic and industrial accidents often result in damage to the frontal cortex, making
disorders of attention, concentration, memory, executive functioning, and emotional regulation common
among those who need neurological rehabilitation. The same difficulties are common in many forms of

221
psychological distress and psychiatric illness.
The traditional split between mind and brain has resulted in the separate development of the fields of
psychotherapy, neuropsychology, and rehabilitation. When psychological difficulties are conceptualized in the
context of a brain-behavior relationship, applying techniques from cognitive rehabilitation in psychotherapy
becomes an interesting possibility. For example, abnormalities of frontal lobe functioning have been found in
OCD, depression, and ADHD. Because these disorders share many symptoms afflicting patients with brain
injury, psychotherapy patients with these and other psychiatric diagnoses may benefit from the strategies of
cognitive rehabilitation (Parente & Herrmann, 1996).
An example of this was given in Chapter 5, when I discussed how the simple memory strategies I used to
assist my patient Sophia in remembering her appointments helped us to establish a solid alliance. My working
assumption was that a combination of decreased hippocampal volume due to chronic stress and
hypometabolism in the temporal lobes related to depression created real, brain-related memory dysfunctions
(Bremner, Scott, et al., 1993; Brody et al., 2001). The success of cognitive-behavioral treatments with
depressed and anxious patients underlines the importance of focusing on basic issues of reality testing, focused
attention, and emotional regulation in order to support prefrontal functioning (Schwartz, 1996).
Findings with borderline clients show damage or dysfunction of the frontal and temporal lobes supports
the use of cognitive rehabilitation techniques with this population (Paris et al., 1999; Swirsky-Sacchetti et al.,
1993). This may help explain why borderline patients require increased levels of structure to scaffold their
erratic executive control and emotional instability. Manipulation and organization of the physical
environment, sensory stimulation, and the type and amount of activity all impact brain functioning.
Psychoeducation and enlisting family and friends in the therapeutic process (as utilized extensively in
rehabilitation after brain damage) are also potential mechanisms of change. A good example of this is
dialectical behavioral therapy (Linehan, 1993), which combines exposure, cognitive modification, skills
development, and problem-solving skills to support prefrontal functioning.
Diagnostic and treatment approaches focused on cognitive deficits serve to decrease shame and help to
create a stronger treatment alliance. Highly structured skill-building techniques, in the context of support and
understanding, may provide disorganized patients the opportunity for early and clearly measurable success
experiences. As our understanding of neural networks related to memory, affect, and behavior expands,
prosthetic aids to these systems will be created and applied in the psychotherapy context. Increasing
interdisciplinary coordination of this kind will require more comprehensive training for psychotherapists, not
only in neuroscience but also in cognition, memory, and rehabilitation science. Removing the traditional
barriers between psychotherapy and rehabilitation may lead to a higher quality of care and greater treatment
success.

Why Neuroscience Matters to Psychotherapists


In science the important thing is to modify and change one’s ideas as science advances.
—Herbert Spencer

The psychotherapist as healer joins a long tradition of rabbis, priests, medicine women, and shamans back
into prehistory. At the same time, research in neuroscience makes it abundantly clear that we are also on the
cutting edge of the scientific mainstream. In contrast to technological medicine, our techniques use both our
personal role in the healing relationship and the importance of the subjective experience of our clients. In the
absence of a brain-based model of change, the pioneers of our field have learned to stimulate and guide
neuroplastic processes in order to help build, integrate, and regulate our brains. But why does an academic
understanding of neuroscience matter in our work? Here are a few thoughts.
On a practical level, adding a neuroscientific perspective to our clinical thinking allows us to talk with
clients about the shortcomings of our human brains instead of the problems with theirs. The truth appears to
be that many human struggles, from phobias to obesity, are consequences of brain evolution and not character
deficits. Identifying problems that therapists and clients hold in common and developing methods to
circumvent or correct them is a solid foundation upon which to base a collaborative therapeutic alliance.
As we come to better understand the neural correlates of mental health and emotional well-being, we may
be able to use this knowledge to aid us in diagnosis and treatment. Neuroscience may also someday provide us
with a rationale for an informed eclecticism as well as additional means of evaluating outcomes. We will be
able to see which combination of treatments impacts targeted neural networks and how changes in the
activation of these circuits correspond with symptom expression. Neuroscience can also provide a common

222
language to communicate with physicians, pharmacologists, and neurologists who may also be treating our
clients. Finally, if you are anything like me, you might find a neuroscientific perspective to be an exciting
addition to many case conceptualizations.
Some therapists bristle at the integration of neuroscience and psychotherapy, calling it irrelevant or
reductionistic. I think I understand their perspective and concerns—if you have a model of therapy that works,
why bother with the brain? Would Rogers, Kohut, or Beck have been better therapists if they had been
trained as neuroscientists? Maybe not. On the other hand, it is hard for me to grasp how the brain could be
irrelevant to changing the mind. And while I dislike reductionism as much as the next person, doesn’t a
tendency toward reductionism say more about the thinker than the nature of natural phenomena? Our
knowledge of neuroscience highlights the fact that we primates have complex and imperfect brains and should
remain skeptical about what we think we know. In other words, primates would be wise to doubt their beliefs
and remain open to new ideas.
It is humbling and more than a little frightening to realize that we rely on what may be the most complex
structure in the universe for our survival with little knowledge of how it works. But even though we are only at
the dawn of understanding the brain, an appreciation of its evolutionary history, developmental sculpting, and
peculiarities of design can surely encourage us to begin to use it more wisely. Practical things—like
understanding the neural damage resulting from drugs, stress, and early deprivation—should influence
everything from personal decision making to public policy. The neural network dissociation that often results
from exposure to combat should make us pay closer attention to those whom we put in harm’s way. Even our
tendencies to distort reality in the direction of personal experience and egocentric needs should lead us to
examine our beliefs and opinions more carefully.
We now know that mind, brain, and body are indivisible and that disorders traditionally thought of as
psychological need to be reconceptualized to include their neurobiological and somatic components. And if
brain dysfunction is central to a client’s difficulties, the “most illuminating interpretation” may not be as
valuable as a little accurate neurobiological knowledge (Yovell, 2000). Self-awareness is a relatively new
phenomenon in evolutionary history. Psychotherapy increases neural integration through challenges that
expand our experience of and perspective on ourselves and the world. The challenge of expanding
consciousness is to move beyond reflex, fear, and prejudice to a mindfulness and compassion for ourselves and
others. Understanding the promise and limitations of our brains is but one essential step in the evolution of
human consciousness.

Summary

Our brains are inescapably social, their structures and functioning deeply embedded in the family, tribe, and
society. And while the brain has many shortcomings and vulnerabilities, our ability to link with, attune to, and
regulate each other’s brains provides us with a way of healing. This is why the power of human relationships is
at the heart of psychotherapy. From my perspective, the value of neuroscience for psychotherapists is not to
explain away the mind or generate new forms of therapy, but to help us grasp the neurobiological substrates of
the talking cure in an optimistic and enthusiastic continuation of Freud’s “Project for a Scientific Psychology.”

223
Chapter 22
How People Change
The curious paradox is that when I accept myself just as I am, then I can change.
—Carl Rogers

Someone once asked me if I knew the difference between a rat and a human being. I was curious, so I played
along; here is what he told me. If you take a hungry rat and put it on a platform surrounded by five tunnels
with cheese hidden down the third tunnel, the rat, smelling the cheese, will explore the tunnels until it finds
the cheese. Rats have excellent spatial memory, so if you put the same rat in the same place the next day, it
will immediately go down the third tunnel. If, in the meantime, you’ve moved the cheese to the fifth tunnel,
the rat will still go down the third tunnel expecting to find the cheese where it was before. So what’s the
difference between a rat and a human being?
Rats are realists and soon come to accept that the cheese is gone and move on to explore the other tunnels.
Humans, on the other hand, will go down the third tunnel forever because they come to believe that’s where
the cheese should be. Within a few generations, humans will develop rituals, philosophies, and religions
focused on the third tunnel, invent gods to rule over it, and create demons to inhabit the other four. The rat’s
simpler brain provides it with no reason to persevere in the face of failure, while humans are experts at
developing and adhering to beliefs. When our brains cause unnecessary suffering because of these beliefs, we
need our minds to come to our rescue.
Our brains are organs of adaptation and survival, designed to do things as fast as possible with the smallest
amount of information. So once they come upon a solution to a problem, like never expressing negative
emotions during childhood, they become shaped to never express negative emotions again. You might also
adapt pacifist philosophies, get into relationships with violent people, and feel like a success because you don’t
express negative emotions. Brains excel in coming to unconscious conclusions and shaping our conscious
experience to reinforce the beliefs we already hold.
Brains are inherently conservative and want to keep doing what’s worked in the past: don’t take risks, fit
in, and do what your parents want you to do. For the lucky souls whose brains are well matched to their
circumstances, life works pretty well. For the rest of us, being stuck between our programming, the
expectations of our tribe, and our own needs for emotional health, can make us physically and mentally ill.
When our life isn’t working and we are anxious, depressed, or engaging in self-destructive behaviors, we go
to therapy seeking change. Often, we are unaware of what is causing our suffering and continue to employ the
same unsuccessful strategies in our lives with continued negative results. Our smartest clients can be the most
difficult to help because they are often well-rewarded for doing things their way. Successful accountants with
OCD and delusional fortunetellers will shower you with examples of how what you are calling a problem has
made them successful.
In contrast to our brains, our minds emerged much farther down the evolutionary path. We still don’t
understand the origins of the mind, but it probably had something to do with groups of brains coming
together to form the superorganisms we call tribes. As attention and memory became stabilized by group
process, our interactions grew into culture, which became the template for mind and the eventual formation of
individual identity. At this point in evolution our best guess is that the human brain is a social organ and the
mind is a product of many interacting brains.

A Social Organ of Adaptation


We are afraid to care too much, for fear that the other person doesn’t care at all.
—Eleanor Roosevelt

224
Human brains are social organs of adaptation—meaning that their growth and organization is shaped and
reshaped in the process of ongoing experience. At the same time, our ability to adapt to new situations is
constrained by habit and prior learning. The dynamic tension between habit and adaptation lies at the heart of
psychotherapy. This is why our clients consciously enlist our help in changing suboptimal functioning,
although their brains automatically work against our efforts. While a client’s resistance may be seen as an
impediment to change, it is the central focus of change in psychotherapy. Being a therapist means always
skating the delicate edge between stability, flexibility, rigidity, and change.
Despite our natural resistance to it, change is a normal part of life. Sometimes we are forced to change;
sometimes we want to change because old patterns of behavior have become too painful or no longer fit who
we’ve become. That’s when we go on a quest to discover something new. Some of us go to the desert; others
find a guru; and still others go to therapy. A brief glance at history reveals that people were changing long
before therapy arrived on the scene. The classic example is the heroic journey from adolescence to adulthood,
in which the hero is able to break away from the constraints of his or her childhood in order to discover new
ways of being. These kinds of life changes have historically been called maturation, redemption, or
transformation.
People change when new experiences disrupt old stimulus-response patterns (habits). Because we depend
on the repetitive execution of habits to feel safe and in control, disrupting these patterns makes us anxious.
This is where the emotional support of a therapist is most important. The reflexive response to change is to
either return to old patterns or grasp some new habit or belief system to escape the anxiety of uncertainty. If
we can tolerate this anxiety and stay on our journey, change is inevitable. We can leverage many aspects of
experience in our quest—the more neural networks we are able to tap into, the more leverage we will have in
service of changing our brains. Because our brains consist of complex and interwoven neural networks, there
are many possible avenues of change.

Habits and Flexibility


When I let go of who I am, I become what I might be.
—Lao-tzu

Robots and humans share the fact that their actions are based on past programming. The actions of robots are
organized in algorithms embedded in lines of code; a robot’s coding is a long list of “if this, then that”
statements that allow it to react to all of the contingencies that the programmer had the foresight to
anticipate. This is how Siri is able to know if we should wear a sweater or if the Cubs are playing at home this
weekend.
In contrast to a robot’s computer code, human habits are maintained by memories stored in ensembles of
neural networks. When triggered by internal or external cues, the associated memories activate our behaviors,
thoughts, and emotions. The unconscious activation of old programming keeps us doing the same old same
old, which is why every psychology student is told that the best predictor of future behavior is past behavior.
Psychotherapists generally believe in three levers of change: feelings, behaviors, and thoughts. Most
traditional forms of psychotherapy use either one or some combination of the three to promote change. For
behaviorists, changes in behavior are believed to lead to changes in emotions and thoughts. This is why B. F.
Skinner believed that modifying reinforcers within the environment would lead to changes in how we think
and feel.
For psychodynamic therapists, the primary lever of change is emotions that will lead to changes in
thoughts and behaviors. In contrast, cognitive therapists believe that thoughts drive feelings and behaviors;
change someone’s thoughts and changes in feelings and behaviors will follow. More innovative forms of
therapy have also discovered the power of movement, body work, art, and meditation to promote change. The
underlying theory that guides each therapeutic school assumes that their particular target of intervention is the
primary driver of change. Over the years, I’ve seen people have great success in therapies that would fall into
all of these categories.
Presented as dogma, each perspective of psychotherapy is simultaneously right and wrong. Each works or
doesn’t depending on the client and the quality of the therapeutic relationship. For many therapists, seeing
this reality is a challenge because most of us choose an orientation to therapy based on our own experiences,
needs, and defenses. This unconscious egocentric bias leads most of us to believe that our view of therapeutic
change is correct to the exclusion of others. This assumption makes us vulnerable to all kinds of biases in
judgment. This tendency toward dogma also places us at risk of interpreting treatment failures as problems in

225
our clients instead of in our techniques or ourselves.
Although my personal bias is in the direction of psychodynamic forms of therapy, if I look closely at what
I actually do, I’m also using behavioral and cognitive therapy. As I create experiments in living with my
clients, I challenge them to engage in new behaviors and ways of thinking. I work with them to alter the
reward contingencies in their lives and challenge dysfunctional patterns of thinking. I encourage clients to
consider meditation, yoga, dance, or any other way they can simultaneously soothe their anxiety and explore
all corners of their minds and bodies. The best cognitive therapists I know invest time in forming solid
relationships with their clients and include discussions of emotions and behaviors as part of treatment.
Behavioral therapists usually connect with and educate their clients as they establish the intellectual and
interpersonal contexts for their work.
What we have learned from neuroscience hasn’t supported any one of these perspectives over the others. In
fact, understanding the interwoven nature of neural networks has challenged us to engage in a higher level of
integrative thinking. This is probably why most well-trained therapists leverage all three neural avenues to
some degree regardless of their orientation. All three schools, with the addition of systems therapy and
adjunctive techniques such as EMDR, can be synthesized into more efficacious interventions.

The Power of Connection


Experience is a biochemical intervention.
—Jason Seidel

The implications of having a social brain are widespread, including the fact that many things we think of as
being objectively true—knowledge, memory, identity, and reality—are largely social constructions. One of the
reasons why things seem more real when we experience them with others is because our experience of reality is
primarily social. Children show us this in a very straightforward way as they implore us to “watch this, watch
this!” as they do cartwheels or perform a magic trick.
Our social brains allow us to link to those around us, connect with the group mind, and regulate one
another’s states of mind. In order to tolerate the anxiety of change, we need to feel safe—this is why the
quality of the therapeutic relationship is so vital to the success of any form of therapy. Secure attachment to
the therapist also activates key drivers of neuroplasticity such as decreases in cortisol, which inhibits
hippocampal functioning, protein synthesis, and new learning.
In addition to the biological consequences of positive relationships, our minds are also more apt to change
when linked to other minds. Having a witness activates mirror neurons and theory-of-mind circuitry, making
us more aware of others and ourselves while reinforcing our identity. The importance of our brains linking
across the social synapse for therapeutic success is probably why the quality of the client-therapist relationship
(as perceived by the client) has the strongest positive correlation with treatment success of any variable to have
been studied.
The therapeutic stance suggested by Carl Rogers over half a century ago is likely the best interpersonal
environment for neural plasticity and social-emotional learning. His focus on warmth, acceptance, and
unconditional positive regard minimizes the need for defensiveness while maximizing expressiveness,
exploration, and risk taking. His orientation to therapy was not to try and solve specific problems, but to help
clients gain the broadest possible awareness of their thoughts and emotions. It is not surprising that these
characteristics have emerged as beneficial for positive child development.
What might be going on in the brain and body of a client who is receiving warmth, acceptance, and
positive regard? Social interactions early in life result in the stimulation of both neurotransmitters and neural
growth hormones that participate in the active building of the brain. It is likely that oxytocin and dopamine
become activated in states of attunement, enhancing a client’s ability to benefit from treatment by stimulating
neuroplasticity. A strong therapeutic bond also increases metabolic functioning that drives blood flow, oxygen
availability, and glucose consumption supportive of new learning.
A Rogerian interpersonal context would allow a client to experience the widest range of emotions within
the dyadic scaffolding of an empathic other. The trust generated in the context of attunement appears to allow
our minds to be open to what we might otherwise reflexively reject. This openness would increase
receptiveness for interventions like supportive rephrasing, clarifications, and interpretations to gain access to
conscious consideration. Clients can then link minds with their therapists to cocreate new narratives
containing the blueprint for more adaptive thoughts, feelings, and behaviors.
It is obvious that we can model the outward behaviors of others and imitate their physical actions. What is

226
less obvious is that the mechanisms of our social brains allow us to attune to the mental activities of those
around us. While this attunement is inhibited with those we fear, trust makes it more likely that we will
spontaneously imitate the actions, thoughts, and feelings of those we like. A therapist who isn’t liked will be
unable to leverage the powerful forces of imitation and emotional resonance to drive positive change.
Those who are nurtured best survive best during childhood and more easily benefit from therapy later in
life. Unfortunately, the social isolation created by many psychological defenses can separate us from the
positive emotional connectedness that drives healing. One of the goals of therapy is to generate trust and
connection so that our clients can rejoin the group mind and benefit from its natural healing properties.

Healing Trauma
Never be afraid to sit a while and think.
—Lorraine Hansberry

Fear and terror change our brains in ways that disrupt the continuity of experience and can lead us to
disconnect from the group mind. I have had clients from Baghdad, Beirut, and London, separated by culture,
language, and generations, who shared the experience of being in buildings as they were destroyed by bombs.
These victims of war describe similar experiences—hearing the whistling growing louder, the explosion, the
violent movements of the floor and walls, followed by long periods of silence, and struggling to breathe
through the clouds of dust. Then the aftermath, digging their way out, climbing over the bodies of relatives
and neighbors, and then prolonged shock that can last for decades.
Another client, fleeing on foot from Eastern Europe to escape the Holocaust, was crossing a field with his
older brother. A Nazi plane spotted them and dropped a bomb that landed within feet of where they had
taken cover. They stared at the unexploded bomb for what seemed like an eternity before they continued to
run. A young woman I worked with was driven to the desert by her sadistic husband and forced to dig her
own grave as he sat and sharpened a butcher knife. These are all experiences that terrify us to the point where
we can lose our words and disconnect from reality. Meanwhile, the trauma can get locked within us and
become the emotional soundtrack of our lives.
The value of someone who is willing to go with us to the ground zero of our pain, a witness to our horror,
should never be underestimated. Communicating our story to another encourages us to articulate a traumatic
experience that may be represented in our brains only as a fragmented collection of images, bodily sensations,
and emotions. Once we have a conscious and articulated story, we gain the possibility of integrating the many
aspects of what has happened to us in order to find a way to heal. Seeing the reactions of the other to our
experiences helps us to grasp their meaning, and having to make them comprehensible to another helps make
them comprehensible to us. In addition, telling the story to others provides us with a new memory of the story
that now includes a witness, making it a public experience, and making it available to editorial changes. All
aspects of this cocreation of our experience support the idea that both reality and memory are social
constructs.
Keep in mind that the ability to heal psychic pain through storytelling has been woven into our brains over
eons of cultural evolution. When young therapists begin to hear their clients’ stories, they usually feel that they
have to do something with the information in order to earn their keep. Over time, we come to appreciate the
fact that simply bearing witness is an important part of our job. Sometimes the best thing to do, especially at
first, is to do nothing. As we weigh in with some of our thoughts and feelings, our clients gradually weave
them into stories.

Turning Your Mind Into an Ally


The body benefits from movement, and the mind benefits from stillness.
—Sakyong Mipham

Being in control of your mind first requires remembering that you have a mind, which is not as easy as it
sounds. While our bowels and bladders remind us of their existence, our minds prefer to go unnoticed. There
is no reflex to orient us to the mind’s existence, no pressure exerted or guilt employed if we ignore it.
Remembering we have a mind involves effort and discipline. This is why many people go through life never
having noticed they have one, let alone knowing they can change it or use it to their advantage.

227
If and when we do remember that we have a mind, what do we do with it? One of the first things we may
notice about our mind is that it keeps generating thoughts all on its own. There may be a momentary
interruption in the flow at first glance, but soon enough, your mind will go back to creating an incessant
stream of words, thoughts, and images. You want to cultivate the ability to observe these thoughts as they
flow by without identifying with them or reacting to them. The goal is to learn to let them go by, to be
replaced by the next, and the next, and the next. This river of thoughts will go on without effort or intention
until you come to realize that it is not you; you’re the one watching.
By gaining distance from this stream of thoughts and feelings, you are in a position to make some choices
that you couldn’t make before. You can evaluate how accurate and useful they are, and whether or not you
choose to believe them. You will soon come to realize that while the brain evolved to help us deal with
potential danger, the speed, intensity, and negative bias of all the thoughts it generates has gone too far. In
many ways, the cortex has grown too smart for our own well-being. The good news is that while brains evolve
over eons, we can change our mind in an instant. It may take decades for that instant to happen, but when it
happens, our mind is capable of discovering new ways of being.

Junk Food Junkie


A change in bad habits leads to a change in life.
—Jenny Craig

Years ago, I was telling my therapist that I wanted to get into better shape, but despite all my efforts, I always
seemed to fall short. “If only I could eat better,” I told her. In good therapeutic form, she replied, “Tell me
more.” “I exercise every day and eat lots of healthy food. On most days, I eat well all day until the evening,
when I become a junk food junkie.” I suppose you could call it binge eating, but because I was in such good
shape and not overweight, I never labeled it that way. But it was clear that the number and especially kind of
calories I was eating most evenings was not good.
The more I spoke of my evening binges, the more I became aware of how automatic and unconscious the
behavior was and how out of control I felt. As the session ended, she suggested that I think about it more,
which, of course, I promptly forgot until a few days later. One morning, as I lay in bed in that middle state
between sleeping and waking, a memory that felt like a daydream began to play in my mind.
I was a young boy, perhaps seven or eight, walking into the kitchen where my grandmother was cleaning
up after dinner. I was telling her that I was sad about something or other. Instead of saying anything, she
turned to the right, opened the refrigerator, reached into the freezer and pulled out a half-gallon box of
spumoni ice cream. As she turned back in my direction, she pulled off the paper zippers, grabbed a spoon, and
pushed it into the ice cream. I reached out, and she placed the box of ice cream in my hands. I turned around
in silence and walked into the living room, where I lay down on the sofa, placed the box of ice cream on my
chest, and ate until I could eat no more.
An important thing to know about my family is that the direct expression of negative feelings was rare,
and discussions about emotions were nonexistent. There was a special injunction against sadness, something I
only realized later in life. There had been enough tragedy, loss, and heartbreak already, and I, as the first child
of a new generation, was to be saved from sadness. This meant that expressing sadness needed to be staved off
at all costs. Everyone used food to distance themselves from emotions, and I learned this lesson well. The
downside was that it left all of us without a way to understand or language to communicate the painful aspects
of our inner worlds.
So the next question, is how do we use our mind to disrupt the negative patterns stored in our brain? We
need to find a tool the mind can use to interrupt the automatic brain’s stimulus-response chains. One that I’ve
found valuable, called HALT, comes from Alcoholics Anonymous. When you feel like having a drink or
engaging in any compulsive behavior, say the word “halt” and ask yourself whether you are hungry, angry,
lonely, or tired. The idea is that if you are going for a drink, there is probably some emotional trigger. Of
course, there are other emotions you should consider, but they wouldn’t spell HALT, which is quite useful.
HALT reminds you not only to interrupt the impulse to drink (or in my case eat), but also to remember
you have a mind, to be self-reflective, and to engage in a caring relationship with yourself. You are doing what
my grandmother was unable to do by asking, “I can see something is wrong. Let’s talk. Tell me what you are
experiencing.” The added awareness momentarily interrupts the stimulus-response chain and provides the
opportunity to reflect, reconsider, and take a healthy detour from our usual route. This is where a well-
established and mature relationship with your inner world really comes in handy.

228
Using the HALT technique is one way to use your mind to change your brain. Asking yourself what you
are feeling instead of engaging in reflexive behavior allows your mind to reconnect with and learn to retrain
the habits of your primitive brain. In a sense, you are giving yourself what you needed as a child—to be seen,
to feel felt, and to be helped to grasp and articulate your experience. This corrective internal reparenting, what
we strive for in therapy, will eventually create new neural circuitry that allows us to replace symptoms with
functional adaptations. In my case, it meant taking better care of myself, investing more in relationships, and
being more willing to directly confront emotional pain.

Constructive Introspection
Loneliness is the poverty of the self; solitude is the richness of self.
—May Sarton

During the early years of life, we experience ourselves as embedded within the group mind of our families,
unaware that we are separate from those around us. For some, especially those in more collectivist cultures,
this frame of mind may last a lifetime. For others, there is a dawning awareness of our separateness. What we
do with this discovery depends on the quality of our relationships, our personalities, and our life experiences.
For some, the awareness of separateness drives us to fear, terror, and despair. For others, solitude and self-
reflection become a cherished retreat from the demands and chaos of the outer world.
Learning to look inward to explore the landscape of our internal worlds isn’t something we are born with.
It requires time, discipline, and more than a little courage. Fortunately, traditions of meditation and prayer
from around the world give us a place to begin. Upon turning inward, we soon discover that our minds are
erratic and unsteady instruments. We become aware that our minds shift among different perspectives,
emotional states, and modes of language. Recall for a moment the three kinds of inner language that arise
during different states of mind: a reflexive social language, an internal narrator, and a language of self-
reflection.
Well into our therapeutic relationship, Shaun sat across from me looking exasperated and hopeless. “Why
am I still haunted by these voices? My life is great. I’ve made it, so why can’t I just enjoy it? When the voices
aren’t criticizing me, they are second guessing everything I do.” I knew Shaun well, and these voices weren’t a
sign of psychosis. “I know—they suck, don’t they!” I replied. “You hear them too?” he asked. I responded,
“Absolutely. Never remember a time without them.” “Well, where do they come from and where do I go for
an exorcism?” he asked. I smiled and said, “I have a theory.”
“The human brain has a long and complex evolutionary history and works in mysterious ways. It also
doesn’t come with an owner’s manual, so we are still in the process of figuring it out. I’ve come to believe that
these voices are a kind of archaeological artifact. What I mean is that they may have once served a survival
purpose at some stage of our evolution but have now become a nuisance. Each of us has two brains, one on
the left and one on the right. Long ago, primates had brains that were largely the same on both sides, but as
brains became larger and more complicated, each hemisphere began to specialize in different skills and
abilities. The right hemisphere is in control of very high and very low levels of emotion (terror and shame) and
is likely a model for what both hemispheres were once like. It also has an early developmental spurt in the first
18 months of life and connects with caretakers for the purposes of attachment, emotional regulation, and
acculturation into our family and tribe.
“The left hemisphere was the experiment that created modern humans. It veered off from this path in
order to specialize in later-evolving abilities such as language, rational thought, social interactions, and self-
awareness. Both hemispheres have language; the left hemisphere has the language we use to think through
problems and to communicate with others. The right has a language that is usually fearful and negative in
tone and programmed very early in our lives. It is the worrier, the critic, and the one designed to keep you in
line.”
“Well, that does suck!” Shaun said. “But why are they so negative?”
“Well, we know that the right hemisphere is biased toward negativity, and people with more activation in
the right than the left prefrontal cortex tend to be depressed. This probably evolved because worrying about
what others in the tribe were thinking about you most likely correlated with sustained connection and survival.
The right hemisphere is completely wrapped up in survival, the way both of our hemispheres once were. My
best guess is that the voices in our heads are the right-hemisphere remnants of voices of parents and tribal
leaders that supported group coordination, cooperation, and cohesion, something Freud called the superego,
an inner supervisor of the self that echoes early conditioning.

229
“Remember that the purpose of the brain is to enhance survival through the prediction and control of
future outcomes. In a social context, the right hemisphere leverages our fear of being shamed to keep us in line
and obedient to the alphas; am I acceptable to others? Am I going to get fired? And so on. Concerns about
being accepted by the group appear to have been woven into our genes, brains, and minds. Some of us have
especially harsh and critical inner voices that never let up. This may be because we had critical parents, have a
bias toward depression, or lack self-confidence and feel ashamed of who we are for other reasons.
“Because these voices seem to come from deep inside of us, we don’t experience them as memories but
rather as part of our self; a very painful and unfortunate misunderstanding of how our brains work. A central
aspect of becoming the CEO of our lives is to see the voices as primitive memory programming and learn how
to interpret, manage, and mitigate their negative effects. I’m not sure they ever go away, perhaps because we
may still need them in some instances when their input is actually helpful. But we all need to discern which of
these voices are counterproductive and learn to tell the destructive ones to get lost. This is one way that
understanding how your brain evolved and developed can help to turn your mind into an ally.”
I’m sure that Shaun had heard the message of not paying attention to these thoughts many times before.
But reasoning with the voices doesn’t seem to work—they are deeper, more primitive, and stronger than
conscious thought. Whether or not the explanation I presented to him is correct, this way of thinking about
his inner voices captured his imagination. This scientific narrative created a way for him to objectify his enemy
—these shaming and critical right-hemisphere voices—and develop a range of strategies to fight them. I
didn’t tell Shaun that he was being irrational. I was saying that the voices in his head were about his parents
and tribe, and they needed to be separated from his sense of self. He now had an enemy to tackle, not just
with his mind but with his body and soul.

Reframing Shame
Nothing you have done is wrong, and nothing you can do can make up for it.
—Gershen Kaufman

Over the last half-century, shame has emerged as the most prevalent and powerful emotion shaping our
clients’ personal narratives. Jung’s shadow, Freud’s neurosis, and Beck’s depression can all be traced to the
origins of shame: abandonment, exploitation, criticism, and myriad other early experiences that make us doubt
our value and legitimacy as a person. In recent decades, a range of disciplines have revealed the importance of
core shame and explored its neurobiological, cognitive, and behavioral dimensions. Shame has come out of the
closet and into the spotlight as a topic of TED talks, best sellers, and new therapies.
While the few who engage in heinous crimes appear to have little or no shame, many who have done
nothing wrong are crushed by it. This incongruity makes us question shame’s evolutionary origins and survival
value. One would think that any human phenomenon so prevalent and so powerful must have some survival
value to counterbalance all of the pain it causes. Although we experience shame as a private and deeply
personal emotion, its survival value may be the result of natural selection at the group level. Let me explain.
As we evolved into social animals, larger and larger groups of individuals needed to cooperate, coordinate
their behavior, and follow a leader. But how is this accomplished by a group of prehumans without culture,
language, or rational thought? How about making most everyone feel anxious and avoidant of the spotlight so
they pay close attention to the thoughts, feelings, and behaviors of an alpha? The value of shame may be to
make us focus on what others are thinking about us in order to maintain cohesive and well-organized groups.
By making us uncertain about ourselves, shame leads us to look to others to make sure we are acceptable and
doing the right thing. This could account not only for how groups are able to organize behind a leader but
also for the power of celebrity, cult leaders, and dictators. The fact that we feel badly about ourselves as a
result of shame may be an accidental by-product of this primitive method of social organization.
If you were going to design a mechanism of social control that would keep betas in line behind alphas, you
couldn’t do better than to instill them with a sense of shame. Those who are psychologically organized around
shame are always worried about what others think of them, try to be perfect in the eyes of others, and only feel
safe and confident when they are following someone else’s orders. What Freud called superego, the inner
narrator, the voice in our head that reminds us of our shortcomings, reinforces the same type of societal
hierarchies that we witness in all mammals.
Understanding and reframing shame is important for change because it is such a powerful agent of the
status quo. We can spend years in therapy trying to discover the origins of our shame—what we have done
wrong and what wrongs have been done to us. Sometimes we can find the smoking gun, and sometimes we

230
can’t. Either way, it doesn’t really make a difference. We have to root out the manifestations of shame in our
day-to-day lives and slowly and systematically alter them, one by one. The manifestations of shame,
perfectionism, low self-esteem, powerlessness, and overvaluing the opinions of others need to be addressed
and combated with energy, assertiveness, and courage. Those of us programmed as betas have to learn that
core shame is a primitive form of social organization and not take it personally. Core shame is not about us as
individuals, it’s a negative consequence of being a social mammal with a large cortex.
An excellent first step to managing shame is to let go of perfectionism and understand that ignorance is a
high state of consciousness. When the oracle at Delphi told Socrates that he was the wisest of men, he
assumed that the oracle was having a bad day because Socrates was certain of his own ignorance. Later, while
watching the folly of those convinced of their knowledge, he realized that the oracle knew that his awareness
of his ignorance was a sure sign of wisdom. This insight is a core teaching of Buddhism in its focus on seeing
past the illusions generated by the mind.
People with core shame spend a great deal of energy avoiding risks that might result in failure, and not
taking risks ensures that change won’t happen. As in many other human struggles, what we resist, persists,
and our fear controls us from the shadows. For change to happen, those with core shame have to convert
mistakes and imperfections from evidence of their lack of value into opportunities for new learning. It is
always amazing to me how many students have to tell me what they know because they are too frightened to
learn something new from me. This is especially salient given the price of tuition.
While Buddhism contains many valuable life lessons, one of the most important is the difference between
pain and suffering. Pain is woven into nature and is an inevitable part of life. To be alive and to love naturally
leads to aging, loss, and death. By contrast, suffering is the anguish we experience from worrying about the
future and regretting what has or hasn’t happened in the past. Shame is a primary cause of suffering. It is
relentless, and the inner narrator never runs out of things to make us feel badly about. Coming to grips with
past trauma, taking on the risk of connecting with others, and turning our minds into our allies are the levers
of change we all need to master to alleviate suffering.
We are simultaneously social and isolated creatures: embedded in our groups and our minds. Because we
are social creatures, we become afraid when we feel isolated, making our shame feel all the more
overwhelming. The presence of another, such as a therapist, allows us to feel safe enough to activate neural
plasticity and change our brains and minds. Feeling safe requires that we become familiar with our inner
world, root out and battle our demons, domesticate our minds, and turn them into allies. How do people
change? We change by connecting with others while cultivating a deeper relationship with ourselves.

Summary

Thank you for taking this journey with me through the neuroscientific foundations of psychotherapy. I hope
that some of what you have read here has helped you to better understand things that you may have found
confusing and has inspired you to think more deeply about your work with your clients. I also hope to have
inspired you to continue your search for better and more useful knowledge. Never stop learning!

231
Credits

Table 6.1: 1. Davidson & Fox, 1982. 2. Canli et al., 1998. 3. Wheeler, Davidson, & Tomarken, 1993. 4. Davidson & Fox, 1982; Harmon-

Jones & Allen, 1998. 5. Tomarken, Davidson, Wheeler, & Dass, 1992. 6. Coan, Allen, & Harmon-Jones, 2001; Davidson et al., 1990; Ekman

& Davidson, 1993. 7. Urry et al., 2004. 8. Fox & Davidson, 1988. 9. Harmon-Jones & Allen, 1998. 10. Harmon-Jones & Sigelman, 2001. 11.

Harmon-Jones & Sigelman, 2001. 12. Davidson et al., 1990. 13. Fox & Davidson, 1988. 14. Canli et al., 1998. 15. Davidson & Fox, 1982. 16.

Wheeler et al., 1993. 17. Kalin, Shelton, Davidson, & Kelley, 2001. 18. Fox & Davidson, 1988. 19. Davidson & Fox, 1989.

Table 7.1: 1. Minagawa-Kawai et al., 2008; Nitschke et al., 2004. 2. Berthoz et al., 2002; Mitchell, Banaji, & Macrae, 2005. 3. Mitchell,

Macrae, & Banaji, 2006. 4. Gusnard et al., 2001. 5. Goel & Dolan, 2001. 6. Frey & Petrides, 2000; Nobre et al., 1999. 7. Öngür & Price, 2000.
8. Bechara et al., 1998; Gallagher, McMahon, & Schoenbaum, 1999; Gehring & Willoughby, 2002; Kringelbach, 2005; Krueger et al., 2006;

O’Doherty, 2004. 9. Bechara et al., 1994; O’Doherty et al., 2002. 10. Matsumoto & Tanaka, 2004. 11. McGuire et al., 1996. 12. Dias,

Robbins, & Roberts, 1996; Simpson, Drevets, et al., 2001; Simpson, Snyder, et al., 2001; Quirk & Beer, 2006. 13. Malloy et al., 1993; Teasdale

et al., 1999; Beer et al., 2006. 14. Koechlin, Ody, & Kouneiher, 2003. 15. Dias et al., 1996; Fuster, 1997; Nagahama et al., 2001. 16. Knight &

Grabowecky, 1995. 17. Rezai et al., 1993; Petrides, Alivisatos, & Frey, 2002. 18. Henson, Shallice, & Dolan, 1999. 19. Lévesque, Eugène, et

al., 2003. 20. Pascual-Leone et al., 1996. 21. Kroger et al., 2002; Malloy et al., 1993; Teasdale et al., 1999. 22. Mitchell et al., 2006. 23. Gray,
Braver, & Raichle, 2002.

Table 7.3: 1. Pollatos, Gramann, & Schandry, 2007. 2. Liotti et al., 2001. 3. Klein et al., 2007. 4. Rosenkranz et al., 2005. 5. Lerner et al.,

2009. 6. Wicker et al., 2003. 7. Sanfey et al., 2003. 8. Berthoz et al., 2002. 9. Watson, Matthews, & Allman, 2007. 10. Liotti et al., 2000. 11.

Benuzzi et al., 2008. 12. Fan et al., 2011. 13. Decety et al., 2008, 2009. 14. Spence et al., 2001. 15. Shin et al., 2000. 16. Stoessel et al., 2011.
17. Najib et al., 2004. 18. Singer et al., 2006. 19. Jackson & Decety, 2004; Jackson et al., 2005; Saarela et al., 2007; Zaki et al., 2007. 20. Aziz-

Zadeh, Kaplan, & Iacoboni, 2009. 21. Dosenbach et al., 2007. 22. Chaudhry et al., 2009; Phan et al., 2002. 23. Xue et al., 2010.

Table 7.4: 1. Tamm, Menon, & Reiss, 2006. 2. Bush et al., 1999; Tamm, Menon, Ringel, & Reiss, 2004. 3. Tamm et al., 2004. 4. Rubia et al.,

1999. 5. Yu-Feng et al., 2007. 6. Tamm et al., 2004. 7. Zang et al., 2005. 8. Yu-Feng et al., 2007. 9. Lou, Henriksen, & Bruhn, 1984. 10. Lee

et al., 2005. 11. Castellanos et al., 2002. 12. Casey, Castellanos, & Giedd, 1997. 13. Li et al., 2007. 14. Markis et al., 2007. 15. Mackie et al.,

2007. 16. Ashtari et al., 2005.

Table 8.1: 1. Dehaene, Molko, Cohen, & Wilson, 2004. 2. Victor & Ropper, 2001. 3. Sirigu et al., 1996. 4. Colby, 1998; Driver & Mattingley,

1998. 5. Newman et al., 2003. 6. Rorden, Mattingley, Karnath, & Driver, 1997; Snyder & Chatterjee, 2004. 7. Karnath, 1997; Ungerleider &

Haxby, 1994. 8. Schwartz et al., 2005. 9. Battelli et al., 2001; Claeys et al., 2003. 10. Griffiths et al., 1998. 11. Snyder & Chatterjee, 2004. 12.

Andersen & Mountcastle, 1983. 13. Pia et al., 2004.

Table 8.2: 1. Griffiths et al., 1998. 2. Chochon et al., 1999. 3. Newman et al., 2003. 4. Uddin et al., 2005. 5. Dehaene et al., 2003. 6. Molko et

al., 2003. 7. Chochon et al., 1999. 8. Molko et al., 2003; Rushworth, Krams, & Passingham, 2001. 9. Newman et al., 2003. 10. Newman et al.,

2003. 11. Antal et al., 2008. 12. Grefkes & Fink, 2005; Wolpert, Goodbody, & Husain, 1998. 13. Wolpert et al., 1998. 14. Jonides et al., 1998.
15. Wagner et al., 2005. 16. Marshuetz et al., 2000; Van Opstal, Verguts, & Fias, 2008. 17. Husain & Nachev, 2007. 18. Astafiev et al., 2003.

19. Mountcastle, 1995. 20. Orban et al., 1999. 21. Castelli, Glaser, & Butterworth, 2006; Fias et al., 2003; Lemer et al., 2003. 22. Fias et al.,

2007. 23. Ruby & Decety, 2001. 24. Iacoboni et al., 2004; Jackson & Decety, 2004. 25. Vogeley et al., 2004.

Table 8.3: 1. Ferrari et al., 2003. 2. Heiser et al., 2003. 3. Cantalupo & Hopkins, 2001. 4. Chaminade et al., 2002. 5. Freund, 2001; Jeannerod

232
et al., 1995; 6. Buccino et al., 2001. 7. Iacoboni et al., 1999.

Table 9.1: 1. Buckner et al., 2008; Wagner, Haxby, & Heatherton, 2012. 2. Greicius & Menon, 2004; Ward et al., 2013. 3. Buckner et al.,
2008; Cavanna & Trimble, 2006; Laureys et al., 2006; Mason et al., 2007; Nielsen et al., 2005; Zysset et al., 2002.

Table 9.2: 1. Horovitz et al., 2009; Laureys et al., 2006. 2. Schneider et al., 2008; van Buuren et al., 2010. 3. Johnson et al., 2006; Whitfield-

Gabrieli et al., 2011. 4. VanBuuren et al., 2010; Yoshimura et al., 2009. 5. Li et al., 2014. 6. Fiset et al., 1999. 7. Jardri et al., 2007. 8.

Argembeau, 2011; Ostby et al., 2012; Pardo et al., 1993; Spreng & Grady, 2009. Maguire & Mummery, 1999. 9. Quin & Northoff, 2011. 10.

Iacoboni et al., 2004. 11. Li et al., 2014; Spreng & Grady, 2009. 12. Uddin et al., 2004. 13. Harrison et al., 2008; Moll et al., 2007. 14. Greicius

et al., 2003. 15. Greicius & Menon, 2004. 16. Andreasen et al., 1995. 17. Yang et al., 2010. 18. Hassabis et al., 2007. 19. Shapira-Lihter et al.,

2012. 20. Laufs et al., 2003. 21. McKiernan et al., 2003. 22. McGuire et al., 1996. 23. Ostby et al., 2012. 24. Spreng et al., 2009. 25. Spreng et
al., 2009.

Table 13.1: 1. Bredy et al., 2003; Champagne et al., 2008. 2. Weaver, Grant, & Meaney, 2002; Weaver et al., 2006. 3. Garoflos et al., 2008;

Menard, Champagne, & Meaney, 2004. 4. Liu et al., 2000. 5. Zhang et al., 2005. 6. Menard et al., 2004. 7. Weaver et al., 2004. 8. Caldji,

Diorio, & Meaney, 2003. 9. Caldji, Diorio, Anisman, & Meaney, 2004; Caldji et al., 1998. 10. Braun & Poeggel, 2001. 11. Champagne et al.,

2003. 12. Champagne et al., 2001, 2003, 2006. 13. Cameron, Fish, & Meaney, 2008.

Table 13.2: 1. Zhang at al., 2002. 2. Marais et al., 2008. 3. Leventopoulos et al., 2007. 4. Caldji et al., 2000; Hsu et al., 2003. 5. Rees, Steiner,

& Fleming, 2006. 6. Blaise et al., 2007. 7. Brake et al., 2004. 8. Kuhn & Schanberg, 1998. 9. Kalinichev et al., 2002. 10. Coutinho et al., 2002.
11. Weaver et al., 2006. 12. Ovtscharoff & Braun, 2001. 13. Akbari et al., 2007.

Table 13.3: 1. McCormick et al., 2000; Meaney et al., 1988, 1991; O’Donnell et al., 1994; Smythe, Rowe, & Meaney, 1994. 2. Sarrieau,

Sharma, & Meaney, 1988. 3. Plotsky & Meaney, 1993. 4. Kosten, Lee, & Kim, 2007. 5. Siviy & Harrison, 2008. 6. Garoflos et al., 2007. 7.

Vallée et al., 1997. 8. Vallée et al., 1999. 9. Costela et al., 1995; Tejedor-Real, Costela, & Gibert-Rahola, 1998. 10. Weaver et al., 2000. 11.
Collette et al., 2000.

Table 15.1: 1. Krugers et al., 2006. 2. Watanabe et al., 1992. 3. Alonso, 2000. 4. Sapolsky, 1990. 5. Dranovsky & Hen, 2006; Kelly, Mullany, &

Lynch, 2000; Pham et al., 2003; Prickaerts et al., 2004. 6. Kuhlmann, Piel, & Wolf, 2005; Kirschbaum et al., 1996; Newcomer et al., 1994,

1999. 7. West, 1993; Lupien et al., 1998. 8. Bremner, Scott, et al., 1993. 9. Bremner, Southwick, et al., 1993; Vythilingam et al., 2002. 10.

Bremner et al., 1995, 1997; Bremner, 2006; de Lanerolle et al., 1989. 11. Villarreal et al., 2002. 12. Falkai & Bogerts, 1986; Nelson et al., 1998.
13. Bourdeau et al., 2002; Condren & Thakore, 2001.

Table 16.1: 1. Brambilla, Soloff, & Sala, 2004; Bremner et al., 1997; Chugani et al., 2001; Dannlowski et al., 2012; Driessen et al., 2000; Gorka
et al., 2015; Hair et al., 2015; Hanson et al., 2015; Irle, Lange, & Sachsse, 2005; Lim, Radua, & Rubia, 2014; Rodrigues et al., 2011; Rogers et
al., 2011; Schmahl et al., 2003, 2009; Soloff et al., 2008; Tebartz van Elst et al., 2003; Teicher, Anderson, & Polcari, 2012; Weniger et al.,

2009; Zetzsche et al., 2007. 2. Shimada et al., 2015; Tomoda et al., 2009. 3. Brambilla et al., 2004; Chanen et al., 2008; Chugani et al., 2001;
Dannlowski et al., 2012; De Brito et al., 2009; Gorka et al., 2015; Hanson et al., 2010; Lim et al., 2014; Lyoo, Han, & Cho, 1998; Rogers et

al., 2011; Tomoda et al., 2009. 4. Chugani et al., 2001; Driessen et al., 2000; Hanson et al., 2015; Schmahl et al., 2003; Tottenham et al., 2010,

2011; Weniger et al., 2009. 5. Edmiston et al., 2011; Hair et al., 2015; Hanson et al., 2010, 2013a, 2013b; van Harmelen et al., 2010. 6. Bäuml

et al., 2014; Hanson et al., 2013; Huang, Gundapuneedi, & Rao, 2012; Smyser et al., 2010, 2013. 7. Brambilla et al., 2004; Brunner et al.,

2010; Goodman et al., 2011; Hazlett et al., 2005; Minzenberg et al., 2007; Tomoda et al., 2009; Whittle et al., 2009. 8. Koc ovská et al., 2013;
Nelson et al., 2013; Suor et al., 2015.

Table 20.1: 1. Bennett, Diamond, Krech, & Rosenzweig, 1964; Diamond et al., 1964. 2. Kempermann et al., 1998; Walsh, Budtz-Olsen,

Penny, & Cummins, 1969. 3. Kolb & Whishaw, 1998. 4. Kolb & Whishaw, 1998. 5. Kolb & Whishaw, 1998. 6. Ickes et al., 2000. 7. Nilsson

et al., 1993. 8. Sirevaag & Greenough, 1988. 9. Sirevaag & Greenough, 1988. 10. Guzowski, Setlow, Wagner, & McGaugh, 2001. 11.
Torasdotter et al., 1998.

233
Table 20.2: 1. Fujikawa et al., 2000. 2. Abercrombie et al., 2003; Andreano & Cahill, 2006; Domes et al., 2005. 3. Takahashi et al., 2004. 4.

Conrad, Lupien, & McEwen, 1999; Kerr, Huggett, & Abraham, 1994; Park et al., 2006; Yau et al., 1995. 5. Diamond et al., 1992. 6. Pavlides

et al., 1995. 7. Sullivan, Wilson, & Leon, 1989. 8. Introini-Collison & McGaugh, 1987.

Table 20.4: 1. Mayberg et al., 2002. 2. Mayberg et al., 2002. 3. Kong et al., 2006. 4. Pariente et al., 2005. 5. Petrovic et al., 2002. 6. Wager et

al., 2004. 7. Wager et al., 2004. 8. Zubieta et al., 2005. 9. Benedetti et al., 2004. 10. Fuente-Fernandez et al., 2001.

Table 21.1: 1. Baxter et al., 1992. 2. Brody et al., 1998. 3. Nakao et al., 2005. 4. Nakatani et al., 2003. 5. Yamanishi et al., 2008. 6. Saxena et al.,

2009. 7. Schwartz et al., 1996. 8. Lázaro et al., 2009. 9. Furmark et al., 2002. 10. Goldin et al., 2014. 11. Paquette et al., 2003. 12. Johanson et

al., 2006. 13. Straube et al., 2006. 14. Levin, Lazrove, & van der Kolk, 1999. 15. Prasko et al., 2004. 16. Sakai et al., 2006. 17. Beutel et al.,

2010. 18. Goldapple et al., 2004. 19. Kennedy et al., 2007. 20. Martin et al., 2001. 21. Brody, Saxena, Stoessel, et al., 2001. 22. Brody et al.,

1998. 23. Penades et al., 2002. 24. Wykes et al., 2002. 25. Laatsch et al., 1999.

234
References

Abbass, A., Nowoweiski, S., Bernier, D., Tarzwell, R., & Beutel, M. (2014). Review of psychodynamic psychotherapy neuroimaging
studies. Psychotherapy and Psychosomatics, 142–147. doi:10.1159/000358841
Abercrombie, H. C., Kalin, N. H., Thurow, M. E., Rosenkranz, M. A., & Davidson, R. J. (2003). Cortisol variation in humans affects
memory for emotionally laden and neutral information. Behavioral Neuroscience, 117(3), 505–516.
Adamec, R. E. (1991). Partial kindling of the ventral hippocampus: Identification of changes in limbic physiology which accompany
changes in feline aggression and defense. Physiology and Behavior, 49, 443–454.
Adams, R. D., Victor, M., & Ropper, A. H. (1997). Principles of neurology. New York: McGraw-Hill.
Adolphs, R. (2010). What does the amygdala contribute to social cognition? Annals of the New York Academy of Sciences, 1191(1), 42–61.
doi:10.1111/j.1749-6632.2010.05445.x
Ahern, G. L., Schomer, D. L., Kleefield, J., Blume, H., Rees-Cosgrove, G., Weintraub, S., & Mesulam, M. (1991). Right hemisphere
advantage for evaluating emotional facial expressions. Cortex, 27, 193–202.
Ainsworth, M. D. S., Blehar, M. C., Waters, E., & Wall, S. (1978). Patterns of attachment: A psychological study of the strange situation.
Hillsdale, NJ: Erlbaum.
Akaneya, Y., Tsumoto, T., Kinoshita, S., & Hatanaka, H. (1997). Brain-derived neurotrophic factor enhances long-term potentiation in rat
visual cortex. Journal of Neuroscience, 17, 6707–6716.
Akbari, E., Chatterjee, D., Levy, F., & Fleming, A. (2007). Experience-dependent cell survival in the maternal rat brain. Behavioral
Neuroscience, 121(5), 1001–1011.
Akers, K. G., Martinez-Canabal, A., Restivo, L., Yiu, A. P., Cristofaro, A. D., Hsiang, H., . . . Frankland, P. W. (2014). Hippocampal
neurogenesis regulates forgetting during adulthood and infancy. Science, 344(6184), 598-602. doi:10.1126/science.1248903
Akirav, I., & Maroun, M. (2007). The role of the medial prefrontal cortex-amygdala circuit in stress effects on the extinction of fear. Neural
Plasticity, 2007, 30873. doi:10.1155/2007/30873.
Alberini, C. M. (2005). Mechanisms of memory stabilization: Are consolidation and reconsolidation similar or distinct processes? Trends in
Neuroscience, 28(1), 51–56.
Alberini, C. M., & LeDoux, J. E. (2013). Memory reconsolidation. Current Biology, 23(17), R746-R750. doi:10.1016/c2010-0-67992-2
Alcoholics Anonymous. (2015). Welcome to Alcoholics Anonymous. Retrieved from http://www.aa.org.
Aleman, A., & Kahn, R. (2005). Strange feelings: Do amygdala abnormalities dysregulate the emotional brain in schizophrenia? Progress in
Neurobiology, 77, 283–298. doi:10.1016/j.pneurobio.2005.11.005
Alexander, G. E., DeLong, M. R., & Strick, P. L. (1986). Parallel organization of functionally segregated circuits linking basal ganglia and
cortex. Annual Review of Neuroscience, 9, 357–381.
Alexander, G. E., Furey, M. L., Grady, C. L., Pietrini, P., Brady, D. R., Mentis, M. J., & Schapiro M. (1997). Association of premorbid
intellectual function with cerebral metabolism in Alzheimer’s disease: Implications for the cognitive reserve hypothesis. American Journal of
Psychiatry, 154, 165–172.
Alexander, M. P., Stuss, D. T., & Benson, D. F. (1979). Capgras syndrome: A reduplicative phenomenon. Neurology, 29, 334–339.
Allen, J. P., McElhaney, K. B., Kuperminc, G. P., & Jodl, K. M. (2004). Stability and change in attachment security across adolescence.
Child Development, 75(6), 1792–1805.
Allen, J. S., Bruss, J., & Damasio, H. (2005). The aging brain: The cognitive reserve hypothesis and hominid evolution. American Journal of
Human Biology, 17, 673–689.
Allen, J. S., Damasio, H., & Grabowski, T. J. (2002). Normal neuroanatomical variation in the human brain: An MRI-volumetric study.
American Journal of Physical Anthropology, 118, 341–358.
Allen, N. J., & Barres, B. A. (2005). Signaling between glia and neurons: Focus on synaptic plasticity. Current Opinion in Neurobiology, 15,
542–548.
Allman, J. M., Hakeem, A., Erwin, J. M., Nimchinsky, E., & Hof, P. (2001). The anterior cingulate cortex: The evolution of an interface
between emotion and cognition. Annals of the New York Academy of Sciences, 935, 107–117.
Allman, J. M., Hakeem, A., Erwin, J. M., Nimchinsky, E., & Hof, P. (2006). The anterior cingulate cortex the evolution of an interface
between emotion and cognition. Annals of the New York Academy of Sciences, 935(1), 107-117. doi:10.1111/j.1749-6632.2001.tb03476.x

235
Allman, J. M., Watson, K. K., Tetreault, N. A., & Hakeem, A. Y. (2005). Intuition and autism: A possible role for Von Economo neurons.
Trends in Cognitive Sciences, 9, 367–373.
Allman, J., Rosin, A., Kumar, R., & Hasenstaub, A. (1998). Parenting and survival in anthropoid primates: Caretakers live longer.
Proceedings of the National Academy of Sciences, USA, 95, 6866–6869.
Al-Mousawi, A. H., Evans, N., Ebmeier, K. P., Roeda, D., Chaloner, F., & Ashcroft, G. W. (1996). Limbic dysfunction in schizophrenia
and mania: A study using 18F-labeled fluorodeoxyglucose and positron emission tomography. British Journal of Psychiatry, 169, 509–516.
Alonso, G. (2000). Prolonged corticosterone treatment of adult rats inhibits the proliferation of oligodendrocyte progenitors present
throughout white and gray matter regions of the brain. Glia, 31, 219–231.
Alonso, M., Vianna, M. R. M., Depino, A. M., e Mello Souza, T. M., Pereira, P., Szapiro, G., . . . Medina, J. H. (2002). BDNF-triggered
events in the rat hippocampus are required for both short- and long-term memory. Hippocampus, 12, 551–560.
Altman, J., Wallace, R. B., Anderson, W. J., & Das, G. D. (1968). Behaviorally induced changes in length of cerebrum in rats.
Developmental Psychobiology, 1, 112–117.
Amaral, D. G. (2002). The primate amygdala and the neurobiology of social behavior: Implications for understanding social anxiety.
Biological Psychiatry, 51, 11–17.
American Psychiatric Association. (2000). Diagnostic and statistical manual of mental disorders (4th ed., text rev.). Washington, DC:
American Psychiatric Association.

American Psychiatric Association (2013). Diagnostic and statistical manual of mental disorders, 5th ed. Washington, DC, American Psychiatric
Association.
Amini, F., Lewis, T., Lannon, R., Louie, A., Baumbacher, G., McGuinness, T., & Schiff, E. Z. (1996). Affect, attachment, memory:
Contributions toward psychobiologic integration. Psychiatry, 59(3), 213–239.
Amodio, D. M., Jost, J. T., Master, S. L., & Yee, C. M (2007). Neurocognitive correlates of liberalism and conservatism. Nature
Neuroscience, 10(10), 1246–1247.
Andersen, R. A., & Cui, H. (2009). Intention, action planning, and decision making in parietal-frontal circuits. Neuron, 63(5), 568–583.
doi:10.1016/j.neuron.2009.08.028
Andersen, R. A., & Mountcastle, V. B. (1983). The influence of the angle of gaze upon the excitability of the light-sensitive neurons of the
posterior parietal cortex. Journal of Neuroscience, 3(3), 532–548.
Andersen, R. A., Snyder, L. H., Bradley, D. C., & Xing, J. (1997). Multi-modal representation of space in the posterior parietal cortex and
its use in planning movements. Annual Review of Neuroscience, 20, 303–330.
Anderson, A. K., Wais, P. E., & Gabrieli, J. D. E. (2006). Emotion enhances remembrance of neutral events past. Proceedings of the
National Academy of Sciences, USA, 103(5), 1599–1604.
Anderson, C. M., Polcari, A., Lowen, S. B., Renshaw, P. F., & Teicher, M. H. (2002). Effects of methyphenidate on functional magnetic
resonance relaxometry of the cerebellar vermis in boys with ADHD. American Journal of Psychiatry, 159(8), 1322–1328.
Anderson, C. R. (1976). Coping behaviors as intervening mechanisms in the inverted-U stress-performance relationship. Journal of Applied
Psychology, 61(1), 30–34.
Anderson, J. S., Ferguson, M. A., Lopez-Larson, M., & Yurgelun-Todd, D. (2011). Connectivity gradients between the default mode and
attention control networks. Brain Connectivity, 1(2), 147-157. doi:10.1089/brain.2011.0007
Anderson, M. C., & Green, C. (2001). Suppressing unwanted memories by executive control. Nature, 410, 366–369.
Anderson, N. D., Ebert, P. L., Jennings, J. M., Grady, C. L., Cabeza, R., & Graham, S. J. (2008). Recollection and familiarity-based
memory in healthy aging and amnestic mild cognitive impairment. Neuropsychology, 22(2), 177–187. doi:10.1037/0894-4105.22.2.177
Andreano, J. M., & Cahill, L. (2006). Glucocorticoid release and memory consolidation in men and women. Psychological Science, 17(6),
466–470.
Andreasen, N. C. (2001). Brave new brain: Conquering mental illness in the era of the genome. New York: Oxford University Press.
Andreoni, J., Harbaugh, W., & Vesterlund, L. (2013). Altruism in experiments. In P. K. Newman (Ed.), The new Palgrave dictionary of
economics (2nd ed.), New York: Palgrave Macmillan.
Anisman, H., Zaharia, M. D., Meaney, M. J., & Merali, Z. (1998). Do early-life events permanently alter behavioral and hormonal
responses to stressors? International Journal of Developmental Neuroscience, 16, 149–164.
Ansorge, M. S., Zhou, M., Lira, A., Hen, R., & Gingrich, J. A. (2004). Early-life blockade of the 5-HT transporter alters emotional
behavior in adult mice. Science, 306, 879–881.
Antal, A., Baudewig, J., Paulus, W., & Dechent, P. (2008). The posterior cingulated cortex and planum temporale/parietal operculum are
activated by coherent visual motion. Visual Neuroscience, 25, 17–26.
Anticevic, A., Cole, M. W., Murray, J. D., Corlett, P. R., Wang, X., & Krystal, J. H. (2012). The role of default network deactivation in
cognition and disease. Trends in Cognitive Sciences, 16(12), 584-592. doi:10.1016/j.tics.2012.10.008

236
Apostolova, I., Block, S., Buchert, R., Osen, B., Conradi, M., Tabrizian, S., . . . Obrocki, J. (2010). Effects of behavioral therapy or
pharmacotherapy on brain glucose metabolism in subjects with obsessive-compulsive disorder as assessed by brain FDG PET. Psychiatry
Research: Neuroimaging, 184(2), 105–116. doi:10.1016/j.pscychresns.2010.08.012
Ardila, A., Ostrosky-Solis, F., Rosselli, M., & Gomez, C. (2000). Age-related cognitive decline during normal aging: The complex effect of
education. Archives of Clinical Neuropsychology, 15, 495–513.
Arnsten, A. F. T. (2000). Genetics of childhood disorders: XVIII. ADHD, Part 2: Norepinephrine has a critical modulatory influence on
prefrontal cortical function. Journal of the American Academy of Child and Adolescent Psychiatry, 39, 374–383.
Arnsten, A. F. T., & Goldman-Rakic, P. S. (1998). Noise stress impairs pre-frontal cortical cognitive function in monkeys: Evidence for a
hyperdopaminergic mechanism. Archives of General Psychiatry, 55, 362–368.
Arnsten, A. F. T., & Li, B. (2005). Neurobiology of executive functions: Catecholamine influences on prefrontal cortical functions.
Biological Psychiatry, 57(11), 1377–1384.
Ashtari, M., Kumra, S., Bhaskar, S. L., Clarke, T., Thaden, E., Cervellione, K. L., . . . Ardekani, B. A. (2005). Attention-
deficit/hyperactivity disorder: A preliminary diffusion tensor imaging study. Biological Psychiatry, 57(5), 448-455.
doi:10.1016/j.biopsych.2004.11.047
Astafiev, S., Shulman, G., Stanley, C., Snyder, A., Van Essen, D., & Corbetta, M. (2003). Functional organization of human intraparietal
and frontal cortex for attending, looking and pointing. Journal of Neuroscience, 23(11), 4689–4699.
Aston-Jones, G., Valentino, R. J., VanBockstaele, E. J., & Meyerson, A. T. (1994). Locus coeruleus, stress, and PTSD: Neurobiology and
clinical parallels. In M. M. Murburg (Ed.), Catecholamine function in posttraumatic stress disorder: Emerging concepts (pp. 17–62). Washington,
DC: American Psychiatric Press.
Augustine, J. R. (1996). Circuitry and functional aspects of the insular lobe in primates including humans. Brain Research Reviews, 22, 229–
244.
Aziz-Zadeh, L., Kaplan, J. T., & Iacoboni, M. (2009). “Aha!”: The neural correlates of verbal insight solutions. Human Brain Mapping, 30,
908–916.
Baars, B. J. (2002). The conscious access hypothesis: Origins and recent evidence. Trends in Cognitive Sciences, 6(1), 47–52.
Bachar, E., Kanyas, K., Latzer, Y., Canetti, L., Bonne, O., & Lerer, B. (2008). Depressive tendencies and lower levels of self-sacrifice in
mothers, and selflessness in their anorexic daughters. European Eating Disorders Review, 16(3), 184–190.
Bagby, R. M., & Taylor, G. J. (1997). Affect dysregulation and alexithymia. In G. J. Taylor, R. M. Bagby, & J. D. A. Parker (Eds.),
Disorders of affect regulation: Alexithymia in medical and psychiatric illness (pp. 26–45). Cambridge: Cambridge University Press.
Baillieux, H., De Smet, H. J., Paquier, P. F., De Deyn, P. P., & Marien, P. (2008). Cerebellar neurocognition: Insights into the bottom of
the brain. Clinical Neurology and Neurosurgery, 110, 763–773.
Baird, A. A., Gruber, S. A., Fein, D. A., Maas, L. C., Steingard, R. J., Renshaw, P. F., . . . Yurgelun-Todd, D. (1999). Functional
magnetic resonance imaging of facial affect recognition in children and adolescents. Journal of the American Academy of Child and Adolescent
Psychiatry, 38(2), 195–199.
Balconi, M., Bortolotti, A., & Gonzaga, L. (2011). Emotional face recognition, EMG response, and medial prefrontal activity in empathic
behaviour. Neuroscience Research, 71(3), 251–259. doi:10.1016/j.neures.2011.07.1833
Baldi, E., & Bucherelli, C. (2005). The inverted “U-shaped” dose-effect relationships in learning and memory: Modulation of arousal and
consolidation. Nonlinearity in Biology, Toxicology, and Medicine, 3(1), 9–21.
Balthazar, M. L., Campos, B. M., Franco, A. R., Damasceno, B. P., & Cendes, F. (2014). Whole cortical and default mode network mean
functional connectivity as potential biomarkers for mild Alzheimer’s disease. Psychiatry Research: Neuroimaging, 221(1), 37-42.
doi:10.1016/j.pscychresns.2013.10.010
Baptista, L. F., & Petrinovich, L. (1986). Song development in the white-crowned sparrow: Social factors and sex differences. Animal
Behavior, 34(5), 1359–1371.
Bar, M., Kassam, K. S., Ghuman, A. S., Boshyan, J., Schmid, A. M., Dale, A. M., . . . Halgren, E. (2006). Top-down facilitation of visual
recognition. Proceedings of the National Academy of Sciences, USA, 103(2), 449–454.
Barad, M. (2000). A biological analysis of transference. Paper presented at the UCLA Annual Review of Neuropsychiatry, February 2, Indian
Wells, California.
Barbas, H. (1995). Anatomic basis of cognitive-emotional interactions in the primate prefrontal cortex. Neuroscience and Biobehavioral
Reviews, 19(3), 499–510.
Barbey, A., Colom, R., Paul, E., & Grafman, J. (2012). Architecture of fluid intelligence and working memory revealed by lesion mapping.
Brain Structure and Function, 219(2), 485–494. doi:10.1007/s00429-013-0512-z
Barde, Y. A. (1989). Trophic factors and neuronal survival. Neuron, 2(6), 1525–1534.
Bargh, J. A., & Chartrand, T. L. (1999). The unbearable automaticity of being. American Psychologist, 54(7), 462–479.

237
Barnow, S., Völker, K. A., Möller, B., Freyberger, H. J., Spitzer, C., Grabe, H. J., & Daskalakis, Z. J. (2009). Neurophysiological correlates
of borderline personality disorder: A transcranial magnetic stimulation study. Biological Psychiatry, 65, 313–318.
Barry, R. J., Clarke, A. R., McCarthy, R., Selikowitz, M., Johnstone, S. J., & Rushby, J. A. (2004). Age and gender effects in EEG
coherence: I. Developmental trends in normal children. Clinical Neurophysiology, 115(10), 2252–2258.
Barsaglini, A., Sartori, G., Benetti, S., Pettersson-Yeo, W., & Mechelli, A. (2013). The effects of psychotherapy on brain function: A
systematic and critical review. Progress in Neurobiology, 114, 1–14. doi:10.1016/j.pneurobio.2013.10.006
Bartels, A., & Zeki, S. (2000). The neural basis of romantic love. NeuroReport, 11(17), 3829–3834.
Bartolomeo, P. (2006). A parietofrontal network for spatial awareness in the right hemisphere of the human brain. Archives of Neurology,
63(9), 1238–1241.
Bartzokis, G., Beckson, M., Lu, P. H., Nuechterlein, K. H., Edwards, N., & Mintz, J. (2001). Age-related changes in frontal and temporal
lobe volumes in men. Archives of General Psychiatry, 58(5), 461–465.
Bateson, G. (1972). Steps to an ecology of mind. New York: Ballantine.
Battelli, L., Cavanagh, P., Intrilligator, J., Tramo, M. J., Henaff, M., Michel, F., & Barton, J. J. S. (2001). Unilateral right parietal damage
leads to bilateral deficit for high-level motion. Neuron, 32(6), 985–995.
Bauer, P. J. (2015). A complementary processes account of the development of childhood amnesia and a personal past. Psychological Review,
122(2), 204–231. doi:10.1037/a0038939
Bauer, P. M., Hanson, J. L., Pierson, R. K., Davidson, R. J., & Pollak, S. D. (2009). Cerebellar volume and cognitive functioning in
children who experienced early deprivation. Biological Psychiatry, 66(12), 1100–1106. doi:10.1016/j.biopsych.2009.06.014
Bäuml, J. G., Daamen, M., Meng, C., Neitzel, J., Scheef, L., Jaekel, J., . . . Sorg, C. (2014). Correspondence between aberrant intrinsic
network connectivity and gray-matter volume in the ventral brain of preterm born adults. Cerebral Cortex, 25(11), 4135–4145.
doi:10.1093/cercor/bhu133
Baxter, L. R., Phelps, M. E., Mazziotta, J. C., Schwartz, J. M., Gerner, R. H., Selin, C. E., & Sumida, R. M. (1985). Cerebral metabolic
rates for glucose metabolism in mood disorders. Archives of General Psychiatry, 42, 441–447.
Baxter, L. R., Schwartz, J. M., Bergman, K. S., Szuba, M. P., Guze, B. H., Mazziotta, J. C., . . . Phelps, M. E. (1992). Caudate glucose
metabolic rate changes with both drug and behavior therapy for obsessive-compulsive disorder. Archives of General Psychiatry, 40(9), 681–689.
Baxter, L. R., Schwartz, J. M., Phelps, M. E., Mazziotta, J. C., Guze, B. H., Selin, C. E., . . . Sumida, R. M. (1989). Reduction of
prefrontal cortex glucose metabolism common to three types of depression. Archives of General Psychiatry, 46(3), 243–250.
Bazanis, E., Rogers, R. D., Dowson, J. H., Taylor, P., Meux, C., Staley, C., . . . Sahakian, B. J. (2002). Neurocognitive deficits in decision-
making and planning of patients with DSM-III-R borderline personality disorder. Psychological Medicine, 32, 1395–1405.
Beason-Held, L. L., Kraut, M. A., & Resnick, S. M. (2009). Stability of Default-Mode Network Activity in the Aging Brain. Brain
Imaging and Behavior, 3(2), 123-131. doi:10.1007/s11682-008-9054-z
Beatty, J. (2001). The human brain: Essentials of behavioral neuroscience. Thousand Oaks, CA: Sage.
Beauregard, M. (2007). Mind does really matter: Evidence from neuroimaging studies of emotional self-regulation, psychotherapy, and
placebo effect. Progress in Neurobiology, 81(4), 218–236.
Beauregard, M., Lévesque, J., & Bourgouin, P. (2001). Neural correlates of conscious self-regulation of emotion. Journal of Neuroscience,
21(RC165), 1–6.
Beblo, T., Saavedra, A. S., Mensebach, C., Lange, W., Markowitsch, H., Rau, H., . . . Driessen, M. (2006). Deficits in visual functions and
neuropsychological inconsistency in borderline personality disorder. Psychiatry Research, 145, 127–135.
Bechara, A., Damasio, A. R., Damasio, H., & Anderson, S. W. (1994). Insensitivity to future consequences following damage to human
prefrontal cortex. Cognition, 50(1–3), 7–15.
Bechara, A., Damasio, H., Tranel, D., & Anderson, S. W. (1998). Dissociation of working memory from decision making within the
human prefrontal cortex. Journal of Neuroscience, 18(1), 428–437.
Bechara, A., Damasio, H., Tranel, D., & Damasio, A. (1997). Deciding advantageously before knowing the advantageous strategy. Science,
275(5304), 1293–1295.
Bechara, A., & Naqvi, N. (2004). Listening to your heart: Interoceptive awareness as a gateway to feeling. Nature Neuroscience, 7(2), 102–
103.
Bechara, A., Tranel, D., Damasio, H., Adolphs, R., Rockland, C., & Damasio, A. R. (1995). Double dissociation of conditioning and
declarative knowledge relative to the amygdala and hippocampus in humans. Science, 269(5227), 1115–1118.
Beck, A. T. (1976). Cognitive therapy and emotional disorders. New York: International University Press.
Beck, A. T., Rush, A. J., Shaw, B. F., & Emery, G. (1979). Cognitive therapy of depression. New York: Guilford.
Beckmann, C. F., DeLuca, M., Devlin, J. T., & Smith, S. M. (2005). Investigations into resting-state connectivity using independent
component analysis. Philosophical Transactions of the Royal Society: Biological Sciences, 360(1457), 1001–1013.

238
Beeghly, M., & Cicchetti, D. (1994). Child maltreatment, attachment, and the self-system: Emergence of an internal state lexicon in
toddlers at high social risk. Development and Psychopathology, 6(1), 5–30.
Beer, J. S. (2007). The default self: Feeling good or being right? Trends in Cognitive Sciences, 11(5), 187-189. doi:10.1016/j.tics.2007.02.004
Beer, J. S., Heerey, E. A., Keltner, D., Scabini, D., & Knight, R. T. (2003). The regulatory function of self-conscious emotion: Insights
from patients with orbitofrontal damage. Journal of Personality and Social Psychology, 85(4), 594–604.
Beer, J. S., John, O. P., Scabini, D., & Knight, R. T. (2006). Orbitofrontal cortex and social behavior: Integrating self-monitoring and
emotion-cognition interactions. Journal of Cognitive Neuroscience, 18(6), 871–879.
Bekinschtein, P., Cammarota, M., Katche, C., Slipczuk, L., Rossato, J. I., Goldin, A., . . . Medina, J. H. (2008). BDNF is essential to
promote persistence of long-term memory storage. Proceedings of the National Academy of Sciences, USA, 105(7), 2711–2716.
Bell, M. A., & Fox, N. A. (1992). The relations between frontal brain electrical activity and cognitive development during infancy. Child
Development, 63(5), 1142–1163.
Belmaker, R. H., & Grisaru, N. (1999). Anti-bipolar potential for transcranial magnetic stimulation. Bipolar Disorders, 1(2), 71–72.
Benedetti, F., Colloca, L., Torre, E., Lanotte, M., Melcarne, A., Pesare, M., . . . Lopiano, L. (2004). Placebo-responsive Parkinson patients
show decreased activity in single neurons of subthalamic nucleus. Nature Neuroscience, 7(6), 587–588.
Benes, F. M. (1989). Myelination of cortical-hippocampal relays during late adolescence. Schizophrenia Bulletin, 15(4), 585–593.
Benes, F. M., Taylor, J. B., & Cunningham, M. C. (2000). Convergence and plasticity of monoaminergic systems in the medial prefrontal
cortex during the postnatal period: Implications for the development of psychopathology. Cerebral Cortex, 10(10), 1014–1027.
Bennett, A. J., Lesch, K. P., Heils, A., Long, J. C., Lorenz, J. G., Shoaf, S. E., . . . Higley, J. D. (2002). Early experience and serotonin
transporter gene variation interact to influence primate CNS function. Molecular Psychiatry, 7(1), 118–122.
Bennett, E. L., Diamond, M. C., Krech, D., & Rosenweig, M. R. (1964). Chemical and anatomical plasticity of brain. Science, 146(3644),
610–619.
Benson, F. D. (1994). The neurology of thinking. New York: Oxford University Press.
Benuzzi, F., Lui, F., Duzzi, D., Nichelli, P. F., & Porro, C. A. (2008). Does it look painful or disgusting? Ask your parietal and cingulate
cortex. Journal of Neuroscience, 28(4), 923–931.
Bermpohl, F., Pascual-Leone, A., Amedi, A., Merabet, L. B., Fregni, F., Wrase, J., Northoff, G. (2008). Novelty seeking modulates medial
prefrontal activity during the anticipation of emotional stimuli. Psychiatry Research, 164, 81–85. 10.1016/j.pscychresns.2007.12.019
Berntson, G. C., Bechara, A., Damasio, H., Tranel, D., & Cacioppo, J. T. (2007). Amygdala contribution to selective dimensions of
emotion. Social Cognitive and Affective Neuroscience, 2(2), 123–129.
Berthier, M. L., Posada, A., & Puentes, C. (2001). Dissociative flashbacks after right frontal injury in a Vietnam veteran with combat-
related posttraumatic stress disorder. Journal of Neuropsychiatry and Clinical Neurosciences, 13(1), 101–105.
Berthoz, S., Armony, J. L., Blair, R. J. R., & Dolan, R. J. (2002). An fMRI study of intentional and unintentional (embarrassing) violations
of social norms. Brain, 125(Pt. 8), 1696–1708.
Berton, O., McClung, C. A., DiLeone, R. J., Krishnan, V., Renthal, W., Russo, S. J., . . . Nestler, E. J. (2006). Essential role of BDNF in
the mesolimbic dopamine pathway in social defeat stress. Science, 311(5762), 864–868.
Beutel, M. E., Stark, R., Pan, H., Silbersweig, D., & Dietrich, S. (2010). Changes of brain activation pre-post short-term psychodynamic
inpatient psychotherapy: An fMRI study of panic disorder patients. Psychiatry Research: Neuroimaging, 184(2), 96–104.
doi:10.1016/j.pscychresns.2010.06.005
Bhide, P. G., & Bedi, K. S. (1982). The effects of environmental diversity on well-fed and previously undernourished rats: I. Body and brain
measurements. Journal of Comparative Neurology, 207(4), 403–409.
Bickart, K. C., Wright, C. I., Dautoff, R. J., Dickerson, B. C., & Barrett, L. F. (2010). Amygdala volume and social network size in
humans. Nature Neuroscience, 14, 163–164.
Birnbaum, S., Gobeske, K. T., Auerbach, J., Taylor, J. R., & Arnsten, A. F. T. (1999). A role for norepinephrine in stress-induced cognitive
deficits: ±-1-adrenoceptor mediation in the prefrontal cortex. Biological Psychiatry, 46(9), 1266–1274.
Bischoff-Grethe, A., Proper, S. M., Mao, H., Daniels, K. A., & Berns, G. S. (2000). Conscious and unconscious processing of nonverbal
predictability in Wernicke’s area. Journal of Neuroscience, 20(5), 1975–1981.
Bishof, H. (1983). Imprinting and cortical plasticity: A comparative review. Neuroscience and Biobehavioral Reviews, 7(2), 213–225.
Bishop, S. J. (2007). Neurocognitive mechanisms of anxiety: An integrative account. Trends in Cognitive Sciences, 11(7), 307–316.
Bishop, S., Duncan, J., & Lawrence, A. D. (2004). State anxiety of the amygdala response to unattended threat-related stimuli. Journal of
Neuroscience, 24(46), 10364–10368.
Bishop, S., Duncan, J., Brett, M., & Lawrence, A. (2004). Prefrontal cortical function and anxiety: Controlling attention to threat-related
stimuli. Nature Neuroscience, 7(2), 184–187.
Bisiach, E., & Luzzatti, C. (1978). Unilateral neglect of representational space. Cortex, 14(1), 129–133.

239
Bisiach, E., Rusconi, M. L., & Vallar, G. (1991). Remission of somatoparaphrenic delusions through vestibular stimulation.
Neuropsychologia, 29, 1029–1031.
Biver, F., Goldman, S., Francois, A., DeLaPorte, C., Luxen, A., Gribomont, B., & Lotstra, F. (1995). Changes in metabolism of cerebral
glucose after stereotactic leukotomy for refractory obsessive-compulsive disorder: A case report. Journal of Neurology, Neurosurgery and Psychiatry,
58(4), 502–505.
Black, J. E. (1998). How a child builds its brain: Some lessons from animal studies of neural plasticity. Preventive Medicine, 27(2), 168–171.
Blaise, J. H., Koranda, J. L., Chow, U., Haines, K. E., & Doward, E. C. (2007). Neonatal isolation stress alters bidirectional long-term
synaptic plasticity in amygdalo-hippocampal synapses in freely behaving adult rats. Brain Research, 1193(25–33), 25–33.
Blanke, O., & Arzy, S. (2005). The out-of-body experience: Disturbed self-processing at the temporo-parietal junction. Neuroscientist,
11(1), 11–24.
Blass, R. B., & Carmeli, Z. (2007). The case against neuropsychoanalysis: On fallacies underlying psychoanalysis’ latest scientific trend and
its negative impact on psychoanalytic discourse. International Journal of Psychoanalysis, 88(Pt. 1), 19–40.
Blonder, L. X., Bowers, D., & Heilman, K. M. (1991). The role of right hemisphere in emotional communication. Brain, 114(3), 1115–
1127.
Bluhm, R. L., Miller, J., Lanius, R. A., Osuch, E. A., Boksman, K., Neufeld, R., . . . Williamson, P. (2007). Spontaneous low-frequency
fluctuations in the BOLD signal in schizophrenic patients: Anomalies in the default network. Schizophrenia Bulletin, 33(4), 1004-1012.
doi:10.1093/schbul/sbm052
Bluhm, R. L., Williamson, P. C., Osuch, E. A., Frewen, P. A., Stevens, T. K., Boksman, K. . . . Lanius, R. A. (2009). Alterations in default
network connectivity in posttraumatic stress disorder related to early-life trauma. Journal Psychiatry and Neuroscience, 34(3), 187-194.
Blum, D. (2002). Love at Goon Park: Harry Harlow and the science of affection. Cambridge, MA: Perseus.
Boehm, C. (2008). Purposive social selection and the evolution of human altruism. Cross-Cultural Research, 42(4), 319–352.
Bohus, M., Limberger, M., Ebner, U., Glocker, F. X., Schwartz, B., Wernz, M., & Lieb, K. (2000). Pain perception during self-reported
distress and calmness in patients with borderline personality disorder and self-mutilating behavior. Psychiatry Research, 95, 251–260.
Bollas, C. (1987). The shadow of the object: Psychoanalytic of the unthought known. New York: Columbia University Press.
Bonda, E., Petrides, M., Frey, S., & Evans, A. C. (1994). Frontal cortex involvement in organized sequences of hand movements: Evidence
from a positron emission tomography study. Social Neuroscience Abstracts, 20, 353.
Bonda, E., Petrides, M., Ostry, D., & Evans, A. (1996). Specific involvement of human parietal systems and the amygdala in the perception
of biological motion. Journal of Neuroscience, 16(11), 3737–3744.
Bonne, O., Brandes, D., Gilboa, M. A., Gomori, M. A., Shenton, M. E., Pitman, R. K., & Shalev, A. Y. (2001). Longitudinal MRI study
of hippocampal volume in trauma survivors with PTSD. American Journal of Psychiatry, 158(8), 1–7. doi:10.1176/appi.ajp.158.8.1248
Bornstein, M. H. (1989). Sensitive periods in development: Structural characteristics and causal interpretations. Psychological Bulletin,
105(2), 179–197.
Borod, J. C., Cicero, B. A., Obler, L. K., Welkowitz, J., Erhan, H. M., Santschi, C., . . . Whalen, J. R. (1998). Right hemisphere emotional
perception: Evidence across multiple channels. Neuropsychology, 12(3), 446–458.
Bourdeau, I., Bard, C., Noel, B., Leclerc, I., Cordeau, M. P., Belair, M., . . . Lacroix, A. (2002). Loss of brain volume in endogenous
Cushing’s syndrome and its reversibility after correction hypercortisolism. Journal of Clinical Endocrinology and Metabolism, 87(5), 1949–1954.
Bowen, M. (1978). Family therapy in clinical practice. Northvale, NJ: Jason Aronson.
Bowlby, J. (1969). Attachment. New York: Basic Books.
Bowlby, J. (1988). A secure base: Clinical applications of attachment theory. London: Routledge.
Bowles, S., Choi, J., & Hopfensitz, A. (2003). The co-evolution of individual behaviors and social institutions. Journal of Theoretical Biology,
223(2003), 135–147.
Bradshaw, J. (1990). Homecoming: Reclaiming and championing your inner child. New York: Bantam.
Brake, W. G., Zhang, T. Y., Diorio, J., Meaney, M. J., & Gratton, A. (2004). Influence of early postnatal rearing conditions on
mesocorticolimbic dopamine and behavioural responses to psychostimulants and stressors in adult rats. European Journal of Neuroscience, 19(7),
1863–1874.
Brambilla, P., Soloff, P. H., & Sala, M. (2004). Anatomical MRI study of borderline personality disorder in patients. Psychiatry Research:
Neuroimaging, 131, 125–133.
Branchi, I., D’Andrea, I., Sietzema, J., Fiore, M., Di Fausto, V., Aloe, L., & Alleva, E. (2006). Early social enrichment augments adult
hippocampal BDNF levels and survival of BrdU-positive cells while increasing anxiety- and “depression”-like behavior. Journal of Neuroscience
Research, 83(6), 965–973.
Branchi, I., Francia, N., & Alleva, E. (2004). Epigenetic control of neurobehavioral plasticity: The role of neurotrophins. Behavioral
Pharmacology, 15(5–6), 353–362.

240
Braun, C., Scweizer, R., Elbert, T., Borbaumer, N., & Taub, E. (2000). Differential activation in somatosensory cortex for different
discrimination tasks. Journal of Neuroscience, 20(1), 446–450.
Braun, K., & Poeggel, G. (2001). Recognition of mother’s voice evokes metabolic activation in the medial prefrontal cortex and lateral
thalamus of Octodon degus pups. Neuroscience, 103(4), 861–864.
Braun, U., Schäfer, A., Walter, H., Erk, S., Romanczuk-Seiferth, N., Haddad, L., . . . Bassett, D. (2015). Dynamic reconfiguration of
frontal brain networks during executive cognition in humans. Proceedings of the National Academy of Sciences, 112(37), 11678–11683.
doi:10.1073/pnas.1422487112
Brázdil, M., Mareek, R., Urbánek, T., Ka párek, T., Mikl, M., Rektor, I., & Zeman, A. (2012). Unveiling the mystery of déjà vu: The
structural anatomy of déjà vu. Cortex, 48(9), 1240–1243. doi:10.1016/j.cortex.2012.03.004
Bredy, T. W., Grant, R. J., Champagne, D. L., & Meaney, M. J. (2003). Maternal care influences neuronal survival in the hippocampus of
the rat. European Journal of Neuroscience, 18(10), 2903–2909.
Bredy, T., Zhang, T., Grant, R., Diorio, J., & Meaney, M. (2004). Peripubertal environmental enrichment reverses the effects of maternal
care on hippocampal development and glutamate receptor subunit expression. European Journal of Neuroscience, 20(5), 1355–1362.
Bremner, J. (2006). Stress and brain atrophy. CNS and Neurological Disorders—Drug Targets, 5(5), 503–512.
Bremner, J. D., Innis, R., Southwick, S., Staib, L., Zoghbi, S., & Charney, D. (2000). Decreased benzodiazepine receptor binding in
prefrontal cortex in combat-related posttraumatic stress disorder. American Journal of Psychiatry, 157, 1120–1126.
Bremner, J. D., & Narayan, M. (1998). The effects of stress on memory and the hippocampus throughout the life cycle: Implications for
childhood development and aging. Development and Psychopathology, 10(4), 871–885.
Bremner, J. D., Randall, P., Scott, T., & Bronen, R. (1995). MRI-based measurement of hippocampal volume in patients with combat-
related posttraumatic stress disorder. American Journal of Psychiatry, 152(7), 973–983.
Bremner, J. D., Randall, P., Vermetten, E., Staib, L., Bronen, R. A., Mazure, C., . . . Charney, D. S. (1997). Magnetic resonance imaging-
based measurement of hippocampal volume in posttraumatic stress disorder related to childhood physical and sexual abuse: A preliminary
report. Biological Psychiatry, 41(1), 23–32.
Bremner, J. D., Scott, T. M., Delaney, R. C., Southwick, S. M., Mason, J. W., Johnson, D. R., . . . Charney, D. S. (1993). Deficits of
short-term memory in posttraumatic stress disorder. American Journal of Psychiatry, 150, 1015–1019.
Bremner, J. D., Southwick, S. M., Johnson, D. R., Yehuda, R., & Charney, D. S. (1993). Childhood physical abuse and combat-related
posttraumatic stress disorder in Vietnam veterans. American Journal of Psychiatry, 150(2), 235–239.
Brennan, K. A., & Shaver, P. R. (1995). Dimensions of adult attachment, affect regulation, and romantic relationship functioning.
Personality and Social Psychology Bulletin, 21(3), 267–283.
Brennan, P. A., Pargas, R., Walker, E. F., Green, P., Newport, D. J., & Stowe, Z. (2008). Maternal depression and infant cortisol:
Influences of timing, comorbidity and treatment. Journal of Child Psychology and Psychiatry, 49(10), 1099–1107.
Brewin, C. R. (2015). Re-experiencing traumatic events in PTSD: New avenues in research on intrusive memories and flashbacks. European
Journal of Psychotraumatology, 6, 1–5. doi:10.3402/ejpt.v6.27180
Brewin, C. R., Dalgleish, T., & Joseph, S. (1996). A dual representation theory of post-traumatic stress disorder. Psychological Research,
103(4), 670–686.
Brewin, C. R., & Smart, L. (2005). Working memory capacity and suppression of intrusive thoughts. Journal of Behavior Therapy and
Experimental Psychiatry, 36(1), 61–68.
Briere, J., Scott, D. C., & Weathers, F. (2005). Peritraumatic and persistent dissociation in the presumed etiology of PTSD. American
Journal of Psychiatry, 162(12), 2295–2301. doi:10.1176/appi.ajp.162.12.2295
Broadhurst, P. L. (1957). Emotionality and the Yerkes-Dodson law. Journal of Experimental Psychology, 54(5), 345–352.
Brodal, P. (1992). The central nervous system. New York: Oxford University Press.
Brodsky, B. S., Groves, S. A., Oquendo, M. A., Mann, J. J., & Stanley, B. (2006). Interpersonal precipitants and suicide attempts in
borderline personality disorder. Suicide and Life-Threatening Behavior, 36(3), 313–322. doi:10.1521/suli.2006.36.3.313
Brody, A. L., Mandelkern, M. A., Olmstead, R. E., Jou, J., Tiongson, E., Allen, V., . . . Cohen, M. S. (2007). Neural substrates of resisting
craving during cigarette cue exposure. Biological Psychiatry, 62(6), 642–651.
Brody, A. L., Saxena, S., Mandelkern, M. A., Fairbanks, L. A., Ho, M. L., & Baxter, L. R. (2001). Brain metabolic changes associated
with symptom factor improvement in major depressive disorder. Biological Psychiatry, 50(3), 171–178.
Brody, A. L., Saxena, S., Schwartz, J. M., Stoessel, P. W., Maidment, K., Phelps, M. E., & Baxter, L. R. (1998). FDG-PET predictors of
response to behavioral therapy and pharmacotherapy in obsessive compulsive disorder. Psychiatry Research: Neuroimaging, 84(1), 1–6.
Brody, A. L., Saxena, S., Stoessel, P., Gillies, L.A., Fairbanks, L. A., Alborzian, S., . . . Baxter, L. R. (2001). Regional brain metabolic
changes in patients with major depression treated with either paroxetine or interpersonal therapy. Archives of General Psychiatry, 58(7), 631–640.
Brody, H. (2000). The placebo response: Recent research and implications for family medicine. Journal of Family Practice, 49(7), 649–654.

241
Brosch, T., Sander, D., & Scherer, K. R. (2007). That baby caught my eye . . . Attention capture by infant faces. Emotion, 7(3), 685–689.
Brothers, L. (1997). Friday’s footprint. New York: Oxford University Press.
Brown, H. D., Rosslyn, S. M., Reiter, H. C., Baer, L., & Janice, M. A. (1994). Can patients with obsessive-compulsive disorder
discriminate between percepts and mental images? A signal detection analysis. Journal of Abnormal Psychology, 103(3), 445–454.
Brown, S. M., Henning, S., & Wellman, C. L. (2005). Mild, short-term stress alters dendritic morphology in rat medial prefrontal cortex.
Cerebral Cortex, 15(11), 1714–1722.
Brozzoli, C., Gentile, G., & Ehrsson, H. (2012). That’s near my hand! Parietal and premotor coding of hand-centered space contributes to
localization and self-attribution of the hand. Journal of Neuroscience, 32(42), 14573–14582.
Bruce, S. E., Buchholz, K. R., Brown, W. J., Yan, L., Durbin, A., & Sheline, Y. I. (2013). Altered emotional interference processing in the
amygdala and insula in women with Post-Traumatic Stress Disorder. NeuroImage: Clinical, 2, 43–49. doi:10.1016/j.nicl.2012.11.003
Bruder, G. E., Stewart, J. W., Mercier, M. A., Agosti, V., Leite, P., Donovan, S., & Quitkin, F. M. (1997). Outcome of cognitive-
behavioral therapy for depression: Relation to hemispheric dominance for verbal processing. Journal of Abnormal Psychology, 106(1), 138–144.
Bruner, J. S. (1990). Acts of meaning. Cambridge, MA: Harvard University Press.
Brunet, A., Orr, S. P., Tremblay, J., Robertson, K., Nader, K., & Pitman, R. K. (2008). Effects of post-retrieval propranolol on
psychophysiologic responding during subsequent script-given traumatic imagery in post-traumatic stress disorder. Journal of Psychiatric Research,
42, 503–506.
Brunner, R., Henze, R., Parzer, P., Kramer, J., Feigl, N., Lutz, K., . . . Stieltjes, B. (2010). Reduced prefrontal and orbitofrontal gray matter
in female adolescents with borderline personality disorder: Is it disorder specific? NeuroImage, 49, 114–120.
Bryant, R. A., Kemp, A. H., Felmingham, K. L., Liddell, B., Oliveri, G., Peduto, A., . . . Williams, L. M. (2008). Enhanced amygdala and
medial prefrontal activation during nonconscious processing of fear in posttraumatic stress disorder: An fMRI study. Human Brain Mapping,
29(5), 517–523.
Buccino, G., Binkofski, F., Fink, G. R., Fadiga, L., Fogassi, L., Gallese, V., . . . Freund, H. (2001). Action observation activates premotor
and parietal areas in a somatotopic manner: An fMRI study. European Journal of Neuroscience, 13(2), 400-404. doi:10.1111/j.1460-
9568.2001.01385.x
Buchanan, T. W., Tranel, D., & Adolphs, R. (2006). Impaired memory retrieval correlates with individual differences in cortisol response
but not autonomic response. Learning and Memory, 13(3), 382–387.
Buchheim, A., Erk, S., George, C., Kächele, H., Kircher, T., Martius, P., . . . Walter, H. (2008). Neural correlates of attachment trauma in
borderline personality disorder: A functional magnetic resonance imaging study. Psychiatry Research: Neuroimaging, 163, 223–235.
Buchheim, A., George, C., Liebl, V., Moser, A., & Benecke, C. (2007). Mimische affektivität von patientinnen mit einer borderline-
persönlichkeitsstörung während des adult attachment projective. Zeitschrift für Psychosomatische Medizin und Psychotherapie, 53, 339–354.
Buckner, J. D., Lopez, C., Dunkel, S., & Joiner, T. E. (2008). Behavior management training for the treatment of reactive attachment
disorder. Child Maltreatment, 13(3), 289-297. doi:10.1177/1077559508318396
Buckner, R. L., & Vincent, J. L. (2007). Unrest at rest: Default activity and spontaneous network correlations. NeuroImage, 37(4), 1091-
1096. doi:10.1016/j.neuroimage.2007.01.010
Buneo, C., & Andersen, R. (2006). The posterior parietal cortex: Sensorimotor interface for the planning and online control of visually
guided movements. Neuropsychologia, 44(13), 2594–2606. doi:10.1016/j.neuropsychologia .2005.10.011
Buonomano, D. V., & Merzenich, M. M. (1998). Cortical plasticity: From synapses to maps. Annual Review of Neuroscience, 21, 149–186.
Bush, G., Frazier, J. A., Rauch, S. L., Seidman, L. J., Whalen, P. J., Jenike, M. A., . . . Biederman, J. (1999). Anterior cingulate cortex
dysfunction in attention deficit/hyperactivity disorder revealed by fMRI and the counting Stroop. Biological Psychiatry, 45(12), 1542–1552.
Bush, G., Luu, P., & Posner, M. I. (2000). Cognitive and emotional influences in anterior cingulate cortex. Trends in Cognitive Sciences,
4(6), 215–222.
Bush, G., Valera, E. M., & Seidman, L. J. (2005). Functional neuroimaging of attention deficit/hyperactivity disorder: A review and
suggested future directions. Biological Psychiatry, 57(11), 1273–1284.
Bush, G., Vogt, B. A., Holmes, J., Dale, A. M., Greve, D., Jenike, M. A., & Rosen, B. R. (2002). Dorsal anterior cingulate cortex: A role
in reward-based decision making. Proceedings of the National Academy of Sciences, USA, 99(1), 507–512.
Butler, R. W., Braff, D. L., Rauch, J. L., Jenkins, M. A., Sprock, J., & Geyer, M. A. (1990). Physiological evidence of exaggerated startle
response in a sub-group of Vietnam veterans with combat-related PTSD. American Journal of Psychiatry, 147(10), 1308–1312.
Buuren, M. V., Gladwin, T. E., Zandbelt, B. B., Kahn, R. S., & Vink, M. (2010). Reduced functional coupling in the default-mode
network during self-referential processing. Human Brain Mapping, 31(8), 1117-1127. doi:10.1002/hbm.20920
Cabeza, R., Ciaramelli, E., & Moscovitch, M. (2012). Cognitive contributions of the ventral parietal cortex: An integrative theoretical
account. Trends in Cognitive Sciences, 16(6), 338–352. doi:10.1016/j.tics.2012.04.008
Cabeza, R., & St. Jacques, P. (2007). Functional neuroimaging of autobiographical memory. Trends in Cognitive Sciences, 11(5), 219–227.

242
Cacioppo, J. T., & Berntson, G. (Eds.). (2004). Social neuroscience: Key readings in social psychology. New York: Psychology Press.
Cacioppo, S., Bianchi-Demicheli, F., Bischof, P., DeZiegler, D., Michel, C. M., & Landis, T. (2013). Hemispheric specialization varies
with EEG brain resting states and phase of menstrual cycle. PLoS One, 8(4), e63196. doi:10.1371/journal.pone.0063196
Cahill, L., & McGaugh, J. L. (1998). Mechanisms of emotional arousal and lasting declarative memory. Trends in Neurosciences, 21(7), 294–
299.
Cahill, L., Prins, B., Weber, M., & McGaugh, J. L. (1994). Beta-adrenergic activation and memory for emotional events. Nature,
371(6499), 702–704.
Calder, A. J., Keane, J., Manly, T., Sprengelmeyer, R., Scott, S., Nimmo-Smith, S., & Young, A. W. (2003). Facial expression recognition
across the adult life span. Neuropsychologia, 41(2), 195–202.
Caldji, C., Diorio, J., Anisman, H., & Meaney, M. (2004). Maternal behavior regulated benzodiazepine/GABA receptor subunit expression
in brain regions associated with fear in BALB/c and C67BL/6 mice. Neuropsychopharmacology, 29(7), 1344–1352.
Caldji, C., Diorio, J., & Meaney, M. J. (2003). Variations in maternal care alter GABA-sub(A) receptor subunit expression in brain regions
associated with fear. Neuropsychopharmacology, 28, 1950–1959.
Caldji, C., Francis, D., Sharma, S., Plotsky, P. M., & Meaney, M. J. (2000). The effects of early rearing environment on the development
of GABA-sub(A) and central benzodiazepine receptor levels and novelty-induced fearfulness in the rat. Neuropsychopharmacology, 22(3), 219–
229.
Caldji, C., Tannenbaum, B., Sharma, S., Francis, D., Plotsky, P. M., & Meaney, M. J. (1998). Maternal care during infancy regulates the
development of neural systems mediating the expression of fearfulness in the rat. Proceedings of the National Academy of Sciences, USA, 95(9),
5335–5340.
Calvo, M. G., & Beltrán, D. (2014). Brain lateralization of holistic versus analytic processing of emotional facial expressions. NeuroImage,
92, 237-247. doi:10.1016/j.neuroimage.2014.01.048
Calvo-Merino, B., Glaser, D. E., Grezes, J., Passingham, R. E., & Haggard, P. (2005). Action observation and acquired motor skills: An
fMRI study with expert dancers. Cerebral Cortex, 15(8), 1243–1249.
Cameron, N., Fish, E., & Meaney, M. (2008). Maternal influences on the sexual behavior and reproductive success of the female rat.
Hormones and Behavior, 54(1), 178–184.
Cameron, N. M., Champagne, F. A., Parent, C., Fish, E. W., Ozaki-Kuroda, K., & Meaney, M. J. (2005). The programming of individual
differences in defensive responses and reproductive strategies in the rat through variations in maternal care. Neuroscience and Biobehavioral
Reviews, 29(4–5), 843–865.
Campbell, J. (1949). The hero with a thousand faces. Novato, CA: New World Library.
Canli, T., Desmond, J. E., Zhao, Z., Glover, G., & Gabrieli, J. D. E. (1998). Hemispheric asymmetry for emotional stimuli detected with
fMRI. NeuroReport, 9(14), 3233–3239.
Canli, T., & Lesch, K. (2007). Long story short: The serotonin transporter in emotion regulation and social cognition. Nature Neuroscience,
10(9), 1103–1109.
Canli, T., Qiu, M., Omura, K., Congdon, E., Haas, B. W., Amin, Z., . . . Lesch, K. P. (2006). Neural correlates of epigenesis. Proceedings of
the National Academy of Sciences, USA, 103(43), 16033–16038.
Cantalupo, C., & Hopkins, W. D. (2001). Asymmetric Broca’s area in great apes: A region of the ape brain is uncannily similar to one
linked with speech in humans. Nature, 414(6863), 505-505. doi:10.1038/35107134
Cappa, S., Sterzi, R., Vallar, G., & Bisiach, E. (1987). Remission of hemineglect and anosognosia during vestibular stimulation.
Neuropsychologia, 25(5), 775–782.
Cappas, N. M., Andres-Hyman, R., & Davidson, L. (2005). What psychotherapists can begin to learn from neuroscience: Seven principles
of a brain-based psychotherapy. Psychotherapy: Theory, Research, Practice, Training, 42(3), 374–383.
Carnevali, L., & Sgoifo, A. (2014). Vagal modulation of resting heart rate in rats: The role of stress, psychosocial factors, and physical
exercise. Frontiers in Physiology, 5(118), 1–12. doi:10.3389/fphys.2014.00118
Carr, L., Iacoboni, M., Dubeau, M. C., Mazziotta, J. C., & Lenzi, G. L. (2003). Neural mechanisms of empathy in humans: A relay from
neural systems for imitation to limbic areas. Proceedings of the National Academy of Sciences, USA, 100(9), 5497–5502.
Carswell, S. (1993). The potential for treating neurodegenerative disorders with NGF-inducing compounds. Experimental Neurology,
124(1), 36–42.
Carvey, P. M. (1998). Drug action in the central nervous system. New York: Oxford University Press.
Casey, B. J., Castellanos, F. X., & Giedd, J. N. (1997). Implications of right frontostriatal circuitry in response inhibition and attention-
deficit/hyperactivity disorder. Journal of the American Academy of Child and Adolescent Psychiatry, 36(3), 374–383.
Casey, B. J., Galvan, A., & Hare, T. A. (2005). Changes in cerebral functional organization during cognitive development. Current Opinion
in Neurobiology, 15(2), 239–244.

243
Cashmore, L., Uomini, N., & Chapelain, A. (2008). The evolution of handedness in humans and great apes: A review and current issues.
Journal of Anthropological Sciences, 86, 7-35.
Caspers, S., Schleicher, A., Bacha-Trams, M., Palomero-Gallagher, N., Amunts, K., & Zilles, K. (2012). Organization of the human
inferior parietal lobule based on receptor architectonics. Cerebral Cortex, 23(3), 615–628. doi:10.1093/cercor/bhs048
Castellanos, F. X., Lee, P., Sharp, W., Jeffries, N., Greenstein, D., Clasen, L., . . . Rapoport, J. L. (2002). Developmental trajectories of
brain volume abnormalities in children and adolescents with attention-deficit/hyperactivity disorder. American Medical Association Journal,
288(14), 1740–1748.
Castelli, F., Glaser, D. E., & Butterworth, B. (2006). Discrete and analogue quantity processing in the parietal lobe: A functional MRI
study. Proceedings of the National Academy of Sciences, USA, 103(12), 4693–4698.
Cattaneo, Z., Vecchi, T., Pascual-Leone, A., & Silvanto, J. (2009). Contrasting early visual cortex activation states causally involved in
visual imagery and short-term memory. European Journal of Neuroscience, 30(7), 1393–1400.
Cavanna, A. E. (2006). The precuneus: A review of its functional anatomy and behavioural correlates. Brain, 129(3), 564-583.
doi:10.1093/brain/awl004
Cavanna, A. E., & Trimble, M. R. (2006). The precuneus: A review of its functional anatomy and behavioural correlates. Brain, 129, 564–
583.
Ceci, S., & Bruch, M. (1993). Suggestibility of the child witness: A historical review and synthesis. Psychological Bulletin, 113(3), 403–439.
Chai, X. J., Ofen, N., Gabrieli, J. D., & Whitfield-Gabrieli, S. (2014). Development of deactivation of the default-mode network during
episodic memory formation. NeuroImage, 84, 932-938. doi:10.1016/j.neuroimage.2013.09.032
Chambers, R. A., Bremner, J. D., Moghaddam, B., Southwick, S. M., Charney, D. S., & Krystal, J. H. (1999). Glutamate and
posttraumatic stress disorder: Toward a psychobiology of dissociation. Seminars in Clinical Neuropsychiatry, 4(4), 274–281.
Chaminade, T., & Decety, J. (2002). Leader or follower? Involvement of the inferior parietal lobule in agency. NeuroReport, 13(15), 1975–
1978.
Chaminade, T., Meltzoff, A. N., & Decety, J. (2002). Does the end justify the means? A PET exploration of the mechanisms involved in
human imitation. NeuroImage, 15(2), 318-328. doi:10.1006/nimg.2001.0981
Champagne, D., Bagot, R., Hasselt, F., Ramakers, G., Meaney, M., Kloet, E., . . . Krugers, H. (2008). Maternal care and hippocampal
plasticity: Evidence for experience-dependent structural plasticity, altered synaptic function, and differential responsiveness to glucocorticoids
and stress. Journal of Neuroscience, 28(23), 6037–6045.
Champagne, F., Diorio, J., Sharma, S., & Meaney, M. J. (2001). Naturally occurring variations in maternal behavior in the rat are associated
with differences in estrogen-inducible central oxytocin receptors. Proceedings of the National Academy of Sciences, USA, 98(22), 12736–12741.
Champagne, F. A., Francis, D. D., Mar, A., & Meaney, M. J. (2003). Variations in maternal care in the rat as a mediating influence for the
effects of environment on development. Physiology and Behavior, 79(3), 359–371.
Champagne, F., Ian, C., Weaver, G., Diorio, J., Dymov, S., Szyf, M., & Meaney, M. J. (2006). Maternal care associated with methylation
of the estrogen receptor-a1b promoter and estrogen receptor-a expression in the medial preoptic area of female offspring. Endocrinology, 147(6),
2909–2915.
Chanen, A. M., Velakoulis, D., Carison, K., Gaunson, K., Wood, S. J., Yuen, H. P., . . . Pantelis, C. (2008). Orbitofrontal, amygdala and
hippocampal volumes in teenagers with first-presentation borderline personality disorder. Psychiatry Research: Neuroimaging, 163, 116–125.
Chapman, L. F., Walter, R. D., Markham, C. H., Rand, R. W., & Crandall, P. H. (1967). Memory changes induced by stimulation of
hippocampus or amygdala in epilepsy patients with implanted electrodes. Transactions of the American Neurological Association, 92, 50–56.
Charnov, E. L., & Berrigan, D. (1993). Why do female primates have such long lifespans and so few babies? Or life in the slow lane.
Evolutionary Anthropology, 1(6), 191–194.
Chaudhry, A. M., Parkinson, J. A., Hinton, E. C., Owen, A. M., & Roberts, A. C. (2009). Preference judgements involve a network of
structures within frontal, cingulate and insula cortices. European Journal of Neuroscience, 29(5), 1047–1055.
Chavez, C. M., McGaugh, J. L., & Weinberger, N. M. (2009). The basolateral amygdala modulates specific sensory memory
representations in the cerebral cortex. Neurobiology of Learning and Memory, 91(4), 382–392.
Chen, A. C., Oathes, D. J., Chang, C., Bradley, T., Zhou, Z., Williams, L. M., . . . Etkin, A. (2013). Causal interactions between fronto-
parietal central executive and default-mode networks in humans. Proceedings of the National Academy of Sciences, 110(49), 19944-19949.
doi:10.1073/pnas.1311772110
Cheng, D. T., Knight, D. C., Smith, C. N., & Helmstetter, F. J. (2006). Human amygdala activity during the expression of fear responses.
Behavioral Neuroscience, 120(5), 1187–1195.
Chiron, C., Jambaque, I., Nabbout, R., Lounes, R., Syrota, A., & Dulac, O. (1997). The right brain is dominant in human infants. Brain,
120(6), 1057–1065.
Chisholm, K. (1998). A three year follow-up of attachment and indiscriminate friendliness in children adopted from Romanian orphanages.

244
Child Development, 69(4), 1092–1106.
Chochon, F., Cohen, L., Demoortele, S., & Dehaene, S. (1999). Differential contributions of the left and right parietal lobules to number
processing. Journal of Cognitive Neuroscience, 11(6), 617–630.
Choi, J., & Bowles, S. (2007). The coevolution of parochial altruism and war. Science, 318, 636–640.
Choi, J., Jeong, B., Rohan, M. L., Polcari, A. M., & Teicher, M. H. (2009). Preliminary evidence for white matter tract abnormalities in
young adults exposed to parental verbal abuse. Biological Psychiatry, 65(3), 227–234. doi:10.1016/j.biopsych.2008.06.022
Christakou, A., Robbins, T. W., & Everitt, B. J. (2004). Prefrontal cortical-ventral striatal interactions involved in affective modulation of
attentional performance: Implications for corticostriatal circuit function. Journal of Neuroscience, 24(4), 773–780.
Christensen, A. J., Edwards, D. L., Wiebe, J. S., Benotsch, E. G., McKelvey, L., Andrews, M., & Lubaroff, D. M. (1996). Effect of verbal
self-disclosure on natural killer cell activity: Moderating influence of cynical hostility. Psychosomatic Medicine, 58(2), 150–155.
Christensen, H., Henderson, A. S., Griffiths, K., & Levings, C. (1997). Does ageing inevitably lead to declines in cognitive performance? A
longitudinal study of elite academics. Personality and Individual Differences, 23(1), 67–78.
Christman, S. D. (1994). The many sides of the two sides of the brain. Brain and Cognition, 26(1), 91–98.
Christodoulou, G. N., & Malliara-Loulakaki, S. (1981). Delusional misidentification syndromes and cerebral “dysrhythmia.” Psychiatrica
Clinica, 14(4), 245–251.
Chrobak, Q. M., & Zaragoza, M. S. (2013). When forced fabrications become truth: Causal explanations and false memory development.
Journal of Experimental Psychology: General, 142(3), 827-844. doi:10.1037/a0030093
Chugani, H. T. (1998). A critical period of brain development: Studies of cerebral glucose utilization with PET. Preventive Medicine, 27(2),
184–188.
Chugani, H. T., Behen, M. E., Muzik, O., Juhász, C., Nagy, F., & Chugani, D. C. (2001). Local brain functional activity following early
deprivation: A study of postinstitutionalized Romanian orphans. NeuroImage, 14(6), 1290–1301. doi:10.1006/nimg.2001.0917
Chugani, H. T., & Phelps, M. E. (1991). Imaging human brain development with positron emission tomography. Journal of Nuclear
Medicine, 32(1), 23–26.
Chugani, H. T., Phelps, M. E., & Mazziotta, J. C. (1987). Positron emission tomography study of human brain functional development.
Annals of Neurology, 22(4), 487–497.
Claeys, K. G., Lindsey, D. T., De Schutter, E., & Orban, G. A. (2003). A higher order motion region in human inferior parietal lobule:
Evidence from fMRI. Neuron, 40(3), 631–642.
Classen, J., Liepert, J., Wise, S. P., Hallett, M., & Cohen, L. G. (1998). Rapid plasticity of human cortical movements representation
induced by practice. Journal of Neurophysiology, 79(2), 1117–1123.
Clovis, C., Pollock, J., Goodman, R., Impey, S., Dunn, J., Mandel, G., . . . Nestler, E. J. (2005). Epigenetic mechanisms and gene networks
in the nervous system. Journal of Neuroscience, 25(45), 10379–10389.
Coan, J. A., Allen, J. B., & Harmon-Jones, E. (2001). Voluntary facial expression and hemispheric asymmetry over the frontal cortex.
Psychophysiology, 38(6), 912–925.
Coan, J. A., Schaefer, H. S., & Davidson, R. J. (2006). Lending a hand: Social regulation of the neural response to threat. Psychological
Science, 17(12), 1032–1039.
Cobb, S. (1944). Foundations of neuropsychiatry. Baltimore: Williams and Wilkins.
Coccaro, E. F., Siever, L. J., Klar, H. M., & Maurer, G. (1989). Serotonergic studies in patients with affective and personality disorders.
Archives of General Psychiatry, 46(7), 587–598.
Cogill, S. R., Caplan, H. L., Alexandra, H., Robson, K. M., & Kumar, R. (1986). Impact of maternal postnatal depression on cognitive
development of young children. British Medical Journal, 292(6529), 1165–1167.
Cohen, R. A., Grieve, S., Hoth, K. F., Paul, R. H., Sweet, L., Tate, D., . . . Williams, L. M. (2006). Early life stress and morphometry of
the adult anterior cingulate cortex and caudate nuclei. Biological Psychiatry, 59(10), 975–982.
Colby, C. L. (1998). Action-oriented spatial reference frames in cortex. Neuron, 20(1), 15–24.
Colby, C. L., & Goldberg, M. E. (1999). Space and attention in parietal cortex. Annual Review of Neuroscience, 22, 319–349.
Collette, J., Millam, R., Klasing, K., & Wakenell, P. (2000). Neonatal handling of Amazon parrots alters the stress response and immune
function. Applied Animal Behavior Science, 66(4), 335–349.
Colom, R., Haier, R. J., Head, K., Álvarez-Linera, J., Quiroga, M. Á., Shih, P. C., & Jung, R. E. (2009). Gray matter correlates of fluid,
crystallized, and spatial intelligence: Testing the P-FIT model. Intelligence, 37(2), 124–135.
Compton, D. M., Bachman, L. D., Brand, D., & Avet, T. L. (2000). Age-associated changes in cognitive function in highly educated
adults: Emerging myths and realities. International Journal of Geriatric Psychiatry, 15(1), 75–85.
Condren, R. M., & Thakore, J. H. (2001). Cushing’s disease and melancholia. Stress, 4(2), 91–119.
Conrad, C. D., Lupien, S. J., & McEwen, B. S. (1999). Support for a bimodal role for type II adrenal steroid receptors in spatial memory.

245
Neurobiology of Learning and Memory, 72(1), 39–46.
Coolidge, F. L., Segal, D. L., Stewart, S. E., & Ellet, J. A. C. (2000). Neuropsychological dysfunction in children with borderline
personality disorder features: A preliminary investigation. Journal of Research in Personality, 34, 554–561.
Coplan, J. D., & Lydiard, R. B. (1998). Brain circuits in panic disorder. Biological Psychiatry, 44(12), 1264–1276.
Corbetta, M., Patel, G., & Shulman, G. L. (2008). The reorienting system of the human brain: From environment to theory of mind.
Neuron, 58(3), 306–324.
Corbetta, M., & Shulman, G. (2002). Control of goal-directed and stimulus-driven attention in the brain. Nature Reviews Neuroscience,
3(3), 201–215.
Corcoran, K. A., & Quirk, G. J. (2007). Activity in prelimbic cortex is necessary for the expression of learned, but not innate, fears. Journal
of Neuroscience, 27(4), 840–844.
Coren, S., & Porac, C. (1977). Fifty centuries of right-handedness: The historical record. Science, 198(4317), 631–632.
Corina, D. P., Vaid, J., & Bellugi, U. (1992). The linguistic basis of left hemisphere specialization. Science, 255(5049), 1258–1260.
Cornette, L., Dupont, P., Salmon, E., & Orban, G. (2001). The neural substrate of orientation working memory. Journal of Cognitive
Neuroscience, 13(6), 813–828.
Corrigan, F. (2004). Psychotherapy as assisted homeostasis: Activation of emotional processing mediated by the anterior cingulate cortex.
Medical Hypothesis, 63(6), 968–973.
Costa, V. D., & Averbeck, B. B. (2013). Frontal-parietal and limbic-striatal activity underlies information sampling in the best choice
problem. Cerebral Cortex, 25(4), 972-982. doi:10.1093/cercor/bht286
Costanzo, E. Y., Villarreal, M., Drucaroff, L. J., Ortiz-Villafañe, M., Castro, M. N., Goldschmidt, M., . . . Guinjoan, S. M. (2015).
Hemispheric specialization in affective responses, cerebral dominance for language, and handedness. Behavioural Brain Research, 288, 11-19.
doi:10.1016/j.bbr.2015.04.006
Costella, C., Tejedor-Real, P., Mico, J. A., & Gibert-Rahola, A. (1995). Effects of neonatal handling on learned helplessness model of
depression. Physiology and Behavior, 57(2), 407–410.
Coull, J., Cotti, J., & Vidal, F. (2014). Increasing activity in left inferior parietal cortex and right prefrontal cortex with increasing temporal
predictability: An fMRI study of the hazard function. Procedia: Social and Behavioral Sciences, 126, 41–44. doi:10.1016/j.sbspro.2014.02.311
Coutinho, S., Plotsky, P., Sablad, M., Miller, J., Zhou, H., Bayati, A., . . . Mayer, E. A. (2002). Neonatal maternal separation alters stress-
induced responses to viscerosomatic nociceptive stimuli in rat. American Journal of Physiology Gastrointestinal and Liver Physiology, 282(2), G307–
G316.
Cowan, W. M., & Kandel, E. R. (2001). A brief history of synapses and synaptic transmission. In W. M. Cowan, T. C. Sudhof, & C. F.
Stevens (Eds.), Synapses (pp. 1–88). Baltimore: Johns Hopkins University Press.
Cowdry, R. W., Pickar, D., & Davies, R. (1985). Symptoms and EEG findings in the borderline syndrome. International Journal of
Psychiatry in Medicine, 15, 201–211.
Cozolino, L. J. (1997). The intrusion of early implicit memory into adult consciousness. Dissociation, 10(1), 44–53.
Cozolino, L. J. (2008). The healthy aging brain: Sustaining attachment, attaining wisdom. New York: Norton.
Cozolino, L. J. (2012). The neuroscience of human relationships: Attachment and the developing social brain (2nd ed.). New York: Norton.
Cozolino, L. J. (2010). The neuroscience of psychotherapy: Healing the social brain (2nd ed.). New York, NY: Norton.
Cozolino, L. J. (2014). Attachment-based teaching: Creating a tribal classroom. New York: Norton.
Cozolino, L. J. (2015). Why therapy works: Using our minds to change our brains. New York: Norton.
Crick, F. (1994). The astonishing hypothesis: The scientific search for the soul. New York: Charles Scribner’s Sons.
Cristinzio, C., N’diaye, K., Seeck, M., Vuilleumier, P., & Sander, D. (2010). Integration of gaze direction and facial expression in patients
with unilateral amygdala damage. Brain, 133(1), 248-261. doi:10.1093/brain/awp255
Critchley, H. (2005). Neural mechanism of autonomic, affective, and cognitive integration. Journal of Comparative Neurology, 493, 154–166.
Critchley, H., Daly, E., Phillips, M., Brammer, M., Bullmore, E., Williams, S., . . . Murphy, D. (2000). Explicit and implicit mechanisms
for processing of social information from facial expressions: A functional magnetic resonance imaging study. Human Brain Mapping, 9, 93–105.
Critchley, H. D., Melmed, R. N., Featherstone, E., Mathias, C. J., & Dolan, R. J. (2002). Volitional control of autonomic arousal: A
functional magnetic resonance study. NeuroImage, 16(4), 909–919.
Critchley, H. D., Wiens, S., Rotshtein, P., Öhman, A., & Dolan, R. J. (2004). Neural systems supporting interoceptive awareness. Nature
Neuroscience, 7, 189–195.
Crittenden, P. M., & DiLalla, D. L. (1988). Compulsive compliance: The development of an inhibitory coping strategy in infancy. Journal
of Abnormal Child Psychology, 16(5), 585–599.
Crowe, D., Goodwin, S., Blackman, R., Sakellaridi, S., Sponheim, S., Macdonald, A., & Chafee, M. (2013). Prefrontal neurons transmit
signals to parietal neurons that reflect executive control of cognition. Nature Neuroscience, 16(10), 1484–1491. doi:10.1038/nn.3509

246
Crowell, J. A., Treboux, D., & Waters, E. (2002). Stability of attachment representations: The transition to marriage. Developmental
Psychology, 38(4), 467–479.
Crowther, A., Smoski, M. J., Minkel, J., Moore, T., Gibbs, D., Petty, C., . . . Dichter, G. S. (2015). Resting-state connectivity predictors of
response to psychotherapy in major depressive disorder. Neuropsychopharmacology, 40(7), 1659–1673. doi:10.1038/npp.2015.12
Culham, J. C., & Kanwisher, N. G. (2001). Neuroimaging of cognitive functions in human parietal cortex. Current Opinion in Neurobiology,
11(2), 157–163.
Cullen, K. R., Vizueta, N., Thomas, K. M., Han, G. J., Lim, K. O., Camchong, J., . . . Schultz, S. C. (2011). Amygdala functional
connectivity in young women with borderline personality disorder. Brain Connectivity, 1(1), 61–71.
Cummings, J. L. (1993). Frontal-subcortical circuits and human behavior. Archives of Neurology, 50(8), 873–880.
Cummings, J. L., & Frankel, M. (1985). Gilles de la Tourette syndrome and the neurological basis of obsessions and compulsions. Biological
Psychiatry, 20(10), 117–126.
Cutting, J. (1992). The role of the right hemisphere in psychiatric disorders. British Journal of Psychiatry, 160, 583–588.
Czéh, B., Müller-Keuker, J. I. H., Rygula, R., Abumaria, N., Hiemke, C., Domenici, E., & Fuchs, E. (2007). Chronic social stress inhibits
cell proliferation in the adult medial prefrontal cortex: Hemispheric asymmetry and reversal by fluoxetine treatment. Neuropsychopharmacology,
32(7), 1490–1503.
D’Argembeau, A., & Salmon, E. (2012). The neural basis of semantic and episodic forms of self-knowledge: Insights from functional
neuroimaging. Advances in Experimental Medicine and Biology Sensing in Nature, 276-290. doi:10.1007/978-1-4614-1704-0_18
Dalla, C., Bangasser, D. A., Edgecomb, C., & Shors, T. J. (2007). Neurogenesis and learning: Acquiring and asymptotic performance
predict how many cells survive in the hippocampus. Neurobiology of Learning and Memory, 88, 143–148.
Damasio, A. (2010). Self comes to mind: Constructing the conscious brain. New York: Pantheon.
Damasio, A. R. (1994). Descartes’ error: Emotion, reason and the human brain. New York: Putnam and Sons.
Damasio, A. R., Grabowski, T. J., Bechara, A. Damasio, H., Ponto, L. L. B., Parvizi, J., & Hichwa, R. D. (2000). Subcortical and cortical
brain activity during the feeling of self-generated emotions. Nature Neuroscience, 3(10), 1049–1056.
Daniels, J. (2011). Default mode alterations in posttraumatic stress disorder related to early-life trauma: A developmental perspective.
Journal of Psychiatry & Neuroscience J Psychiatry Neurosci, 36(1), 56-59. doi:10.1503/jpn.100050
Dannlowski, U., Stuhrmann, A., Beutelmann, V., Zwanzger, P., Lenzen, T., Grotegerd, D., . . . Kugel, H. (2012). Limbic scars: Long-
term consequences of childhood maltreatment revealed by functional and structural magnetic resonance imaging. Biological Psychiatry, 71(4),
286–293. doi:10.1016/j.biopsych.2011.10.021
Daskalakis, Z. J., Christensen, B. K., Fitzgerald, P. B., & Chen, R. (2002). Transcranial magnetic stimulation: A new investigational and
treatment tool in psychiatry. Journal of Neuropsychiatry Clinical Neuroscience, 14(4), 406–415.
Davidson, R. J. (1999). The neurobiology of personality and personality disorders. In D. S. Charney, E. J. Nestler, & B. S. Bunney (Eds.),
Neurobiology of mental illness (pp. 841–854). New York: Oxford University Press.
Davidson, R. J. (2000). Affective style, psychopathology, and resilience: Brain mechanisms and plasticity. American Psychologist, 55(11),
1196–1214.
Davidson, R. J. (2002). Anxiety and affective style: Role of prefrontal cortex and amygdala. Biological Psychiatry, 51(1), 68–80.
Davidson, R. J. (2004). Well-being and affective style: Neural substrates and biobehavioural correlates. Philosophical Transactions of the Royal
Society: Biological Sciences, 359(1449), 1395–1411.
Davidson, R. J., Ekman, P., Saron, C. D., Senulis, J. A., & Friesen, W. V. (1990). Approach-withdrawal and cerebral asymmetry:
Emotional expression and brain physiology I. Journal of Personality and Social Psychology, 58(2), 330–341.
Davidson, R. J., & Fox, N. A. (1982). Asymmetrical brain activity discriminates between positive and negative affective stimuli in human
infants. Science, 218(4578), 1235–1237.
Davidson, R. J., & Fox, N. A. (1989). Frontal brain asymmetry predicts infants’ response to maternal separation. Journal of Abnormal
Psychology, 98(2), 127–131.
Davidson, R. J., Irwin, W., Anderle, M. J., & Kalin, N. H. (2003). The neural substrates of affective processing in depressed patients treated
with venlafaxine. American Journal of Psychiatry, 160(1), 64–75.
Davidson, R. J., Jackson, D. C., & Kalin, N. H. (2000). Emotion, plasticity, context, regulation: Perspectives from affective neuroscience.
Psychological Bulletin, 126, 890–909.
Davidson, R. J., Kabat-Zinn, J., Schumacher, M. S., Rosenkranz, M., Muller, D., Santorelli, S. F., . . . Sheridan, J. F. (2003). Alterations in
brain and immune function produced by mindfulness meditation. Psychosomatic Medicine, 65(4), 564–570.
Davis, M. (1992). The role of the amygdala in fear and anxiety. Annual Review of Neuroscience, 15, 353–375.
Davis, M. (1997). Neurobiology of fear responses: The role of the amygdala. Journal of Neuropsychiatry and Clinical Neurosciences, 9(3), 382–
402.

247
Davis, M. (1998). Are different parts of the extended amygdala involved in fear versus anxiety? Biological Psychiatry, 44(12), 1239–1247.
Davis, M., Myers, K. M., Chhatwal, J., & Ressler, K. J. (2006). Pharmacological treatments that facilitate extinction of fear: Relevance to
psychotherapy. Journal of the American Society for Experimental NeuroTherapeutics, 3(1), 82–96.
De Bellis, M. D., Baum, A. S., Birmaher, B., Keshavan, M. S., Eccard, C. H., Boring, A. M., . . . Ryan, N. D. (1999). Developmental
traumatology part I: Biological stress systems. Biological Psychiatry, 45(10), 1259–1270.
De Bellis, M. D., Keshavan, M. S., Clark, D. B., Casey, B. J., Giedd, J. N., Boring, A. M., . . . Ryan, N. D. (1999). Developmental
traumatology part II: Brain development. Biological Psychiatry, 45(10), 1271–1284.
De Brito, S. A., Mechelli, A., Wilke, M., Laurens, K. R., Jones, A. P., Barker, G. J., . . . Viding, E. (2008). Size matters: Increased grey
matter in boys with conduct problems and callous-unemotional traits. Brain, 132(4), 843-852. doi:10.1093/brain/awp011
De Brito, S. A., Viding, E., Sebastian, C. L., Kelly, P. A., Mechelli, A., Maris, H., & McCrory, E. J. (2013). Reduced orbitofrontal and
temporal grey matter in a community sample of maltreated children. Journal of Child Psychology and Psychiatry 54(1), 105–112.
doi:10.1111/j.1469-7610.2012.02597.x
de Casper, A. J., & Fifer, W. P. (1980). Of human bonding: Newborns prefer their mother’s voices. Science, 208(4448), 1174–1176.
De Pisapia, N., Serra, M., Rigo, P., Jager, J., Papinutto, N., Esposito, G., . . . Bornstein, M. H. (2014). Interpersonal competence in young
adulthood and right laterality in white matter. Journal of Cognitive Neuroscience, 26(6), 1257-1265. doi:10.1162/jocn_a_00534
Decety, J. (1994). Mapping motor representations with positron emission tomography. Nature, 371(6498), 600–602.
Decety, J., Chaminade, T., Grèzes, J., & Meltzoff, A. N. (2002). A PET exploration of the neural mechanisms involved in reciprocal
imitation. NeuroImage, 15(1), 265–272.
Decety, J., Echols, S., & Correll, J. (2009). The blame game: The effect of responsibility and social stigma on empathy for pain. Journal of
Cognitive Neuroscience, 22(5), 985–997.
Decety, J., & Jackson, P. L. (2004). The functional architecture of human empathy. Behavioral and Cognitive Neuroscience Reviews, 3(2), 71–
100.
Decety, J., & Lamm, C. (2006). Human empathy through the lens of social neuroscience. Scientific World Journal, 6, 1146–1163.
Decety, J., & Meyer, M. (2008). From emotional resonance to empathic understanding: A social developmental neuroscience account.
Development and Psychopathology, 20, 1053–1080.
Decety, J., Michalska, K. J., & Akitsuki, Y. (2008). Who caused the pain? An fMRI investigation of empathy and intentionality in children.
Neuropsychologia, 46(11), 2607-2614. doi:10.1016/j.neuropsychologia.2008.05.026
De Dreu, C. K. W., Greer, L. L., Handgraaf, M. J. J., Shalvi, S., Van Kleef, G. A., Matthijs Baas, . . . Feith, S. W. W. (2010). The
neuropeptide oxytocin regulates parochial altruism in intergroup conflict among humans. Science, 328, 1408–1411. doi:
10.1126/science.1189047
Dehaene, S., Molko, N., Cohen, L., & Wilson, A. J. (2004). Arithmetic and the brain. Current Opinion in Neurobiology, 14(2), 218–224.
Dehaene, S., Piazza, M., Pinel, P., & Cohen, L. (2003). Three parietal circuits for number processing. Cognitive Neuropsychology, 20(3),
487–506.
de la Fuente, J. M., Goldman, S., Stanus, E., Vizuete, C., Morlan, I., Bobes, J., & Mendlewicz, J. (1997). Brain glucose metabolism in
borderline personality disorder. Journal of Psychiatric Research, 31, 531–541.
de Lanerolle, N. C., Kim, J. H., Robbins, R. J., & Spencer, D. D. (1989). Hippocampal interneuron loss and plasticity in human temporal
lobe epilepsy. Brain Research, 495(2), 387–395.
Dennett, D. C. (1991). Consciousness explained. Boston: Little, Brown.
Derryberry, D., & Reed, M. A. (2002). Anxiety-related attentional biases and their regulation by attentional control. Journal of Abnormal
Psychology, 111(2), 225–236.
DeRubeis, R. J., Hollon, S. D., Amsterdam, J. D., Shelton, R. C., Young, P. R., Salomon, R. M., . . . Gallop, R. (2005). Cognitive therapy
vs. medications in the treatment of moderate to severe depression. Archives of General Psychiatry, 62(4), 409–416.
Desimone, R. (1991). Face-selective cells in the temporal cortex of monkeys. Journal of Cognitive Neuroscience, 3(1), 1–8.
de Vignemont, F., & Singer, T. (2006). The empathic brain: How, when and why? TRENDS in Cognitive Sciences, 10(10), 435–441.
doi:10.1016/j.tics.2006.08.008
Devinsky, O. (2000). Right cerebral hemisphere dominance for a sense of corporeal and emotional self. Epilepsy and Behavior, 1(1), 60–73.
Devinsky, O., Morrell, M. J., & Vogt, B. A. (1995). Contributions of anterior cingulate cortex to behavior. Brain, 118, 279–306.
De Waal, F. (1989). Peacemaking among primates. New York: Penguin.
de Waal., F. B. M. (2008). Putting the altruism back in altruism: The evolution of empathy. Annual Review of Psychology, 59, 279–300.
Diamond, D. M., Bennett, M. C., Fleshner, M., & Rose, G. M. (1992). Inverted-U relationships between the level of peripheral
corticosterone and the magnitude of hippocampal primed burst potentiation. Hippocampus, 2(4), 421–430.
Diamond, M. C., Krech, D., & Rosenweig, M. R. (1964). The effects of enriched environment on the histology of the rat cerebral cortex.

248
Journal of Comparative Neurology, 123, 111–119.
Diamond, M. C., Law, F., Rhodes, H., Lindner, B., Rosenweig, M. R., Krech, D., & Bennett, E. L. (1966). Increases of cortical depth and
glia numbers in rats subjected to enriched environments. Journal of Comparative Neurology, 128, 117–126.
Diamond, M. C., Scheibel, A. B., Murphy, G. M., & Harvey, T. (1985). On the brain of a scientist: Albert Einstein. Experimental
Neurology, 88(1), 198–204.
Dias, R., Robbins, T. W., & Roberts, A. C. (1996). Dissociation in prefrontal cortex of affective and attentional shifts. Nature, 380(6569),
69–72.
Díaz-Marsá, M., Carrasco, J. L., López-Ibor, M., Moratti, S., Montes, A., Ortiz, T., & López-Ibor, J. J. (2011). Orbitofrontal dysfunction
related to depressive symptomatology in subjects with borderline personality disorder. Journal of Affective Disorders, 134, 410–415.
Di Martino, A., Ross, K., Uddin, L. Q., Sklar, A. B., Castellanos, F. X., & Milham, M. P. (2009b). Functional brain correlates of social
and nonsocial processes in autism spectrum disorders: An activation likelihood estimation meta-analysis. Biological Psychiatry, 65(1), 63-74.
doi:10.1016/j.biopsych.2008.09.022
Di Martino, A., Shehzad, Z., Kelly, C., Roy, A. K., Gee, D. G., Uddin, L. Q., . . . Milham, M. P. (2009a). Relationship between cingulo-
insular functional connectivity and autistic traits in neurotypical adults. American Journal of Psychiatry, 166(8), 891-899.
doi:10.1176/appi.ajp.2009.08121894
Dimond, S. J., & Farrington, L. (1977). Emotional response to films shown to the right or left hemisphere of the brain measured by heart
rate. Acta Psychologica, 41(4), 255–260.
Dinn, W. M., Harris, C. L., Aycicegi, A., Greene, P. B., Kirkley, S. M., & Reilly, C. (2004). Neurocognitive function in borderline
personality disorder. Progress in Neuropsychopharmacology and Biological Psychiatry, 28, 329–341.
Diorio, J., & Meaney, M. (2007). Maternal programming of defensive responses through sustained effects on gene expression. Journal of
Psychiatry and Neuroscience, 32(4), 275–285.
Distel, M. A., Middeldorp, C. M., Trull, T. J., Derom, C. A., Willemsen, G., & Boomsma, D. I. (2011). Life events and borderline
personality features: The influence of gene-environment interaction and gene-environment correlation. Psychological Medicine, 41, 849–860.
Dixon-Gordon, K. L., Chapman, A. L., Lovasz, N., & Walters, K. (2011). Too upset to think: The interplay of borderline personality
features, negative emotions, and social problem solving in the laboratory. Personality Disorders: Theory, Research, and Treatment, 2(4), 243–260.
Dolan, R. (2007). Keynote address: Revaluing the orbital prefrontal cortex. Annals of the New York Academy of Sciences, 1121, 1–9.
Dolan, R. J. (1999). On the neurology of morals. Nature Neuroscience, 2(11), 927–929.
Dolcos, F., & McCarthy, G. (2006). Brain systems mediating cognitive interference by emotional distraction. Journal of Neuroscience, 26(7),
2072–2079.
Doll, A., Sorg, C., Manoliu, A., Wöller, A., Meng, C., Förstl, H., . . . Riedl, V. (2013). Shifted intrinsic connectivity of central executive
and salience network in borderline personality disorder. Frontiers in Human Neuroscience, 7(727), 1–13. doi:10.3389/fnhum.2013.00727
Domes, G., Rothfischer, J., Reichwald, U., & Hautzinger, M. (2005). Inverted-U function between salivary cortisol and retrieval of verbal
memory after hydrocortisone treatment. Behavioral Neuroscience, 119(2), 512–517.
Donegan, N. H., Sanislow, C. A., Blumberg, H. P., Fulbright, R. K., Lacadie, C., Skudlarski, P., . . . Wexler, B. E. (2003). Amygdala
hyperreactivity in borderline personality disorder: Implications for emotional dysregulation. Biological Psychiatry, 54, 1284–1293.
Dosenbach, N. U. F., Fair, D. A., Miezin, F. M., Cohen, A. L., Wenger, K. K., Dosenbach, R. A. T., . . . Petersen, S. E. (2007). Distinct
brain networks for adaptive and stable task control in humans. PNAS, 104(26), 11073–11078.
Dougherty, D. D., Rauch, S. L., Deckerbach, T., Marci, C., Loh, R., Shin, L. M., . . . Fava, M. (2004). Ventromedial prefrontal cortex and
amygdala dysfunction during an anger induction positron emission tomography study in patients with major depressive disorder with anger
attacks. Archives of General Psychology, 61(8), 795–804.
Dougherty, R. F., Ben-Shachar, M., Deutsch, G. K., Hernandez, A., Fox, G. R., & Wandell, B. A. (2007). Temporal-callosal pathway
diffusivity predicts phonological skills in children. Proceedings of the National Academy of Sciences, USA, 104(20), 8556–8561.
Douglas, R. J. (1967). The hippocampus and behavior. Psychological Bulletin, 67(6), 416–442.
Douglas, R. J., & Pribram, K. H. (1966). Learning and limbic lesions. Neuropsychologia, 4(3), 197–220.
Drake, R. A. (1984). Lateral asymmetry of personal optimism. Journal of Research in Personality, 18(4), 497–507.
Drake, R. A., & Seligman, M. E. P. (1989). Self-serving biases in causal attributions as a function of altered activation asymmetry.
International Journal of Neuroscience, 45(3–4), 199–204.
Dranovsky, A., & Hen, R. (2006). Hippocampal neurogenesis: Regulation by stress and antidepressants. Biological Psychiatry, 59(12), 1136–
1143.
Drevets, W. C. (1998). Functional neuroimaging studies of depression: The anatomy of melancholia. Annual Review of Medicine, 49, 341–
361.
Drevets, W. C., & Raichle, M. E. (1998). Reciprocal suppression of regional cerebral blood during emotional versus higher cognitive

249
processes: Implications for interactions between emotion and cognition. Cognition and Emotion, 12(3), 353–385.
Driessen, M., Herrmann, J., Stahl, K., Zwaan, M., Meier, S., Hill, A., . . . Peterson, D. (2000). Magnetic resonance imaging volumes of the
hippocampus and the amygdala in women with borderline personality disorder and early traumatization. Archives of General Psychiatry, 57,
1115–1122.
Driver, J., & Mattingley, J. B. (1998). Parietal neglect and visual awareness. Nature Neuroscience, 1(1), 17–22.
Dudai, Y. (2006). Reconsolidation: The advantage of being refocused. Current Opinion in Neurobiology, 16(2), 174–178.
Dunbar, R. I. (1996). Grooming, gossip, and the evolution of language. Cambridge, MA: Harvard University Press.
Dunbar, R. I. (2014). The social brain: Psychological underpinnings and implications for the structure of organizations. Current Directions in
Psychological Science, 23(2), 109–114. doi: 10.1177/0963721413517118
Durston, S., Tottenham, N. T., Thomas, K. M., Davidson, M. C., Eigsti, I., Yang, Y., Ulug, A. M., & Casey, B. J. (2003). Differential
patterns of striatal activation in young children with and without ADHD. Biological Psychiatry, 53(10), 871–878.
Dwivedi, Y., Rizavi, H. S., Conley, R. R., Roberts, R. C., Tamminga, C. A., & Pandey, G. N. (2003). Altered gene expression of brain-
derived neurotrophic factor and receptor tyrosine kinase B in postmortem brain of suicide subjects. Archives of General Psychiatry, 60(8), 804–
815.
Dziobek, I., Preissler, S., Grozdanovic, Z., Heuser, I., Heekeren, H. R., & Roepke, S. (2011). Neuronal correlates of altered empathy and
social cognition in borderline personality disorder. NeuroImage, 57, 539–548.
Eales, L. A. (1985). Song learning in zebra finches: Some effects of song model availability on what is learnt and when. Animal Behavior,
33(4), 1293–1300.
Edelman, G. M. (1987). Neural Darwinism: The theory of neuronal group selection. New York: Basic Books.
Edelman, G. M. (1989). The remembered present: A biological theory of consciousness. New York: Basic Books.
Edin, F., Macoveanu, J., Olesen, P., Tegner, J., & Klingberg, T. (2007). Stronger synaptic connectivity as a mechanism behind development
of working memory-related brain activity during childhood. Journal of Cognitive Neuroscience, 19(5), 750–760.
Edmiston, E. E., Wang, F., Mazure, C. M., Guiney, J., Sinha, R., Mayes, L. C., & Blumberg, H. P. (2011). Corticostriatal-limbic gray
matter morphology in adolescents with self-reported exposure to childhood maltreatment. Archives of Pediatrics and Adolescent Medicine, 165(12),
1069–1077. doi:10.1001/archpediatrics.2011.565
Egeland, B., & Farber, E. A. (1984). Infant-mother attachment: Factors related to its development and changes over time. Child
Development, 55(3), 753–771.
Eichenbaum, H. (1992). The hippocampal system and declarative memory in animals. Journal of Cognitive Neuroscience, 4(3), 217–231.
Eisenberg, L. (1995). The social construction of the human brain. American Journal of Psychiatry, 152(11), 1563–1575.
Eisch, A. J., & Petrik, D. (2012). Depression and hippocampal neurogenesis: A road to remission? Science, 338(6103), 72-75.
doi:10.1126/science.1222941
Ekman, P., & Davidson, R. J. (1993). Voluntary smiling changes regional brain activity. Psychological Science, 4(5), 342–345.
Elbert, T., Flor, H., Birbaumer, N., Knecht, S., Hampson, S., Larbig, W., & Taub, E. (1994). Extensive reorganization of the
somatosensory cortex in adult humans after nervous system injury. NeuroReport, 5(18), 2593–2597.
Elbert, T., Pantev, C., Wienbruch, C., Rockstroh, B., & Taub, E. (1995). Increased cortical representation of the fingers of the left hand in
string players. Science, 270(5234), 305–307.
Eliot, L. (1999). What’s going on in there? How the brain and mind develop in the first five years of life. New York: Bantam.
Ellenberger, H. F. (1970). The discovery of the unconscious. New York: Basic Books.
Elliott, R., Agnew, Z., & Deakin, J. (2008). Medial orbitofrontal cortex codes relative rather than absolute value of financial rewards in
humans. European Journal of Neuroscience, 27(9), 2213–2218.
Elliott, R., Friston, K. J., & Dolan, R. J. (2000). Dissociable neural responses in human reward systems. Journal of Neuroscience, 20(16),
6159–6165.
Ellis, A. (1962). Reason and emotion in psychotherapy. Secaucus, NJ: Lyle Stuart.
Elton, A., & Gao, W. (2014). Divergent task-dependent functional connectivity of executive control and salience networks. Cortex, 51, 56–
66. doi:10.1016/j.cortex.2013.10.012
Elvander-Tottie, E., Eriksson, T. M., Sandin, J., & Ögren, S. O. (2006). N-methyl-d-aspartate receptors in the medial septal area have a
role in spatial and emotional learning in the rat. Neuroscience, 142(4), 963–978.
Encinas, J. M., Vaahtokari, A., & Enikolopov, G. (2006). Fluoxetine targets early progenitor cells in the adult brain. Proceedings of the
National Academy of Sciences, USA, 103(21), 8233–8238.
Engdahl, B., Leuthold, A. C., Tan, H. M., Lewis, S. M., Winskowski, A. M., Dikel, T. N., & Georgopoulos, A. P. (2010). Post-traumatic
stress disorder: A right temporal lobe syndrome? Journal of Neural Engineering, 7(6), 1-8. doi:10.1088/1741-2560/7/6/066005
Engle, P. L., & Black, M. M. (2008). The effect of poverty on child development and educational outcomes. Annals of the New York Academy

250
of Sciences, 1136(1), 243–256. doi:10.1196/annals.1425.023
Enlow, M., Egeland, B., Carlson, E., Blood, E., & Wright, R. (2013). Mother-infant attachment and the intergenerational transmission of
posttraumatic stress disorder. Development and Psychopathology, 26(1), 41–65. doi:10.1017/S0954579413000515
Eriksson, P. S., Perfileva, E., Bjork-Eriksson, T., Alborn, A. M., Nordborg, C., Peterson, D. A., . . . Gage, F. (1998). Neurogenesis in the
adult human hippocampus. Nature Medicine, 4, 1313–1317.
Ernst, A., Alkass, K., Bernard, S., Salehpour, M., Perl, S., Tisdale, J., . . . Frisén, J. (2014). Neurogenesis in the striatum of the adult human
brain. Cell, 156(5), 1072–1083. doi:10.1016/j.cell.2014.01.044
Esch, T., & Stefano, G. B. (2005). The neurobiology of love. Neuroendocrinology Letters, 26(3), 175–192.
Eskine, K., Kacinik, N., & Prinz, J. (2011). A bad taste in the mouth: Gustatory disgust influences moral judgment. Psychological Science,
22(3), 295–299. doi:10.1177/0956797611398497
Eslinger, P. J. (1998). Neurological and neuropsychological bases of empathy. European Neurology, 39(4), 193–199.
Etchison, M., & Kleist, D. (2000). Review of narrative therapy: Research and utility. Family Journal, 8(1), 61–66.
Etkin, A., Egner, T., & Kalisch, R. (2011). Emotional processing in anterior cingulate and medial prefrontal cortex. Trends in Cognitive
Sciences, 15(2), 85–93. doi:10.1016/j.tics.2010.11.004
Etkin, A., Phil, M., Pittenger, C., Polan, H. J., & Kandel, E. R. (2005). Toward a neurobiology of psychotherapy: Basic science and clinical
applications. Journal of Neuropsychiatry Clinical Neuroscience, 17(2), 145–158.
Fair, D. A., Cohen, A. L., Dosenbach, N. U., Church, J. A., Miezin, F. M., Barch, D. M., . . . Schlaggar, B. L. (2008). The maturing
architecture of the brain’s default network. Proceedings of the National Academy of Sciences, 105(10), 4028-4032. doi:10.1073/pnas.0800376105
Falkai, P., & Bogerts, B. (1986). Cell loss in the hippocampus of schizophrenics. European Archives of Psychiatry and Neurological Sciences,
236(3), 154–161.
Fan, J., Gu, X., Liu, X., Guise, K. G., Park, Y., Martin, L., . . . Hof, P. R. (2011). Involvement of the anterior cingulate and frontoinsular
cortices in rapid processing of salient facial emotional information. NeuroImage, 54, 2539–2546.
Fanselow, M. S. (1986). Conditioned fear-induced opiate analgesia: A competing motivational state theory of stress analgesia. Annals of the
New York Academy of Sciences, 467, 40–54.
Fantuzzo, J. W., Rouse, H. L., McDermott, P. A., Sekino, Y., Childs, S., & Weiss, A. (2005). Early childhood experiences and
kindergarten success: A population-based study of large urban setting. Social Psychology Review, 34(4), 571–588.
Farrer, C., & Frith, C. D. (2002). Experiencing oneself vs. another person as being the cause of an action: The neural correlates of the
experience of agency. Neuroimage, 15, 596–603.
Federici, R. (1998). Help for the hopeless child: A guide for families. Alexandria, VA: Federici and Associates.
Federspiel, A., Volpe, U., Horn, H., Dierks, T., Franck, A., Vannini, P., . . . Maj, M. (2005). Motion standstill leads to activation of
inferior parietal lobe. Human Brain Mapping, 27(4), 340–349.
Feinberg, T. E., & Shapiro, R. M. (1989). Misidentification-reduplication and the right hemisphere. Neuropsychiatry, Neuropsychology, and
Behavioral Neurology, 2(1), 39–48.
Feldman, R. (2012). Oxytocin and social affiliation in humans. Hormones and Behavior, 61(3), 380–391.
Feldman, R., Greenbaum, C. W., & Yirimiya, N. (1999). Mother-infant affect synchrony as an antecedent of the emergence of self-control.
Developmental Psychology, 35(1), 223–231.
Felitti, V. J., Anda, R. F., Nordenberg, D., Williamson, D. F., Spitz, A. M., Edwards, V., . . . Marks, J. S. (1998). Relationship of
childhood abuse and household dysfunction to many of the leading causes of death in adults. American Journal of Preventive Medicine, 14(4),
245–258. doi:10.1016/s0749-3797(98)00017-8
Fellin, T., Pascual, O., & Haydon, P. G. (2006). Astrocytes coordinate synaptic networks: Balanced excitation and inhibition. Physiology,
21(3), 208–215.
Fernandes, C. C., Pinto-Duarte, A., Ribeiro, J. A., & Sebastião, A. M. (2008). Postsynaptic action of brain-derived neurotrophic factor
attenuates a7 nicotinic acetylcholine receptor-mediated responses in hippocampal interneurons. Journal of Neuroscience, 28(21), 5611–5618.
Ferrari, P. F., Gallese, V., Rizzolatti, G., & Fogassi, L. (2003). Mirror neurons responding to the observation of ingestive and
communicative mouth actions in the monkey ventral premotor cortex. European Journal of Neuroscience, 17(8), 1703-1714. doi:10.1046/j.1460-
9568.2003.02601.x
Fias, W., Lamartine, J., Caessens, B., & Orban, G. (2007). Processing of abstract knowledge in the horizontal segment of the intraparietal
sulcus. Journal of Neuroscience, 27(33), 8952–8957.
Fias, W., Lammertyn, J., Reynvoet, B., Dupont, P., & Orban, G. (2003). Parietal representation of symbolic and nonsymbolic magnitude.
Journal of Cognitive Neuroscience, 15(1), 47–56.
Field, T. (1997). The treatment of depressed mothers and their infants. In L. Murry & P. J. Cooper (Eds.), Postpartum depression and child
development (pp. 221–236). New York: Guilford.

251
Field, T., & Diego, M. (2008a). Cortisol: The culprit prenatal stress variable. International Journal of Neuroscience, 118(8), 1181–1205.
Field, T., & Diego, M. (2008b). Maternal depression effects on infant frontal EEG asymmetry. International Journal of Neuroscience, 118(8),
1081–1108.
Field, T., Diego, M., & Hernandez-Reif, M. (2006). Prenatal depression effects on the fetus and newborn: A review. Infant Behavior and
Development, 29(3), 445–455.
Field, T. M., Gizzle, N., Scafidi, F., Abrams, S., Richardson, S., Kuhn, C., & Schanberg, S. (1996). Massage therapy for infants of
depressed mothers. Infant Behavior and Development, 19(1), 107–112.
Field, T. M., Healy, B., Goldstein, S., & Guthertz, M. (1990). Behavior-state matching and synchrony in mother-infant interactions of
nondepressed versus depressed dyads. Developmental Psychology, 26(1), 7–14.
Field, T. M., Healy, B., Goldstein, S., Perry, S., & Bendell, D. (1988). Infants of depressed mothers show “depressed” behavior even with
nondepressed adults. Child Development, 59(6), 1569–1579.
Field, T. M., Woodson, R., Greenberg, R., & Cohen, D. (1982). Discrimination and imitation of facial expressions by neonates. Science,
218(4568), 179–181.
Figiel, G. S., Epstein, C., McDonald, W. M., Amazon-Leece, J., Figiel, L., Saldivia, A., & Glover, M. D. (1998). The use of rapid-rate
transcranial magnetic stimulation (rTMS) in refractory depression. Journal of Clinical Neuropsychiatry and Clinical Neurosciences, 10(1), 20–25.
Fine, M. L. (1989). Embryonic, larval and adult development of the sonic neuromuscular system in the oyster toadfish. Brain, Behavior and
Evolutions, 34(1), 13–24.
Fischer, K. W. (1987). Relations between brain and cognitive development. Child Development, 58(3), 623–632.
Fischer, K. W., Shaver, P. R., & Carnochan, P. (1990). How emotions develop and how they organise development. Cognition and Emotion,
4(2), 81–127.
Fiset, P., Paus, T., Daloze, T., Plourde, G., Meuret, P., Bonhomme, V., . . . Evans, A. C. (1999). Brain mechanisms of propofol-induced
loss of consciousness in humans: A positron emission tomographic study. The Journal of Neuroscience, 19(13), 5506-5513.
Fish, E. W., Shahrokh, D., Bagot, R., Caldji, C., Bredy, T., Szyf, M., & Meaney, M. J. (2004). Epigenetic programming of stress responses
through variations in maternal care. Annals of the New York Academy of Sciences, 1036, 167–180.
Fisher, H. E. (1998). Lust, attraction, and attachment in mammalian reproduction. Human Nature, 9(1), 23–52.
Fisher, H. E. (2004). Why we love: The nature and chemistry of romantic love. New York: Holt Paperbacks.
Fisher, P. M., Meltzer, C. C., Ziolko, S. K., Price, J. C., Moses-Kolko, E. L., Berga, S. L., & Harari, A. R. (2006). Capacity for 5-HT1A-
mediated autoregulation predicts amygdala reactivity. Nature Neuroscience, 9(11), 1362–1363.
Fish-Murry, C. C., Koby, E. V., & van der Kolk, B. A. (1987). Evolving ideas: The effects of abuse on children’s thought. In B. A. van der
Kolk (Ed.), Psychological trauma (pp. 89–110). Washington, DC: American Psychiatric Press.
Fitzgerald, P. B., Laird, A. R., Maller, J., & Daskalakis, Z. J. (2008). A meta-analytic study of changes in brain activation in depression.
Human Brain Mapping, 29(6), 683–695. doi:10.1002/hbm.20426
Fleming, A. S., & Korsmit, M. (1996). Plasticity in the maternal circuit: Effects of maternal experience on Fos-lir in hypothalamic, limbic,
and cortical structures in the postpartum rat. Behavioral Neuroscience, 110(3), 567–582.
Fonagy, P., Gergely, G., Jurist, E., & Target, M. (2002). Affect regulation, mentalization, and the development of self. New York: Other Press.
Fonagy, P., Steele, H., & Steele, M. (1991). Maternal representations of attachment during pregnancy predict the organization of infant-
mother attachment at one year of age. Child Development, 62(5), 891–905.
Fonagy, P., Steele, M., Steele, H., Moran, G. S., & Higgitt, A. C. (1991). The capacity for understanding mental states: The reflective self
in parent and child and its significance for security of attachment. Infant Mental Health Journal, 12(3), 201–218.
Fonagy, P., Target, M., & Gergely, G. (2000). Attachment and borderline personality disorder: A theory and some evidence. Borderline
Personality Disorder, 23, 103–122.
Forbes, E. E., Shaw, D. S., Silk, J. S., Feng, X., Cohn, J. F., Fox, N. A., & Kovacs, M. (2008). Children’s affect expression and frontal
EEG asymmetry: Transactional associations with mothers’ depressive symptoms. Journal of Abnormal Child Psychology, 36(2), 207–221.
Fowler, C. D., Liu, Y., Ouimet, C., & Wang, Z. (2002). The effects of social environment on adult neurogenesis in the female prairie vole.
Journal of Neurobiology, 51(2), 115–128.
Fox, M. D., Snyder, A. Z., Vincent, J. L., Corbetta, M., Van Essen, D. C., & Raichle, M. E. (2005). The human brain is intrinsically
organized into dynamic, anticorrelated functional networks. Proceedings of the National Academy of Sciences, USA, 102(27), 9673–9678.
Fox, N. A. (1991). If it’s not left it’s right: Electroencephalograph asymmetry and the development of emotion. American Psychologist, 46(8),
863–872.
Fox, N. A., & Davidson, R. J. (1988). Patterns of brain electrical activity during facial signs of emotion in 10-month-old infants.
Developmental Psychology, 24(2), 230–236.
Francis, D., Diorio, J., Plotsky, P., & Meaney, M. (2002). Environmental enrichment reverses the effects of maternal separation on stress

252
reactivity. Journal of Neuroscience, 22(18), 7840–7843.
Frank, J. (1963). Persuasion and healing. New York: Schocken.
Fransson, P. (2005). Spontaneous low-frequency BOLD signal fluctuations: An fMRI investigation of the resting-state default mode of
brain function hypothesis. Human Brain Mapping, 26(1), 15-29. doi:10.1002/hbm.20113
Freedman, L. J., Insel, T. R., & Smith, Y. (2000). Subcortical projections of area 25 (subgenual cortex) of the macaque monkey. Journal of
Comparative Neurology, 421(2), 172–188.
Freeman, T. W., & Kimbrell, T. (2001). A “cure” for chronic combat-related posttraumatic stress disorder secondary to a right frontal lobe
infarct: A case report. Journal of Neuropsychiatry and Clinical Neurosciences, 13(1), 106–109.
Freud, S. (1968). Project for a scientific psychology. In J. Strachey (Ed.), The standard edition of the complete psychological works of Sigmund
Freud (Vol. 1, pp. 3–182). London: Hogarth. (Original work published 1895)
Freud, S. (1975). The dynamics of transference. In J. Strachey (Ed.), The standard edition of the complete psychological works of Sigmund Freud
(Vol. 12, pp. 99–108). London: Hogarth. (Original work published 1912)
Freund, H. (2001). The parietal lobe as a sensorimotor interface: A perspective from clinical and neuroimaging data. NeuroImage, 14(1).
doi:10.1006/nimg.2001.0863
Frewen, P. A., Dozois, D. J., Neufeld, R. W., Lane, R. D., Densmore, M., Stevens, T. K., & Lanius, R. A. (2011). Emotional numbing in
posttraumatic stress disorder. Journal of Clinical Psychiatry, 73(4), 431–436. doi:10.4088/jcp.10m06477
Frey, S., & Petrides, M. (2000). Orbitofrontal cortex: A key prefrontal region for encoding information. Proceedings of the National Academy
of Sciences, USA, 97(15), 8723–8727.
Freyd, J. J. (1987). Dynamic mental representations. Psychological Reviews, 94(4), 427–438.
Friberg, L., Olsen, T. S., Roland, P. E., Paulsen, O. B., & Lassen, N. A. (1985). Focal increase of blood flow in the cerebral cortex of man
during vestibular stimulation. Brain, 108(Pt. 3), 609–623.
Fricchione, G., & Stefano, G. B. (2005). Placebo neural systems: Nitric oxide, morphine and the dopamine brain reward and motivation
circuitries. Medical Science Monitor, 11(5), MS54–65.
Frick, R. B. (1982). The ego and the vestibulocerebellar system: Some theoretical perspectives. Psychoanalytic Quarterly, 51(1), 93–122.
Friedman, D., Goldman, R., Stern, Y., & Brown, T. R. (2009). The brain’s orienting response: An event-related functional magnetic
resonance imaging investigation. Human Brain Mapping, 30(4), 1144–1154. doi:10.1002/hbm .20587
Frith, C. D., & Frith, U. (1999). Interacting minds: A biological basis. Science, 286, 1692– 1695.
Frith, U., & Frith, C. (2010). The social brain: Allowing humans to boldly go where no other species has been. Philosophical Transactions of
the Royal Society, 365, 165–176. doi: 10.1098/rstb.2009.0160
Frühbeis, C., Fröhlich, D., Kuo, W. P., & Krämer-Albers, E. (2013). Extracellular vesicles as mediators of neuron-glia communication.
Frontiers in Cellular Neuroscience, 7(182), 1–6. doi:10.3389/fncel.2013.00182
Fuente-Fernández, R., Ruth, T. J., Sossi, V., Schulzer, M., Calne, D. B., & Stoessl, A. J. (2001). Expectation and dopamine release:
Mechanism of the placebo effect in Parkinson’s disease. Science, 293(5532), 1164–1166.
Fujikawa, T., Soya, H., Fukuoka, H., Alam, K. S. M., Yoshizato, H., McEwan, B. S., & Nakashima, K. (2000). A biphasic regulation of
receptor mRNA expressions for growth hormone, glucocorticoid and mineralocorticoid in the rat dentate gyrus during acute stress. Brain
Research, 874(2), 186–193.
Furmark, T., Tillfors, M., Marteinsdottir, I., Fischer, H., Pissiota, A., Långström, B., & Fredrikson, M. (2002). Common changes in
cerebral blood flow in patients with social phobia treated with citalopram or cognitive-behavioral therapy. Archives of General Psychiatry, 59(5),
425–433.
Fuster, J. M. (1996). Frontal lobe and the cognitive foundation of behavioral action. In A. R. Damasio, H. Damasio, & Y. Christen (Eds.),
Neurobiology of decision-making (pp. 47–61). Berlin: Springer-Verlag.
Fuster, J. M. (1997). The prefrontal cortex: Anatomy, physiology, and neuropsychology of the frontal lobe (3rd ed.). Philadelphia: Lippincott-
Raven.
Fuster, J. M. (2004). Upper processing stages of the perception-action cycle. Trends in Cognitive Science, 8(4), 143–145.
Fuster, J. M., Bonder, M., & Kroger, J. K. (2000). Cross-modal and cross-temporal association in neurons of frontal cortex. Nature,
405(6784), 347–351.
Gablik, S. (1985). Magritte. New York: Thames and Hudson.
Gainotti, G. (1972). Emotional behavior and hemispheric side of the lesion. Cortex, 8(1), 41–55.
Gainotti, G. (2012). Unconscious processing of emotions and the right hemisphere. Neuropsychologia, 50(2), 205-218.
doi:10.1016/j.neuropsychologia.2011.12.005
Galin, D. (1974). Implications for psychiatry of left and right cerebral specialization: A neurophysiological context for unconscious
processes. Archives of General Psychiatry, 31(4), 572–583.

253
Galin, D., Johnstone, J., Nakell, L., & Herron, J. (1979). Development for the capacity for tactile information transfer between hemispheres
in normal children. Science, 204, 1330–1331.
Gallagher, M., McMahon, R. W., & Schoenbaum, G. (1999). Orbitofrontal cortex and representation of incentive value in associative
learning. Journal of Neuroscience, 19(15), 6610–6614.
Gallese, V., Fadiga, L., Fogassi, L., & Rizolatti, G., (1996). Action recognotion in the premotor cortex. Brain, 119(2), 593–609.
Gallese, V., & Goldman, A. (1998). Mirror neurons and the simulation theory of mind-reading. Trends in Cognitive Sciences, 2(12), 493-
501. doi:10.1016/s1364-6613(98)01262-5
Gallese, V., & Keysers, C. (2001). Mirror neurons: A sensorimotor representation system. Behavioral and Brain Sciences, 24(05), 983-984.
doi:10.1017/s0140525x01340116
Gallese, V., Fadiga, L., Fogassi, L., & Rizzolatti, G. (1996). Action recognition in the premotor cortex. Brain, 119(2), 593-609.
doi:10.1093/brain/119 .2.593
Gallese, V., Rochat, M., Cossu, G., & Sinigaglia, C. (2009). Motor cognition and its role in the phylogeny and ontogeny of action
understanding. Developmental Psychology, 45(1), 103–113.
Galovski, T., & Lyons, J. A. (2004). Psychological sequelae of combat violence: A review of the impact of PTSD on the veteran’s family and
possible interventions. Aggression and Violent Behavior, 9(5), 477–501. doi:10.1016/s1359-1789(03)00045-4
Galynker, I. I., Cai, J., Ongseng, F., Fineston, H., Dutta, E., & Serseni, D. (1998). Hypofrontality and negative symptoms in major
depressive disorder. Journal of Nuclear Medicine, 39(4), 608–612.
Ganis, G., Kosslyn, S. M., Stose, S., Thompson, W. L., & Yurgelun-Todd, D. A. (2003). Neural correlates of different types of deception:
An fMRI investigation. Cerebral Cortex, 13(8), 830–836.
Gao, W., Zhu, H., Giovanello, K. S., Smith, J. K., Shen, D., Gilmore, J. H., & Lin, W. (2009). Evidence on the emergence of the brain’s
default network from 2-week-old to 2-year-old healthy pediatric subjects. Proceedings of the National Academy of Sciences, 106(16), 6790-6795.
doi:10.1073/pnas.0811221106
Garavan, H., Ross, T. J., & Stein, E. A. (1999). Right hemisphere dominance of inhibitory control: An event-related functional MRI study.
Proceedings of the National Academy of Sciences, USA, 96, 8301–8306.
Garoflos, E., Stamatakis, A., Pondiki, S., Apostolou, A., Philippidis, H., & Sylianopoulou, F. (2007). Cellular mechanism underlying the
effect of a single exposure to neonatal handling on neurotrophin-3 in the brain of 1-day-old rats. Neuroscience, 148, 349–358.
Garoflos, E., Stamatakis, A., Rafrogianni, A., Pondiki, S., & Sylianopoulou, F. (2008). Neonatal handling on the first postnatal day leads to
increased maternal behavior and fos levels in the brain of the newborn rat. Developmental Psychobiology, 50(7), 704–713.
Garrity, A. (2007). Aberrant “default mode” functional connectivity in schizophrenia. American Journal of Psychiatry Am J Psychiatry, 164(3),
450. doi:10.1176/appi.ajp.164.3.450
Gartside, S. E., Leitch, M. M., McQuade, R., & Swarbrick, D. J. (2003). Flattening the glucocorticoid rhythm causes changes in
hippocampal expression of messenger RNAs coding structural and functional proteins: Implications for aging and depression.
Neuropsychopharmacology, 28(5), 821–829.
Gauthier, I., Tarr, M. J., Moylan, J., Skudlarski, P., Gore, J. C., & Anderson, A. W. (2000). The fusiform “face area” is part of a network
that processes faces at the individual level. Journal of Cognitive Neuroscience, 12(3), 495–504.
Gazzaley, A., Rissman, J., Cooney, J., Aaron, R., Seibert, T., Clapp, W., & D’Esposito, M. (2007). Functional interactions between
prefrontal and visual association cortex contribute to top-down modulation of visual processing. Cerebral Cortex, 17, i125–i135.
Gazzaniga, M. S. (1989). Organization of the human brain. Science, 245(4921), 947–952.
Gazzaniga, M. S. (1995). Consciousness and the cerebral hemispheres. In M. S. Gazzaniga (Ed.), The cognitive neurosciences (pp. 1391–
1400). Cambridge, MA: MIT Press.
Gazzaniga, M. S., LeDoux, J. E., & Wilson, D. H. (1977). Language, praxis, and the right hemisphere: Clues to some mechanisms of
consciousness. Neurology, 27(12), 1144–1147.
Ge, W., Miyawaki, A., Gage, F. H., Jan, Y. N., & Jan, L. Y. (2012). Local generation of glia is a major astrocyte source in postnatal cortex.
Nature, 484(7394), 376-380. doi:10.1038/nature10959
Geday, J., Kupers, R., & Gjedde, A. (2007). As time goes by: Temporal constraints on emotional activation of inferior medial prefrontal
cortex. Cerebral Cortex, 17(12), 2753–2759.
Gedo, J. E. (1991). The biology of clinical encounters: Psychoanalysis as a science of mind. Hillsdale, NJ: Analytic Press.
Gehring, W. J., & Willoughby, A. R. (2002). The medial frontal cortex and the rapid processing of monetary gains and losses. Science,
295(5563), 2279–2282.
Genovesio, A., Wise, S., & Passingham, R. (2014). Prefrontal-parietal function: From foraging to foresight. Trends in Cognitive Sciences,
18(2), 72–81. doi:10.1016/j.tics.2013.11.007
Gentili, C., Ricciardi, E., Gobbini, M. I., Santarelli, M. F., Haxby, J. V., Pietrini, P., & Guzzelli, M. (2009). Beyond amygdala: Default

254
mode network activity differs between patients with social phobia and healthy controls. Brain Research Bulletin, 79, 409–413.
George, M. D., Wasserman, M. D., Kimbrell, J. T., Little, M. D., Williams, W. E., Danielson, A. L., . . . Post, R. M. (1997). Mood
improvement following daily left prefrontal repetitive transcranial magnetic stimulation in patients with depression: A placebo-controlled
crossover trial. American Journal of Psychiatry, 154(12), 1752–1756.
Geschwind, N., & Galaburda, A. M. (1985). Cerebral lateralization: Biological mechanisms, associations and pathology: I. A hypothesis
and a program for research. Archives of Neurology, 42(5), 428–459.
Geuze, E., Vermetten, E., & Bremner, J. D. (2005). MR-based in vivo hippocampal volumetrics: 2. Findings in neuropsychiatric disorders.
Molecular Psychiatry, 10(2), 160–184.
Ghashghaei, H. T., & Barbas, H. (2002). Pathways for emotion: Interactions of prefrontal and anterior temporal pathways in the amygdala
of the rhesus monkey. Neuroscience, 115(4), 1261–1279.
Ghashghaei, H. T., Hilgetag, C. C., & Barbas, H. (2007). Sequence of information processing for emotions based on the anatomic dialogue
between prefrontal cortex and amygdala. NeuroImage, 34(3), 905–923.
Gibson, J. J. (1966). The senses considered as perceptual systems. Boston: Houghton Mifflin.
Giesen-Bloo, J., & Arntz, A. (2005). World assumptions and the role of trauma in borderline personality disorder. Journal of Behavior
Therapy and Experimental Psychiatry, 36(3), 197–208. doi:10.1016/j.jbtep.2005.05.003
Gilbertson, M. W., Shenon, M. E., Ciszewski, A., Kasai, K., Lasko, N. B., Orr, S. P., . . . Pitman, R. K. (2002). Smaller hippocampal
volume predicts pathologic vulnerability to psychological trauma. Nature Neuroscience, 5(11), 1242–1247.
Gilboa, A., Shalev, A., Laor, L., Lester, H., Louzoun, Y., Chisin, R., . . . Bonne, O.(2004). Functional connectivity of the prefrontal cortex
and the amygdala in posttraumatic stress disorder. Biological Psychiatry, 55(3), 263–272.
Gilliland, B. E., & James, R. K. (1998). Theories and strategies in counseling and psychotherapy. Boston: Allyn and Bacon.
Gintis, H. (2003). The hitchhiker’s guide to altruism: Gene-culture coevolution, and the internalization of norms. Journal of Theoretical
Biology, 220, 407–418. doi:10.1006/jtbi.2003.3104
Gitlin, M. J. (2007). The psychotherapist’s guide to psychopharmacology. New York: Free Press.
Glascher, J., Hampton, A. N., & O’Doherty, J. P. (2009). Determining a role for ventromedial prefrontal cortex in encoding action-based
value signals during reward-related decision making. Cerebral Cortex, 19(2), 483–495. doi:10.1093/cercor/bhn098
Glaser, D. (2000). Child abuse and neglect and the brain—a review. Journal of Child Psychiatry and Allied Disciplines, 41(1), 97–116.
Gloor, P. (1978). Inputs and outputs of the amygdala: What the amygdala is trying to tell the rest of the brain. In K. E. Livingston & O.
Hornykiewicz (Eds.), Limbic mechanisms: The continuing evolution of the limbic system concept (pp. 189–209). New York: Plenum.
Goel, V., & Dolan, R. J. (2003). Reciprocal neural response within lateral and ventral medial prefrontal cortex during hot and cold
reasoning. NeuroImage, 20(4), 2314–2321.
Goel, V., Grafman, J., Sadato, N., & Hallett, M. (1995). Modeling other minds. NeuroReport, 6(13), 1741–1746.
Goldapple, K., Segal, Z., Garson, C., Lau, M., Bieling, P., Kennedy, S., . . . Mayberg H. (2004). Modulation of cortical-limbic pathways in
major depression. Archives of General Psychiatry, 61(1), 34–41.
Goldberg, E., & Costa, L. D. (1981). Hemispheric differences in the acquisition and use of descriptive systems. Brain and Language, 14,
144–173.
Goldin, P. R., Ziv, M., Jazaieri, H., Weeks, J., Heimberg, R. G., & Gross, J. J. (2014). Impact of cognitive-behavioral therapy for social
anxiety disorder on the neural bases of emotional reactivity to and regulation of social evaluation. Behaviour Research and Therapy, 62, 97–106.
doi:10.1016/j.brat.2014.08.005
Goldman, A., & De Vignemont, F. (2009). Is social cognition embodied? Trends in Cognitive Sciences, 13(4), 218-232.
doi:10.1093/acprof:osobl/ 9780199874187.003.0010
Goldman, P. S. (1971). Functional development of the prefrontal cortex in early life and the problem of neural plasticity. Experimental
Neurology, 32, 366–387.
Goldman, P. S., & Galkin, T. W. (1978). Prenatal removal of frontal association cortex in the fetal rhesus monkey: Anatomical and
functional consequences in postnatal life. Brain Research, 152(3), 451–485.
Goldstein, K. (1939). The organism: A holistic approach to biology derived from pathological data in man. New York: American Books.
Goldstein, L. E., Rasmusson, A. M., Bunney, B. S., & Roth, R. H. (1996). Role of the amygdala in the coordination of behavioral,
neuroendocrine, and prefrontal cortical monoamine responses to psychological stress in the rat. Journal of Neuroscience, 16(15), 4787–4798.
Goleman, D. (2006). Emotional intelligence (10th ed.). New York: Bantam.
Golomb, J., de Leon, M. J., Kluger, A., George, A. E., Tarshish, C., & Ferris, S. H. (1993). Hippocampal atrophy in normal aging: An
association with recent memory impairment. Archives of Neurology, 50(9), 967–973.
Goodman, R. R., Snyder, S. H., Kuhar, M. J., & Young, W. S., III. (1980). Differential of delta and mu opiate receptor localizations by
light microscope autoradiography. Proceedings of the National Academy of Sciences, USA, 77, 2167–2174.

255
Goodwin, S., Blackman, R., Sakellaridi, S., & Chafee, M. (2012). Executive control over cognition: Stronger and earlier rule-based
modulation of spatial category signals in prefrontal cortex relative to parietal cortex. Journal of Neuroscience, 32(10), 3499–3515.
doi:10.1523/jneurosci.3585-11.2012
Gorka, A. X., Hanson, J. L., Radtke, S. R., & Hariri, A. R. (2014). Reduced hippocampal and medial prefrontal gray matter mediate the
association between reported childhood maltreatment and trait anxiety in adulthood and predict sensitivity to future life stress. Biology of Mood
and Anxiety Disorders, 4(1), 12. doi:10.1186/2045-5380-4-12
Gottfried, J. A., & Dolan, R. J. (2004). Human orbitofrontal cortex mediates extinction learning while accessing conditioned
representations of value. Nature Neuroscience, 7(10), 1145–1153.
Gould, E. (2007). How widespread is adult neurogenesis in mammals? Nature Reviews Neuroscience, 8, 481–488.
Gould, E., McEwen, B. S., Tanapat, P., Galea, L. A. M., & Fuchs, E. (1997). Neurogenesis in the dentate gyrus of the adult tree shrew is
regulated by psychosocial stress and NMDA receptor activation. Journal of Neuroscience, 17(7), 2492–2498.
Gould, E., Reeves, A. J., Fallah, M., Tanapat, P., Gross, C. G., & Fuchs, E. (1999). Hippocampal neurogenesis in adult old world
primates. Proceedings of the National Academy of Sciences, USA, 96(9), 5263–5267.
Gould, E., Reeves, A. J., Graziano, M. S. A., & Gross, C. G. (1999). Neurogenesis in the neocortex of adult primates. Science, 628(5439),
548–552.
Gould, E., Tanapat, P., Hastings, N. B., & Shors, T. J. (1999). Neurogenesis in adulthood: A possible role in learning. Trends in Cognitive
Sciences, 3(5), 186–191.
Gould, E., Woolley, C., & McEwan, B. (1990). Short-term glucocorticoid manipulations affect neuronal morphology and survival in the
adult dentate gyrus. Neuroscience, 37(2), 367–375.
Gould, S. J. (1977). Ontogeny and phylogeny. Cambridge, MA: Belknap.
Goyer, P. F., Andreason, P. J., Semple, W. E., Clayton, A. H., King, A. C., Compton-Toth, B. A., . . . Cohen, R. M. (1994). Positron-
emission tomography and personality disorders. Neuropsychopharmacology, 10, 21–28.
Grafton, S. T., Arbib, M. A., Fadiga, L., & Rizzolatti, G. (1996). Localization of grasp representations in humans by positron emission
tomography. 2: Observation compared with imagination. Experimental Brain Research, 112, 103–111.
Gray, J. R., Braver, T. S., & Raichle, M. E. (2002). Integration of emotion and cognition in the lateral prefrontal cortex. Proceedings of the
National Academy of Sciences, USA, 99(6), 4115–4120.
Green, A. (1978). Self-destructive behavior in battered children. American Journal of Psychiatry, 135(5), 579–582.
Green, A. (1981). Neurological impairments in maltreated children. Child Abuse and Neglect, 5, 129–134.
Greenough, W. T. (1987). Experience effects on the developing and mature brain: Dendritic branching and synaptogenesis. In N. A.
Krasnegor, E. M. Blass, M. A. Hofer, & W. P. Smotherman (Eds.), Perinatal development: A psychobiological perspective (pp. 195–221). Orlando:
Academic Press.
Grefkes, C., & Fink, G. R. (2005). The functional organization of the intra-parietal sulcus in humans and monkeys. Journal of Anatomy,
207(1), 3–17.
Greicius, M. D., Krasnow, B., Reiss, A. L., & Menon, V. (2003). Functional connectivity in the resting brain: A network analysis of the
default mode hypothesis. Proceedings of the National Academy of Sciences, 100(1), 253-258. doi:10.1073/pnas.0135058100
Greicius, M. D., & Menon, V. (2004). Default-mode activity during a passive sensory task: Uncoupled from deactivation but impacting
activation. Journal of Cognitive Neuroscience, 16(9), 1484-1492. doi:10.1162/0898929042568532
Greicius, M. D., Supekar, K., Menon, V., & Dougherty, R. F. (2008). Resting-state functional connectivity reflects structural connectivity
in the default mode network. Cerebral Cortex, 19(1), 72-78. doi:10.1093/cercor/bhn059
Griffiths, T. D., Rees, G., Rees, A., Green, G., Witton, C., Rowe, D., . . . Frackowiak, R. S. J. (1998). Right parietal cortex is involved in
the perception of sound movement in humans. Nature Neuroscience, 1(1), 74–79.
Grimm, S., Boesiger, P., Beck, J., Schuepbach, D., Bermpohl, F., Walter, M., . . . Northoff, G. (2008). Altered negative BOLD responses
in the default-mode network during emotion processing in depressed subjects. Neuropsychopharmacology, 34(4), 932-843.
doi:10.1038/npp.2008.81
Grimm, S., Ernst, J., Boesiger, P., Schuepbach, D., Boeker, H., & Northoff, G. (2011). Reduced negative BOLD responses in the default-
mode network and increased self-focus in depression. The World Journal of Biological Psychiatry, 12(8), 627-637.
doi:10.3109/15622975.2010.545145
Grimm, S., Ernst, J., Boesiger, P., Schuepbach, D., Hell, D., Boeker, H., & Northoff, G. (2009). Increased self-focus in major depressive
disorder is related to neural abnormalities in subcortical-cortical midline structures. Human Brain Mapping Hum. Brain Mapp., 30(8), 2617-
2627. doi:10.1002/hbm .20693
Grimshaw, G. M., & Carmel, D. (2014, May 23). An asymmetric inhibition model of hemispheric differences in emotional processing.
Frontiers in Psychology, 5. doi:10.3389/fpsyg.2014.00489

256
Grinband, J., Savitskaya, J., Wager, T. D., Teichert, T., Ferrera, V. P., & Hirsch, J. (2011). The dorsal medial frontal cortex is sensitive to
time on task, not response conflict or error likelihood. NeuroImage, 57(2), 303-311. doi:10.1016/j.neuroimage.2010.12.027
Grisaru, N., Chudakov, B., Yaroslavsky, Y., & Belmaker, R. H. (1998). Transcranial magnetic stimulation in mania: A controlled study.
American Journal of Psychiatry, 155(11), 1608–1610.
Gross, C. G. (2000). Neurogenesis in the adult brain: Death of a dogma. Nature Review of Neuroscience, 1, 67–73.
Gujar, N., Yoo, S., Hu, P., & Walker, M. P. (2010). The un-rested resting brain: Sleep-deprivation alters activity within the default-mode
network. Journal of Cognitive Neuroscience, 22(8), 1637-1648. doi:10.1162/jocn.2009.21331
Gundel, H., Lopez-Sala, A., & Ceballos-Baumann, A. O. (2004). Alexithymia correlates with the size of the right anterior cingulate.
Psychosomatic Medicine, 66(1), 132–140.
Gunnar, M. R. (1992). Reactivity of the hypothalamic-pituitary-adrenocortical system to stressors in normal infants and children. Pediatrics,
90(Suppl. 3), 491–479.
Gunnar, M. R. (1998). Quality of care and buffering of neuroendocrine stress reactions: Potential effects on the developing human brain.
Preventive Medicine, 27, 208–211.
Gunnar, M. R., & Stone, C. (1984). The effects of positive maternal affect on infant responses to pleasant, ambiguous, and fear-provoking
toys. Child Development, 55(4), 1231–1236.
Guo, W., Liu, F., Zhang, J., Zhang, Z., Yu, L., Liu, J., . . . Xiao, C. (2014). Abnormal default-mode network homogeneity in first-
episode, drug-naive major depressive disorder. PLoS ONE, 9(3). doi:10.1371/journal.pone.0091102
Gurvits, T. V., Gilbertson, M. W., Lasko, N. B., Tarhan, A. S., Simeon, D., Maclin, M. L., . . . Pitman, R. K. (2000). Neurological soft
signs in chronic posttraumatic stress disorder. Archives of General Psychiatry, 57(2), 181–183.
Gusnard, D. A., & Raichle, M. E. (2001). Searching for a baseline: Functional imaging and the resting human brain. Nature Reviews
Neuroscience, 2, 685–694.
Guzowski, J. F., Setlow, B., Wagner, E. K., & McGaugh, J. L. (2001). Experience-dependent gene expression in the rat hippocampus after
spatial learning: A comparison of the immediate-early genes Arc, c-fos, and zif268. Journal of Neuroscience, 21(14), 5089–5098.
Hahn, T., Kircher, T., Straube, B., Wittchen, H., Konrad, C., Ströhle, A., . . . Lueken, U. (2015). Predicting treatment response to
cognitive behavioral therapy in panic disorder with agoraphobia by integrating local neural information. JAMA Psychiatry, 72(1), 68.
doi:10.1001/jamapsychiatry.2014.1741
Haier, R., Jung, R., Yeo, R., Head, K., & Alkire, M. (2004). Structural brain variation and general intelligence. NeuroImage, 23(1), 425–
433. doi:10.1016/j.neuroimage.2004.04.025
Hair, N. L., Hanson, J. L., Wolfe, B. L., & Pollak, S. D. (2015). Association of child poverty, brain development, and academic
achievement. Journal of the American Medical Association: Pediatrics, 169(9), 822. doi:10.1001/jamapediatrics.2015.1475
Halassa, M. M., Fellin, T., & Haydon, P. G. (2007). The tripartite synapse: Roles for gliotransmission in health and disease. Trends in
Molecular Medicine, 13(2), 54–63.
Halgren, E., Dale, A. M., Sereno, M. I., Tootell, R. B. H., Marinkovic, K., & Rosen, B. R. (1999). Location of human face-selective cortex
with respect to retinotopic areas. Human Brain Mapping, 7(1), 29–37.
Halgren, E., Walter, R. D., Cherlow, D. G., & Crandall, P. H. (1978). Mental phenomena evoked by electrical stimulation of the human
hippocampal formation and amygdala. Brain, 101(1), 83–117.
Hamilton, C. E. (2000). Continuity and discontinuity of attachment from infancy through adolescence. Child Development, 71(3), 690–694.
Hampden-Turner, C. (1981). Maps of the mind. New York: Macmillan.
Hane, A., & Fox, N. (2006). Ordinary variations in maternal caregiving influence human infants’ stress reactivity. Psychological Science, 17(6),
550–556.
Hanson, J. L., Chung, M. K., Avants, B. B., Shirtcliff, E. A., Gee, J. C., Davidson, R. J., & Pollak, S. D. (2010). Early stress is associated
with alterations in the orbitofrontal cortex: A tensor-based morphometry investigation of brain structure and behavioral risk. Journal of
Neuroscience, 30(22), 7466–7472. doi:10.1523/jneurosci.0859-10.2010
Hanson, J. L., Hair, N., Shen, D. G., Shi, F., Gilmore, J. H., Wolfe, B. L., & Pollak, S. D. (2013a). Family poverty affects the rate of
human infant brain growth. PLoS One, 8(12). doi:10.1371/journal.pone.0080954
Hanson, J. L., Hair, N., Shen, D. G., Shi, F., Gilmore, J. H., Wolfe, B. L., & Pollak, S. D. (2013b). Correction: Family poverty affects the
rate of human infant brain growth. PLoS One, 10(12), e80954. doi:10.1371/journal.pone.0146434
Hanson, J. L., Nacewicz, B. M., Sutterer, M. J., Cayo, A. A., Schaefer, S. M., Rudolph, K. D., . . . Davidson, R. J. (2015). Behavioral
problems after early life stress: Contributions of the hippocampus and amygdala. Biological Psychiatry, 77(4), 314–323.
doi:10.1016/j.biopsych.2014.04.020
Hardingham, G. E., & Bading, H. (2003). The yin and yang of NMDA receptor signaling. Trends in Neurosciences, 26(2), 81–89.
Hariri, A. R., Bookheimer, S. Y., & Mazziotta, J. C. (2000). Modulating emotional responses: Effects of a neocortical network on the

257
limbic system. NeuroReport, 11(1), 43–48.
Hariri, A. R., Drabant, E. M., & Weinberger, D. R. (2006). Imaging genetics: Perspectives from studies of genetically driven variation in
serotonin function and corticolimbic affective processing. Biological Psychiatry, 59(10), 888–897.
Hariri, A. R., Mattay, V. S., Tessitore, A., Fera, F., & Weinberger, D. R. (2003). Neocortical modulation of the amygdala response to
fearful stimuli. Biological Psychiatry, 53(6), 494–501.
Harlow, H. F., & Suomi, S. J. (1971). Social recovery by isolation-reared monkeys. Proceedings of the National Academy of Sciences, USA,
68(7), 1534–1538.
Harlow, J. (1868). Recovery from the passage of an iron bar through the head. Publication of the Massachusetts Medical Society, 2, 329–346.
Harmon-Jones, E., & Allen, J. J. B. (1998). Anger and frontal brain activity: EEG asymmetry consistent with approach motivation despite
negative affective valence. Journal of Personality and Social Psychology, 74, 1310–1316.
Harmon-Jones, E., Gable, P. A., & Peterson, C. K. (2010). The role of asymmetric frontal cortical activity in emotion-related phenomena:
A review and update. Biological Psychology, 84(3), 451-462. doi:10.1016/j.biopsycho.2009.08.010
Harmon-Jones, E., & Sigelman, J. (2001). State anger and prefrontal brain activity: Evidence that insult-related relative left-prefrontal
activation is associated with experienced anger and aggression. Journal of Personality and Social Psychology, 80(5), 797–803.
Harris, C. L., Dinn, W. M., & Marcinkiewicz, J. A. (2002). Partial seizure-like symptoms in borderline personality disorder. Epilepsy and
Behavior, 3, 433–438.
Harrison, B. J., Pujol, J., López-Solà, M., Hernández-Ribas, R., Deus, J., Ortiz, H., . . . Cardoner, N. (2008). Consistency and functional
specialization in the default mode brain network. Proceedings of the National Academy of Sciences, USA, 105(28), 9781–9786.
Hassabis, D., & Maguire, E. A. (2007). Deconstructing episodic memory with construction. Trends in Cognitive Sciences, 11(7), 299-306.
doi:10.1016/j.tics.2007.05.001
Hasselmo, M. E., Rolls, E. T., & Baylis, G. C. (1989). The role of expression and identity in the face-selective responses of neurons in the
temporal visual cortex of the monkey. Behavior Brain Research, 32(3), 203–218.
Hawkes, K., O’Connell, J. F., & Jones, N. G. B. (1997). Hadza women’s time allocation, offspring provisioning, and the evolution of long
post-menopausal life spans. Current Anthropology, 38(4), 551–577.
Hazan, C., & Shaver, P. R. (1990). Love and work: An attachment-theoretical perspective. Journal of Personality and Social Psychology, 59(2),
270–280.
Hazlett, E. A., New, A. S., Newmark, R., Haznedar, M. M., Lo, J. N., Speiser, L. J., . . . Buchsbaum, M. S. (2005). Reduced anterior and
posterior cingulate gray matter in borderline personality disorder. Biological Psychiatry, 58, 614–623.
Hazlett, E. A., Speiser, L. J., Goodman, M., Roy, M., Carrizal, M., Wynn, J. K., . . . New, A. S. (2007). Exaggerated affect-modulated
startle during unpleasant stimuli in borderline personality disorder. Biological Psychiatry, 62, 250–255.
Hebb, D. O. (1949). The organization of behavior: A neuropsychological theory. New York: Wiley.
Heft, H. (1989). Affordances and the body: An intentional analysis of Gibson’s ecological approach to visual perception. Journal for the
Theory of Social Behaviour, 19(1), 1–30.
Heider, F. (1958). The psychology of interpersonal relations. New York: Wiley.
Heimer, L., & Van Hoesen, G. W. (2006). The limbic lobe and its output channels: Implications for emotional functions and adaptive
behavior. Neuroscience and Biobehavioral Reviews, 30(2), 126–147.
Heimer, L., Van Hoesen, G. W., Trimble, M., & Zahm, D. S. (2008). Anatomy of neuropsychiatry: The new anatomy of the basal forebrain and
its implications for neuropsychiatric illness. Amsterdam: Academic Press.
Heinrichs, M., & Domes, G. (2008). Neuropeptides and social behavior: Effects of oxytocin and vasopressin in humans. Progress in Brain
Research, 170, 337–350.
Heinz, A., Braus, D. F., Smolka, M. N., Wrase, J., Puls, I., Hermann, D., . . . Büchel, C. (2005). Amygdala-prefrontal coupling depends on
a genetic variation of the serotonin transporter. Nature Neuroscience, 8(1), 20–21.
Heiser, M., Iacoboni, M., Maeda, F., Marcus, J., & Mazziotta, J. C. (2003). The essential role of Broca’s area in imitation. European Journal
of Neuroscience, 17(5), 1123-1128. doi:10.1046/j.1460-9568.2003.02530.x
Henry, R. R., Satz, P., & Saslow, E. (1984). Early brain damage and the ontogenesis of functional asymmetry. Early Brain Damage, 1, 253–
275.
Hensch, T. K. (2004). Critical period regulation. Annual Review of Neuroscience, 27, 549–579.
Henson, R. N. A., Shallice, T., & Dolan, R. J. (1999). Right prefrontal cortex and episodic memory retrieval: A functional MRI test of the
monitoring hypothesis. Brain, 122(7), 1367–1381.
Herman, B. A., & Panksepp, J. (1978). Effects of morphine and naloxone on separation distress and approach attachment: Evidence for
opiate mediation of social effect. Pharmacology, Biochemistry and Behavior, 9, 213–220.
Herman, J. L. (1992). Complex PTSD: A syndrome in survivors of prolonged and repeated trauma. Journal of Traumatic Stress, 5(3), 377–

258
391.
Herpertz, S. C., Dietrich, T. M., Wenning, B., Krings, T., Erberich, S. G., Willmes, K., . . . Sass, H. (2001). Evidence of abnormal
amygdala functioning in borderline personality disorder: A functional MRI study. Biological Psychiatry, 50, 292–298.
Herpertz, S. C., Kunert, H. J., Schwenger, U. B., & Sass, H. (1999). Affective responsiveness in borderline personality disorder: A
psychophysiological approach. American Journal of Psychiatry, 156, 1550–1556.
Herrick, S., Brown, J., & Concepcion, E. (2014). The impact of traumatic brain injury on sexual offending behavior. Journal of Law
Enforcement, 4(1), 1–4.
Herrmann, E., Call, J., Hernandez-Lloreda, M. V., Hare, B., & Tomasello, M. (2007). Humans have evolved specialized skills of social
cognition: The cultural intelligence hypothesis. Science, 317(5843), 1360–1366. doi:10.1126/science.1146282
Herschkowitz, N., Kegan, J., & Zilles, K. (1997). Neurobiological basis of behavioral development in the first year. Neuropediatrics, 28, 296–
306.
Hesse, E. (1999). The adult attachment interview: Historical and current perspectives. In J. Cassidy & P. R. Shaver (Eds.), Handbook of
attachment: Theory, research, and clinical applications (pp. 395–433). New York: Guilford.
Hirsten, W., & Ramachandran, V. S. (1997). Capgras syndrome: A novel probe for understanding the neural representation of the identity
and familiarity of persons. Proceedings of the Royal Society of London: Biological Sciences, 264(1380), 437–444.
Hodge, C. J., & Boakye, M. (2001). Biological plasticity: The future of science in neurosurgery. Neurosurgery, 48(1), 2–16.
Hoffman, A., & Spengler, D. (2014). DNA memories of early social life. Neuroscience, 264, 64–75. doi:10.1016/j.neuroscience.2012.04.003
Holtforth, M. G., Grawe, K., Egger, O., & Berking, M. (2005). Reducing the dreaded: Change of avoidance motivation in psychotherapy.
Psychotherapy Research, 15(3), 261–271.
Holthoff, V. A., Beuthien-Baumann, B., Zündorf, G., Triemer, A., Lüdecke, S., Winiecki, P., . . . Herholz, K. (2004). Changes in brain
metabolism associated with remission in unipolar major depression. Acta Psychiatrica Scandinavica, 110(3), 184–194.
Hood, K. E., Dreschel, N. A., & Granger, D. A. (2003). Maternal behavior changes after immune challenge of neonates with
developmental effects on adult social behavior. Developmental Psychobiology, 42(1), 17–34.
Hoppe, K. D. (1977). Split-brains and psychoanalysis. Psychoanalytic Quarterly, 46(2), 220–244.
Hoppe, K. D., & Bogen, J. E. (1977). Alexithymia in twelve commissurotomized patients. Psychotherapy and Psychosomatics, 28(1–4), 148–
155.
Horovitz, S. G., Braun, A. R., Carr, W. S., Picchioni, D., Balkin, T. J., Fukunaga, M., & Duyn, J. H. (2009). Decoupling of the brain’s
default mode network during deep sleep. Proceedings of the National Academy of Sciences, 106(27), 11376-11381. doi:10.1073/pnas.0901435106
Hoshaw, B. A., Malberg, J. E., & Lucki, I. (2005). Central administration of IGF-I and BDNF leads to long-lasting antidepressant-like
effects. Brain Research, 1037(1–2), 204–208.
Houdé, O., Rossi, S., Lubin, A., & Joliot, M. (2010). Mapping numerical processing, reading, and executive functions in the developing
brain: An fMRI meta-analysis of 52 studies including 842 children. Developmental Science, 13(6), 876-885. doi:10.1111/j.1467-
7687.2009.00938.x
Houston, R. J., Ceballos, N. A., Hesselbrock, V. M., & Bauer, L. O. (2005). Borderline personality disorder features in adolescent girls:
P300 evidence of altered brain maturation. Clinical Neurophysiology, 116, 1424–1432.
Hoyle, R. L., Bromberger, B., Groversman, H. D., Klauber, M. R., Dixon, S. D., & Snyder, J. M. (1983). Regional anesthesia during
newborn circumcision: Effect on infant pain response. Clinical Pediatrics (Philadelphia), 22(12), 813–818.
Hsu, F., Zhang, G., Raol, Y., Valentino, R., Coulter, D., & Brooks-Kayal, A. (2003). Repeated neonatal handling with maternal separation
permanently alters hippocampal GABA receptors and behavioral stress responses. Proceedings of the National Academy of Sciences, USA, 100(21),
12213–12218.
Hu, P., Fan, J., Xu, P., Zhou, S., Zhang, L., Tian, Y., & Wang, K. (2013). Attention network impairments in patients with focal frontal or
parietal lesions. Neuroscience Letters, 534, 177–181. doi:10.1016/j.neulet.2012.12.038
Huang, H., Gundapuneedi, T., & Rao, U. (2012). White matter disruptions in adolescents exposed to childhood maltreatment and
vulnerability to psychopathology. Neuropsychopharmacology, 37(12), 2693–2701. doi:10.1038/npp.2012.133
Huang, X., Huang, P., Li, D., Zhang, Y., Wang, T., Mu, J., . . . Xie, P. (2014). Early brain changes associated with psychotherapy in major
depressive disorder revealed by resting-state fMRI: Evidence for the top-down regulation theory. International Journal of Psychophysiology, 94(3),
437–444. doi:10.1016/j.ijpsycho.2014.10.011
Huang, Z. J., Kirkwood, A., Pizzarusso, T., Porciatti, V., Morales, B., Bear, M. F., . . . Tonegawa, S. (1999). BDNF regulates the
maturation of inhibition and the critical period of plasticity in mouse visual cortex. Cell, 98(6), 739–755.
Hubel, D. H., & Wiesel, T. N. (1962). Receptive field binocular interaction and functional architecture in the cat’s visual cortex. Journal of
Physiology, 160, 106–154.
Hugo, V. (2013). Le miserables. (L. Fahnestock & N. MacAfee, Trans.). New York: Penguin. (Original work published 1862)

259
Hurley, R. A., Taber, K. H., Zhang, J., & Hayman, L. A. (1999). Neuropsychiatric presentation of multiple sclerosis. Journal of
Neuropsychiatry and Clinical Neurosciences, 11(1), 5–7.
Husain, M., & Nachev, P. (2007). Space and the parietal cortex. Trends in Cognitive Sciences, 11(1), 30–36.
Huttenlocher, P. R. (1994). Synaptogenesis in human cerebral cortex. In G. Dawson & K. W. Fischer (Eds.), Human behavior and the
developing brain (pp. 137–152). New York: Guilford.
Iacoboni, M. (2008). Mirroring people. New York: Farrar, Straus and Giroux.
Iacoboni, M., Koski, L. M., Brass, M., Bekkering, H., Woods, R. P., Dubeau, M., . . . Rizzolatti, G. (2001). Reafferent copies of imitated
actions in the right superior temporal cortex. Proceedings of the National Academy of Sciences, 98(24), 13995-13999. doi:10.1073/pnas.241474598
Iacoboni, M., Lieberman, M., Knowlton, I., Moritz, M., Throop, C., & Fiske, A. (2004). Watching social interactions produces
dorsomedial prefrontal and medial parietal BOLD fMRI signal increases compared to a resting baseline. NeuroImage, 21(3), 1167–1173.
Ickes, B. R., Pham, T. M., Sanders, L. A., Albeck, D.S., Mohammed, A. H., & Grandholm, A. C. (2000). Long-term environmental
enrichment leads to regional increases in neurotrophin levels in rat brains. Experimental Neurology, 164, 45–52.
Iidaka, T., Harada, T., & Sadato, N. (2010). Forming a negative impression of another person correlates with activation in medial
prefrontal cortex and amygdala. Social Cognitive and Affective Neuroscience, 6(4), 516-525. doi:10.1093/scan/nsq072
Ince, P. G. (2001). Pathological correlates of late-onset dementia in a multicentre, community-based population in England and Wales.
Lancet, 357(9251), 169–175.
Ingvar, D. H. (1985). “Memory for the future”: An essay on the temporal organization of conscious awareness. Human Neurobiology, 4(3),
127–136.
Introini-Collison, I., & McGaugh, J. L. (1987). Naloxone and beta-endorphin alter the effects of post-training epinephrine on retention of
an inhibitory avoidance response. Psychopharmacology, 92, 229–235.
Irle, E., Exner, C., Thielen, K., Weniger, G., & Ruther, E. (1998). Obsessive-compulsive disorder and ventromedial frontal lesions:
Clinical and neuropsychological findings. American Journal of Psychiatry, 155(2), 255–263.
Irle, E., Lange, C., & Sachsse, U. (2005). Reduced size and abnormal asymmetry of parietal cortex in women with borderline personality
disorder. Biological Psychiatry, 57, 173–182.
Ito, Y., Teicher, M. H., Glod, C. A., Harper, D., Magnus, E., & Gelbard, H. A. (1993). Increased prevalence of electrophysiological
abnormalities in children with psychological, physical, and sexual abuse. Journal of Neuropsychiatry, 5(4), 401–408.
Izard, C. E., Porges, S. W., Simons, R. F., Haynes, O. M., Hyde, C., Parisi, M., & Cohen, B. (1991). Infant cardiac activity:
Developmental changes and relations with attachment. Developmental Psychology, 27(3), 432–439.
Jablonska, B., Gierdalski, M., Kossut, M., & Skangiel-Kramska, J. (1999). Partial blocking of NMDA receptors reduces plastic changes
induced by short-lasting classical conditioning in the SL barrel cortex of adult mice. Cerebral Cortex, 9, 222–231.
Jack, A. I., Dawson, A. J., Begany, K. L., Leckie, R. L., Barry, K. P., Ciccia, A. H., & Snyder, A. Z. (2013). fMRI reveals reciprocal
inhibition between social and physical cognitive domains. NeuroImage, 66, 385–401.
Jackson, D. C., Mueller, C. J., Dolski, I., Dalton, K. M., Nitschke, J. B., Urry, H. L., . . . Davidson, R. J. (2003). Now you feel it, now you
don’t: Frontal brain electrical asymmetry and individual differences in emotion regulation. Psychological Science, 14(6), 612–617.
Jackson, P. L., & Decety, J. (2004). Motor cognition: A new paradigm to study self-other interactions. Current Opinion in Neurobiology,
14(2), 259–263.
Jackson, P. L., Meltzoff, A. N., & Decety, J. (2005). How do we perceive the pain of others? A window into the neural processes involved
in empathy. NeuroImage, 24, 771–779.
Jacobs, B., Driscoll, L., & Schall, M. (1997). Life-span dendritic and spine changes in areas 10 and 18 of human cortex: A quantitative
Golgi study. Journal of Comparative Neurology, 386(4), 661–680.
Jacobs, B., Schall, M., & Scheibel, A. B. (1993). A quantitative dendritic analysis of Wernicke’s area in humans: II. Gender, hemispheric,
and environmental factors. Journal of Comparative Neurology, 327(1), 97–111.
Jacobs, B., & Scheibel, A. B. (1993). A quantitative dendritic analysis of Wernicke’s area in humans: I. Lifespan changes. Journal of
Comparative Neurology, 327(1), 83–96.
Jacobs, B. L., van Praag, H., & Gage, F. H. (2000). Depression and the birth and death of brain cells. American Scientist, 88(4), 340–345.
Jacobs, W. J., & Nadel, L. (1985). Stress-induced recovery of fears and phobias. Psychological Review, 92(4), 512–531.
Jang, J. H., Jung, W. H., Choi, J., Choi, C., Kang, D., Shin, N. Y., . . . Kwon, J. S. (2011). Reduced prefrontal functional connectivity in
the default mode network is related to greater psychopathology in subjects with high genetic loading for schizophrenia. Schizophrenia Research,
127(1-3), 58-65. doi:10.1016/j.schres.2010.12.022
Janoff-Bulman, R. (1992). Shattered assumptions: Towards a new psychology of trauma. New York: Free Press.
Jardri, R., Pins, D., Bubrovszky, M., Despretz, P., Pruvo, J., Steinling, M., & Thomas, P. (2007). Self awareness and speech processing: An
fMRI study. NeuroImage, 35(4), 1645-1653. doi:10.1016/j.neuroimage.2007.02.002

260
Jason, G., & Pajurkova, E. (1992). Failure of metacontrol: Breakdown in behavioral unity after lesions of the corpus callosum and
inferomedial frontal lobes. Cortex, 28, 241–260.
Jaynes, J. (1976). The origin of consciousness in the breakdown of the bicameral mind. Boston: Houghton Mifflin.
Jeannerod, M. (2001). Neural Simulation of Action: A Unifying Mechanism for Motor Cognition. NeuroImage, 14(1).
doi:10.1006/nimg.2001.0832
Jeannerod, M., Arbib, M. A., Rizzolatti, G., & Sakata, H. (1995). Grasping objects: The cortical mechanisms of visuomotor
transformation. Trends in Neurosciences, 18(7), 314–320.
Jellema, T., Baker, C. I., Wicker, B., & Perrett, D. I. (2000). Neural representation for the perception of the intentionality of actions. Brain
and Cognition, 44(2), 280–302.
Jellema, T., Maassen, F., & Perrett, D. I. (2004). Single cell integration of animate form, motion and location in the superior temporal
cortex of the macaque monkey. Cerebral Cortex, 14(7), 781–790.
Jellema, T., & Perrett, D. I. (2003). Perceptual history influences neural responses to face and body postures. Journal of Cognitive
Neuroscience, 15, 961–971.
Jenkins, A. C., & Mitchell, J. P. (2011). Medial prefrontal cortex subserves diverse forms of self-reflection. Social Neuroscience, 6(3), 211-
218. doi:10.1080/17470919.2010.507948
Ji, J., & Maren, S. (2007). Hippocampal involvement in contextual modulation of fear extinction. Hippocampus, 17(9), 749–758.
Johanson, A., Gustafson, L., Passant, U., Risberg, J., Smith, G., Warkentin, S., & Tucker, D. (1998). Brain function in spider phobia.
Psychiatry Research: Neuroimaging Section, 84(2–3), 101–111.
Johanson, A., Risberg, J., Tucker, D. M., & Gustafson, L. (2006). Changes in frontal lobe activity with cognitive therapy for spider phobia.
Applied Neuropsychology, 13(1), 34–41.
Johansson, B. B. (2000). Brain plasticity and stroke rehabilitation: The Willis lecture. Stroke, 31(1), 223–230.
Johnson, A. (2006). Healing shame. The Humanistic Psychologist, 34(3), 223-242. doi:10.1207/s15473333thp3403_2
Johnson, M. (1987). The body in the mind. Chicago: University of Chicago Press.
Johnson, P. A., Hurley, R. A., Benkelfat, C., Herpertz, S. C., & Taber, K. H. (2003). Understanding emotion regulation in borderline
personality disorder: Contributions of neuroimaging. Journal of Neuropsychiatry of Clinical Neuroscience, 15, 397–402.
Johnstone, T., van Reekum, C. M., Urry, H. L., Kalin, N. H., & Davidson, R. J. (2007). Failure to regulate: Counterproductive recruitment
of top-down prefrontal-subcortical circuitry in major depression. Journal of Neuroscience, 27(33), 8877–8884.
Jonides, J., Schumacher, E. H., Smith, E. E., Koeppe, R. A., Awh, E., Reuter-Lorenz, P. A., Willis, C. R. (1998). The role of parietal
cortex in verbal working memory. Journal of Neuroscience, 18(3), 5026–5034.
Joseph, R. (1996). Neuropsychiatry, neuropsychology, and clinical neuroscience. Baltimore: Williams and Wilkins.
Jost, J. T., & Amodio, D. M. (2012). Political ideology as motivated social cognition: Behavioral and neuroscientific evidence. Motivation
and Emotion, 36, 55–64. doi:10.1007/s11031-011-9260-7
Judd, P. H. (2005). Neurocognitive impairment as a moderator in the development of borderline personality disorder. Development and
Psychopathology, 17, 1173–1196.
Juengling, F. D., Schmahl, C., Hesslinger, B., Ebert, D., Bremner, J. D., Gostomzyk, J., . . . Lieb, K. (2003). Positron emission tomography
in female patients with borderline personality disorder. Journal of Psychiatric Research, 37, 109–115.
Jung, R. E., & Haier, R. J. (2007). The pirate-frontal integration theory (P-FIT) of intelligence: Converging neuroimaging evidence. The
Behavioral and Brain Sciences, 30(2), 135–154. doi:10.1017/S0140525X07001185
Kaehler, L. A., & Freyd, J. J. (2009). Borderline personality characteristics: A betrayal trauma approach. Psychological Trauma: Theory,
Research, Practice, and Policy, 1(4), 261–268.
Kaehler, L. A., & Freyd, J. J. (2011). Betrayal trauma and borderline personality characteristics: Gender differences. Psychological Trauma:
Theory, Research, Practice, and Policy, 4(4), 379–385.
Kaler, S. R., & Freeman, B. J. (1994). Analysis of environmental deprivation: Cognitive and social development in Romanian orphans.
Journal of Child Psychology and Psychiatry, 35(4), 769–781. doi:10.1111/j.1469-7610.1994.tb01220.x
Kalia, M. (2005). Neurobiological basis of depression: An update. Metabolism, 54(5), 24–27.
Kalin, N. H., Larson, C., Shelton, S. E., & Davidson, R. J. (1998). Asymmetric frontal brain activity, cortisol, and behavior associated with
fearful temperament in rhesus monkeys. Behavioral Neuroscience, 112(2), 286–292.
Kalin, N. H., Shelton, S. E., Davidson, R. J., & Kelley, A. E. (2001). The primate amygdala mediates acute fear but not the behavioral and
physiological components of anxious temperament. Journal of Neuroscience, 21(6), 2067–2074.
Kalin, N. H., Shelton, S. E., & Lynn, D. E. (1995). Opiate systems in mother and infant primates coordinate intimate contact during
reunion. Psychoneuroendocrinology, 20(7), 735–742.
Kalin, N. H., Shelton, S. E., & Snowdon, C. T. (1993). Social factors regulating security and fear in infant rhesus monkeys. Depression,

261
1(3), 137–142.
Kalinichev, M., Easterlin, K., Plotsky, P., & Holtzman, S. (2002). Long-lasting changes in stress-induced corticosteron response and
anxiety-like behaviors as a consequence of neonatal maternal separation in Long-Evans rats. Pharmacology, Biochemistry and Behavior, 73(1),
131–141.
Kalisch, R., Korenfeld, E., Stephan, K. E., Weiskopf, N., Seymour, B., & Dolan, R. J. (2006). Context-dependent human extinction
memory is mediated by a ventromedial prefrontal and hippocampal network. Journal of Neuroscience, 26(37), 9503–9511.
Kampe, K. K. W., Frith, C. D., Dolan, R. J., & Frith, U. (2001). Reward value of attractiveness and gaze. Nature, 413(6856), 589–590.
Kanai, R., Feilden, T., Firth, C., & Rees, G. (2011). Political orientations are correlated with brain structure in young adults. Current
Biology, 21, 677–680. doi:10.1016/j.cub.2011.03.017
Kandel, E. R. (1998). A new intellectual framework for psychiatry. American Journal of Psychiatry, 155(4), 457–469.
Kandel, E., Dudai, Y., & Mayford, M. (2014). The molecular and systems biology of memory. Cell, 157(1), 163–186.
doi:10.1016/j.cell.2014.03.001
Kang, H., & Schuman, E. (1995). Long-lasting neurotrophin-induced enhancement of synaptic transmission in the adult hippocampus.
Science, 267(5204), 1658–1662.
Kaplan, H. S., & Robson, A. J. (2002). The emergence of humans: The coevolution of intelligence and longevity with intergenerational
transfers. Proceedings of the National Academy of Sciences, USA, 99(15), 10221–10226.
Karmiloff-Smith, A., Klima, E., Bellugi, U., Grant, J., & Baron-Cohen, S. (1995). Is there a social module? Language, face processing, and
theory of mind in individuals with Williams syndrome. Journal of Cognitive Neuroscience, 7(2), 196–208.
Karnath, H. O. (1997). Spatial orientation and the representation of space with parietal lobe lesions. Philosophical Transactions of the Royal
Society, Biological Sciences, 352(1360), 1411–1419.
Karni, A., Meyer, G., Jezzard, P., Adams, M. M., Turner, R., & Ungerleider, L. G. (1995). Functional MRI evidence for adult cortex
plasticity during motor skill learning. Nature, 377, 155–158.
Karten, Y. J. G., Olariu, A., & Cameron, H. A. (2005). Stress in early life inhibits neurogenesis in adulthood. Trends in Neurosciences, 28(4),
171–172.
Katz, L. C., & Shatz, C. J. (1996). Synaptic activity and the construction of cortical circuits. Science, 274(5290), 1133–1138.
Katzman, R., Aronson, M., Fuld, P., Kawas, C., Brown, T., Morgenstern, H., . . . Ooi, W. (1989). Development of dementing illness in an
80-year-old volunteer cohort. Annals of Neurology, 25, 317–324.
Kay, C., & Green, J. (2012). Reactive attachment disorder following early maltreatment: Systematic evidence beyond the institution. Journal
of Abnormal Child Psychology, 41(4), 571–581. doi:10.1007/s10802-012-9705-9
Keenan, J. P., McCutcheon, B., Freund, S., Gallup, G. G., Sanders, G., & Pascual-Leone, A. (1999). Left hand advantage in a self-face
recognition task. Neuropsychologia, 37(12), 1421–1425.
Kehoe, P., & Blass, E. M. (1989). Conditioned opioid release in ten-day-old rats: Reversal of stress with maternal stimuli. Developmental
Psychobiology, 19, 385–398.
Kelly, A., Mullany, P. M., & Lynch, M. A. (2000). Protein synthesis in entorhinal cortex and long-term potentiation in dentate gyrus.
Hippocampus, 10(4), 431–437.
Kempermann, G., Kuhn, H. G., & Gage, F. H. (1997). More hippocampal neurons in adult mice living in an enriched environment.
Nature, 386, 493–495.
Kempermann, G., Kuhn, H. G., & Gage, F. H. (1998). Experience-induced neurogenesis in the senescent dentate gyrus. Journal of
Neuroscience, 18(9), 3206–3212.
Kennard, M. A. (1955). The cingulate gyrus in relation to consciousness. Journal of Nervous and Mental Disease, 121(1), 34–39.
Kennedy, S. H., Evans, K. R., Kruger, S., Mayberg, H. S., Meyer, J. H., McCann, S., . . . Vaccarino, F. (2001). Changes in regional brain
glucose metabolism measured with positron emission tomography after paroxetine treatment of major depression. American Journal of Psychiatry,
158(6), 899–905.
Kennedy, S. H., Konarski, J. Z., Segal, Z. V., Lau, M. A., Bieling, P. J., McIntyre, R. S., . . . Mayberg, H. (2007). Differences in brain
glucose metabolism between responders to CBT and venlafaxine in a 16-week randomized controlled trial. American Journal of Psychiatry,
164(5), 778–788.
Kern, S., Oakes, T., Stone, C., McAuliff,E., Kirschbaum, C., & Davidson, R. (2008). Glucose metabolic changes in the prefrontal cortex
are associated with HPA axis response to psychosocial stressor. Psychoneuroendocrinology, 33(4), 517–529.
Kerr, D. S., Huggett, A. M., & Abraham, W. C. (1994). Modulation of hippocampal long-term potentiation and long-term depression by
corticosteroid receptor activation. Psychobiology, 22(2), 123–133.
Kessler, R. C., Berglund, P., Demler, O., Jin, R., Koretz, D., Merikangas, K. R., . . . Wang, P. S. (2003). The epidemiology of major
depressive disorder: Results from the National Comorbidity Survey Replication (NCS-R). Journal of the American Medical Association, 289(23),

262
3095–3105.
Keverne, E. (2014). Significance of epigenetics for understanding brain development, brain evolution and behaviour. Neuroscience, 264, 207–
217. doi:10.1016/j.neuroscience.2012.11.030
Keverne, E. B., Martens, N. D., & Tuite, B. (1989). Beta-endorphin concentrations in cerebrospinal fluid of monkeys are influenced by
grooming relationships. Psychoneuroendocrinology, 14(1–2), 155–161.
Kilgard, M. P., & Merzenich, M. M. (1998). Cortical map reorganization enabled by nucleus basalis activity. Science, 279(5357), 1714–
1718.
Kilner, J. M., Neal, A., Weiskopf, N., Friston, K. J., & Frith, C. D. (2009). Evidence of mirror neurons in human inferior frontal gyrus.
Journal of Neuroscience, 29(32), 10153–10159.
Kim, H., Somerville, L. H., Johnstone, T., Alexander, A. L., & Whalen, P. J. (2003). Inverse amygdala and medial prefrontal cortex
responses to surprised faces. NeuroReport, 14(18), 2317–2322.
Kim, J. J., & Diamond, D. M. (2002). The stressed hippocampus, synaptic plasticity and lost memories. Nature Reviews Neuroscience, 3(6),
453–462.
Kim, J. J., Koo, J. W., Lee, H. J., & Han, J. S. (2005). Amygdalar inactivation blocks stress-induced impairments in hippocampal long-term
potentiation and spatial memory. Journal of Neuroscience, 25(6), 1532–1539.
Kim, J. J., Lee, H. J., Han, J., & Packard, M. G. (2001). Amygdala is critical for stress-induced modulation of hippocampal long-term
potentiation and learning. Journal of Neuroscience, 21(14), 5222–5228.
Kilner, J. M., Vargas, C., Duval, S., Blakemore, S., & Sirigu, A. (2004). Motor activation prior to observation of a predicted movement.
Nature Neuroscience, 7(12), 1299-1301. doi:10.1038/nn1355
Kimble, D. P. (1968). Hippocampus and internal inhibition. Psychological Bulletin, 70(5), 285–295.
King, V., & Elder, G. H., Jr. (1997). The legacy of grandparenting: Childhood experiences with grandparents and current involvement with
grandchildren. Journal of Marriage and the Family, 59(4), 848–859.
King-Casas, B., Sharp, C., Lomax-Bream, L., Lohrenz, T., Fonagy, P., & Montague, P. R. (2008). The rupture and repair of cooperation
in borderline personality disorder. Science, 321(5890), 806–810.
Kinsley, C. H., Trainer, R., Stafisso-Sandoz, G., Quadros, P., Keyser Marcus, L., Hearon, C., . . . Lambert, K. G. (2006). Motherhood and
the hormones of pregnancy modify concentrations of hippocampal neuronal dendritic spines. Hormones and Behaviour, 49(2), 131–142.
Kirkpatrick, L. A., & Davis, K. E. (1994). Attachment style, gender, and relationship stability: A longitudinal analysis. Journal of Personality
and Social Psychology, 66(3), 502–512.
Kirkwood, A., Rozas, C., Kirkwood, J., Perez, F., & Bear, M. F. (1999). Modulation of long-term synaptic depression in visual cortex by
acetylcholine and norepinephrine. Journal of Neuroscience, 19(5), 1599–1609.
Kirmayer, L. J., & Carroll, J. (1987). A neurobiological hypothesis on the nature of chronic self-mutilation. Integrative Psychiatry, 5, 212–
213.
Kirschbaum, C., Wolf, O. T., May, M., Wippich, W., & Hellhammer, D. H. (1996). Stress and treatment-induced elevations of cortisol
levels associated with impaired declarative memory in healthy adults. Life Sciences, 58(17), 1475–1483.
Klein, E., Kreinin, I., Chistyakov, A., Koren, D., Mecz, L., Marmur, S., . . . Feinsod, M. (1999). Therapeutic efficacy of right prefrontal
slow repetitive transcranial magnetic stimulation in major depression. Archives of General Psychiatry, 56(4), 315–320.
Klein, T. A., Endrass, T., Kathmann, N., Neurmann, J., von Cramon, Y., & Ullsperger, M. (2007). Neural correlates of error awareness.
NeuroImage, 34(4), 1774–1781.
Kling, A., & Steklis, H. D. (1976). A neural substrate for affiliative behavior in nonhuman primates. Brain Behaviors, 13(2–3), 216–238.
Klingberg, T. (2006). Development of a superior frontal-intraparietal network for visuo-spatial working memory. Neuropsychologia, 44(11),
2171–2177.
Klingberg, T., Forssberg, H., & Westerberg, H. (2002). Increased brain activity in frontal and parietal cortex underlies the development of
visuospatial working memory capacity during childhood. Journal of Cognitive Neuroscience, 14(1), 1–10.
Knight, R. T., & Grabowecky, M. (1995). Escape from linear time: Prefrontal cortex and conscious experience. In M. S. Gazzaniga (Ed.),
The cognitive neurosciences (pp. 1357–1372). Cambridge, MA: MIT Press.
Knight, R. T., Staines, R. W., Swick, D., & Chao, L. L. (1999). Prefrontal cortex inhibition and excitation in distributed neural networks.
Acta Psychologica, 101(2–3), 159–178.
Knowles, P. A., Conner, R. L., & Panksepp, J. (1989). Opiate effects on social behavior of juvenile dogs as a function of social deprivation.
Pharmacology, Biochemistry and Behavior, 33(3), 533–537.
Knutson, K. M., Mah, L., Manly, C. F., & Grafman, J. (2007). Neural correlates of automatic beliefs about gender and race. Human Brain
Mapping, 28(10), 915–930.
Koc ovská, E., Wilson, P., Young, D., Wallace, A. M., Gorski, C., Follan, M., . . . Minnis, H. (2013). Cortisol secretion in children with

263
symptoms of reactive attachment disorder. Psychiatry Research, 209(1), 74–77. doi:10.1016/j.psychres.2012.12.011
Koechlin, E., Ody, C., & Kouneiher, F. (2003). The architecture of cognitive control in the human prefrontal cortex. Science, 302(5648),
1181–1185.
Koenigs, M. (2012). The role of prefrontal cortex in psychopathy. Reviews in the Neurosciences, 23(3), 253–262. doi:10.1515/revneuro-2012-
0036
Koenigsberg, H. W., Siever, L. J., Lee, H., Pizzarello, S., New, A. S., Goodman, M., . . . Prohovnik, I. (2009). Neural correlates of emotion
processing in borderline personality disorder. Psychiatry Research: Neuroimaging, 172, 192–199.
Kohut, H. (1984). How does analysis cure? Chicago: University of Chicago Press.
Kolb, B., & Gibb, R. (1991). Environmental enrichment and cortical injury: Behavioral and anatomical consequences of frontal cortex
lesions. Cerebral Cortex, 1(2), 189–198.
Kolb, B., & Gibb, R. (2002). Frontal lobe plasticity and behavior. In T. Donald & T. Robert (Eds.), Principles of frontal lobe function (pp.
541–556). New York: Oxford University Press.
Kolb, B., & Whishaw, I. Q. (1998). Brain plasticity and behavior. Annual Review of Psychology, 49, 43–64.
Kong, J., Gollub, R. L., Rosman, I. S., Webb, J. M., Vangel, M. J., Kirsch, I., & Kaptchik, T. (2006). Brain activity associated with
expectancy-enhanced placebo analgesia as measured by functional magnetic resonance imaging. Journal of Neuroscience, 26(2), 381–388.
Konig, P., & Engel, A. K. (1995). Correlated firing in sensory-motor systems. Current Opinions in Neurobiology, 5(4), 511–519.
Koopman, C., Classen, C., & Spiegel, D. (1994). Predictors of posttraumatic stress symptoms among survivors of the Oakland/Berkeley,
Calif. firestorm. American Journal of Psychiatry, 151(6), 888–894.
Kosten, T., Lee, H., & Kim, J. (2007). Neonatal handling alters learning in adult male and female rats in a task-specific manner. Brain
Research, 1154, 144–153.
Koukkou, M., & Lehmann, D. (2006). Experience-dependent brain plasticity: A key concept for studying nonconscious decisions.
International Congress Series, 1286, 45–52.
Koziol, L., Budding, D., & Chidekel, D. (2011). From movement to thought: Executive function, embodied cognition, and the cerebellum.
Cerebellum, 11(2), 505–525. doi:10.1007/s12311-011-0321-y
Kringelbach, M. L. (2005). The human orbitofrontal cortex: Linking reward to hedonic experience. Nature Reviews Neuroscience, 6(9), 691–
702.
Kroger, J. K., Sabb, F. W., Fales, C. L., Bookeimer, S. Y., Cohen, M. S., & Holyoak, K. J. (2002). Recruitment of anterior dorsolateral
prefrontal cortex in human reasoning: A parametric study of relational complexity. Cerebral Cortex, 12(5), 477–485.
Krueger, F., Moll, J., Zahn, R., Heinecke, A., & Grafman, J. (2006). Event frequency modulates the processing of daily life activities in
human medial prefrontal cortex. Cerebral Cortex, 17(10), 2346–2353.
Krugers, H. J., Goltstein, P. M., van der Linden, S., & Joels, M. (2006). Blockade of glucocorticoid receptors rapidly restores hippocampal
CA1 synaptic plasticity after exposure to chronic stress. European Journal of Neuroscience, 23(11), 3051–3055.
Krystal, J. H., Bremner, J. D., Southwick, S. M., & Charney, D. S. (1998). The emerging neurobiology of dissociation: Implication for
treatment of posttraumatic stress disorder. In J. D. Bremner & C. R. Marmar (Eds.), Trauma, memory, and dissociation (pp. 321–364).
Washington, DC: American Psychiatric Press.
Kuhlmann, S., Piel, M., & Wolf, O. T. (2005). Impaired memory retrieval after psychosocial stress in healthy young men. Journal of
Neuroscience, 25(11), 2977–2982.
Kuhn, C. M., & Schanberg, S. M. (1998). Responses to maternal separation: Mechanisms and mediators. International Journal of
Developmental Neuroscience, 16(3–4), 261–270.
Kuijpers, K. F., van der Knaap, L. M., Winkel, F. W., Pemberton, A., & Baldry, A. C. (2011). Borderline traits and symptoms of post-
traumatic stress in a sample of female victims of intimate partner violence, Stress and Health, 27, 206–215.
Kukolja, J., Schlapfer, T., Keysers, C., Klingmuller, D., Maier, W., Fink, G., . . . Hurlemann, R. (2008). Modeling a negative response bias
in the human amygdala by noradrenergic-glucocorticoid interactions. Journal of Neuroscience, 28(48), 12868–12876.
Kytta, M. (2002). Affordances of children’s environments in the context of cities, small towns, suburbs and rural villages in Finland and
Belarus. Journal of Environmental Psychology, 22(1–2), 109–123.
Laatsch, L., Pavel, D., Jobe, T., Lin, Q., & Quintana, J. C. (1999). Incorporation of SPECT imaging in a longitudinal cognitive
rehabilitation therapy programme. Brain Injury, 13(8), 555–570.
LaBar, K. S., Gatenby, J. C., Gore, J. C., LeDoux, J. E., & Phelps, A. E. (1998). Human amygdala activation during conditioned fear
acquisition and extinction: A mixed-trial FMRI study. Neuron, 20(5), 937–945.
LaBar, K. S., LeDoux, J. E., Spencer, D. D., & Phelps, E. A. (1995). Impaired fear conditioning following unilateral temporal lobectomy in
humans. Journal of Neuroscience, 15(10), 6846–6855.
Lachmann, F. M., & Beebe B. A. (1996). Three principles of salience in the organization of the patient-analyst interaction. Psychoanalytic

264
Psychology, 13(1), 1–22.
Ladd, C., Thrivikraman, K., Hout, R., & Plotsky, P. (2005). Differential neuroendocrine responses to chronic variable stress in adult Long
Evans rats exposed to handling-maternal separation as neonates. Psychoneuroendocrinology, 30(6), 520–533.
Lahdenperä, M., Lummaa, V., Helle, S., Tremblay, M., & Russell, A. F. (2004). Fitness benefits of prolonged post-reproductive lifespan in
women. Nature, 428(6979), 178–181.
Lamm, C., Batson, C. D., & Decety, J. (2007). The neural substrate of human empathy: Effects of perspective-taking and cognitive
appraisal. Journal of Cognitive Neuroscience, 19(1), 42-58. doi:10.1162/jocn.2007.19.1.42
Lane, R. D., Reiman, E. M., Bradley, M. M., Lang, P. J., Ahern, G. L., Davidson, R. J., & Schwartz, G. E. (1997). Neuroanatomical
correlates of pleasant and unpleasant emotion. Neuropsychologia, 35, 1437–1444.
Lange, C., Kracht, L., Herholz, K., Sachsse, U., & Irle, E. (2005). Reduced glucose metabolism in temporo-parietal cortices of women with
borderline personality disorder. Psychiatric Research: Neuroimaging, 139, 115–126.
Langer, E. J. (1978). Rethinking the role of thought in social interaction. In J. H. Harvey, W. Ickes, & R. F. Kidd (Eds.), New directions in
attribution research (Vol. 2, pp. 35–58). Hillsdale, NJ: Erlbaum.
Langer, N., Pedroni, A., Gianotti, L. R., Hänggi, J., Knoch, D., & Jäncke, L. (2012). Functional brain network efficiency predicts
intelligence. Human brain mapping, 33(6), 1393–1406.
Langeslag, S. J., Schmidt, M., Ghassabian, A., Jaddoe, V. W., Hofman, A., Lugt, A. V., . . . White, T. J. (2012). Functional connectivity
between parietal and frontal brain regions and intelligence in young children: The Generation R study. Human Brain Mapping, 34(12), 3299-
3307. doi:10.1002/hbm.22143
Lanius, R. A., Bluhm, R. L., Coupland, N. J., Hegadoren, K. M., Rowe, B., Thã©Berge, J., . . . Brimson, M. (2010). Default mode
network connectivity as a predictor of post-traumatic stress disorder symptom severity in acutely traumatized subjects. Acta Psychiatrica
Scandinavica, 121(1), 33-40. doi:10.1111/j.1600-0447.2009.01391.x
Lanius, R. A., Williamson, P. C., Bluhm, R. L., Densmore, M., Boksman, K., Neufeld, R. W. J., . . . Menon, R. S. (2005). Functional
connectivity of dissociative responses in posttraumatic stress disorder: A functional magnetic resonance imaging investigation. Biological
Psychiatry, 57(8), 873–884.
Lanius, R. A., Williamson, P. C., Densmore, M., Boksman, K., Gupta, M. A., Neufeld, R. W., . . . Menon, R. S. (2001). Neural
correlation of traumatic memories in posttraumatic stress disorder: A functional MRI investigation. American Journal of Psychiatry, 158(11),
1920–1922.
Larson, C., Schaefer, H., Siegle, G., Jackson, C., Anderle, M., & Davidson, R. (2006). Fear is fast in phobic individuals: Amygdala
activation in response to fear-relevant stimuli. Biological Psychiatry, 60(4), 410–417.
Laufs, H., Krakow, K., Sterzer, P., Eger, E., Beyerle, A., Salek-Haddadi, A., & Kleinschmidt, A. (2003). Electroencephalographic
signatures of attentional and cognitive default modes in spontaneous brain activity fluctuations at rest. Proceedings of the National Academy of
Sciences, 100(19), 11053-11058. doi:10.1073/pnas.1831638100
Laureys, S. (2006). Tracking the recovery of consciousness from coma. Journal of Clinical Investigation, 116(7), 1823-1825.
doi:10.1172/jci29172
Lawson, D. M., Barnes, A. D., Madkins, J. P., & Francios-Lamonte, B. M. (2006). Changes in male partner abuser attachment styles in
group treatment. Psychotherapy: Theory, Research, Practice, Training, 43(2), 232–237.
Lázaro, L., Bargalló, N., Castro-Fornieles, J., Falcón, C., Andrés, S., Calvo, R., . . . Junque, C. (2009). Brain changes in children and
adolescents with obsessive-compulsive disorder before and after treatment: A voxel-based morphometric MRI study. Psychiatry Research:
Neuroimaging, 172(2), 140–146.
Lebow, M. A., & Chen, A. (2016). Overshadowed by the amygdala: The bed nucleus of the stria terminalis emerges as key to psychiatric
disorders. Molecular Psychiatry, 21(4), 450-463. doi:10.1038/mp.2016.1
Le Carret, N., Lafont, S., Letenneur, L., Dartigues, J. F., Mayo, W., & Fabrigoule, C. (2003). The effect of education on cognitive
performances and its implication for the constitution of the cognitive reserve. Developmental Neuropsychology, 23(3), 317–337.
LeDoux, J. E. (1986). Sensory systems and emotion: A model of affective processing. Integrative Psychiatry, 4(4), 237–243.
LeDoux, J. E. (1994). Emotion, memory and the brain. Scientific American, 270(6), 32–39.
LeDoux, J. E. (1996). The emotional brain. New York: Simon and Schuster.
LeDoux, J. E., Romanski, L. M., & Xagoraris, A. E. (1989). Indelibility of subcortical emotional memories. Journal of Cognitive
Neuroscience, 1(3), 238–243.
LeDoux, J. E., Wilson, D. H., & Gazzaniga, M. S. (1977). A divided mind: Observations on the conscious properties of the separated
hemispheres. Annals of Neurology, 2(5), 417–421.
Leech, R., Kamourieh, S., Beckmann, C. F., & Sharp, D. J. (2011). Fractionating the default mode network: Distinct contributions of the
ventral and dorsal posterior cingulate cortex to cognitive control. Journal of Neuroscience, 31(9), 3217-3224. doi:10.1523/jneurosci.5626-10.2011

265
Lee, R. D. (2003). Rethinking the evolutionary theory of aging: Transfers, not births, shape senescence in social species. Proceedings of the
National Academy of Sciences, USA, 100(16), 9637–9642.
Lee, T. M. C., Liu, H. L., Chan, C. C. H., Fang, S. Y., & Gao, J. H. (2005). Neural activities associated with emotion recognition
observed in men and women. Molecular Psychiatry, 10(5), 450–455.
Lee, T. W., Josephs, O., Dolan, R. J., & Critchley, H. D. (2006). Imitating expressions: Emotion-specific neural substrates in facial
mimicry. SCAN, 1, 122–135.
Lee, Y., & Davis, M. (1997). Role of the hippocampus, the bed nucleus of the stria terminalis and the amygdala in the excitatory effect of
corticotropin-releasing hormone on the acoustic startle reflex. Journal of Neuroscience, 17(16), 6434–6446.
Lemer, C., Dehaene, S., Spelke, E., & Cohen, L. (2003). Approximate quantities and exact number words: Dissociable systems.
Neuropsychologica, 41(14), 1942–1958.
Leonard, C. M., Rolls, E. T., Wilson, F. A. W., & Baylis, G. C. (1985). Neurons in the amygdala of the monkey with responses selective
for faces. Behavioral Brain Research, 15(2), 159–176.
Lerner, A., Bagic, A., Hanakawa, T., Boudreau, E. A., Pagan, F., Mari, Z., . . . Hallett, M. (2009). Involvement of insula and cingulate
cortices in control and suppression of natural urges. Cerebral Cortex, 19(1), 218–223.
Leventopoulos, M., Rüedi-Bettschen, D., Knuesel, I., Feldon, J., Pryce, C. R., & Opacka-Juffry, J. (2007). Long-term effects of early life
deprivation on brain glia in Fischer rats. Brain Research, 1142, 119–126.
Lévesque, J., Eugène, F., Joanette, Y., Paquette, V., Mensour, B., Beaudoin, G., . . . Beauregard, M. (2003). Neural circuitry underlying
voluntary suppression of sadness. Biological Psychiatry, 53(6), 502–510.
Lévesque, J., Joanette, Y., Mensour, B., Beaudoin, G., Leroux, J. M., Bourgoin, P., . . . Beauregard, M. (2003). Neural correlates of sad
feelings in healthy girls. Neuroscience, 121(3), 545–551.
Lévesque, J., Joanette, Y., Mensour, B., Beaudoin, G., Leroux, J. M., Bourgouin, P., & Beauregard, M. (2004). Neural basis of emotional
self-regulation in childhood. Neuroscience, 129(2), 361–369.
Levin, P., Lazrove, S., & van der Kolk, B. (1999). What psychological testing and neuroimaging tell us about the treatment of posttraumatic
stress disorder by eye movement desensitization and reprocessing. Journal of Anxiety Disorders, 13(1–2), 159–172.
Levy, D. A. (1997). Tools of critical thinking. Boston: Allyn and Bacon.
Levy, J., Trevarthen, C., & Sperry, R. W. (1972). Perception of bilateral chimeric figures following hemispheric disconnection. Brain, 95,
61–78.
Lewicki, P., Hill, T., & Czyzewska, M. (1992). Nonconscious acquisition of information. American Psychologist, 47(6), 796–801.
Lewis, M., Feiring, C., & Rosenthal, S. (2000). Attachment over time. Child Development, 71(3), 707–720.
Lewis, P., Rezaie, R., Brown, R., Roberts, N., & Dunbar, R. (2011). Ventromedial prefrontal volume predicts understanding of others and
social network size. NeuroImage, 57(4), 1624–1629. doi:10.1016/j.neuroimage.2011.05.030
Li, H., Weiss, S. R. B., Chaung, D. M., Post, R. M., & Rogawski, M. A. (1998). Bidirectional synaptic plasticity in the rat basolateral
amygdala: Characterization of an activity-dependent switch sensitive to the presynaptic metabotropic glutamate receptor antagonist 2S-alpha-
ethyglutamic acid. Journal of Neuroscience, 18, 1662–1670.
Li, W., Mai, X., & Liu, C. (2014). The default mode network and social understanding of others: What do brain connectivity studies tell
us. Frontiers in Human Neuroscience Front. Hum. Neurosci., 8. doi:10.3389/fnhum.2014.00074
Li, X., Jiang, J., Zhu, W., Yu, C., Sui, M., Wang, Y., & Jiang, T. (2007). Asymmetry of prefrontal cortical convolution complexity in males
with attention-deficit/hyperactivity disorder using fractal information dimension. Brain and Development, 29(10), 649–655.
Libet, B., Gleason, C. A., Wright, E. W., & Pearl, D. K. (1983). Time of conscious intention to act in relation to onset of cerebral activity
(readiness-potential): The unconscious initiation of a freely voluntary act. Brain, 106, 623–642.
Libet, B., Gleason, C. A., Wright, E. W., & Pearl, D. K. (1993). Time of conscious intention to act in relation to onset of cerebral activity
(readiness-potential). Neurophysiology of Consciousness, 249-268. doi:10.1007/978-1-4612-0355-1_15
Lieb, K., Rexhausen, J. E., Kahhl, K. G., Schweiger, U., Philipsen, A., Hellhammer, D. H., & Bohus, M. (2004). Increased diurnal salivary
cortisol in women with borderline personality disorder. Journal of Psychiatric Research, 38, 559–565.
Lieberman, M. D., Eisenberger, N. I., Crockett, M. J., Tom, S. M., Pfeifer, J. H., & Way, B. M. (2007). Putting feelings into words:
Affect labeling disrupts amygdala activity in response to affective stimuli. Psychological Science, 18(5), 421–428.
Lim, L., Radua, J., & Rubia, K. (2014). Gray matter abnormalities in childhood maltreatment: A voxel-wise meta-analysis. American Journal
of Psychiatry, 171(8), 854–863. doi:10.1176/appi.ajp.2014.13101427
Linden, D. E. J. (2006). How psychotherapy changes the brain—the contribution of functional neuroimaging. Molecular Psychiatry, 11(6),
528–538.
Linehan, M. (1993). Cognitive-behavioral treatment of borderline personality disorder. New York: Guilford.
Liotti, M., Brannan, S., Egan, G., Shade, R., Madden, L., Abplanalp, B., . . . Denton, D. (2001). Brain responses associated with

266
consciousness of breathlessness (air hunger). PNAS, 98(4), 2035–2040.
Liotti, M., Mayberg, H. S., Brannan, S. K., McGinnis, S., Jerabek, P., & Fox, P. T. (2000). Differential limbic-cortical correlates of sadness
and anxiety in healthy subjects: Implications for affective disorders. Biological Psychiatry, 48, 30–42.
Liu, D., Diorio, J., Day, J. C., Francis, D. D., & Meaney, M. J. (2000). Maternal care, hippocampal synaptogenesis and cognitive
development in rats. Nature Neuroscience, 3, 799–806.
Liu, D., Diorio, J., Tannenbaum, B., Caldji, C., Francis, D., Freedman, A., . . . Meaney, M. J. (1997). Maternal care, hippocampal
glucocorticoid receptors, and hypothalamic-pituitary-adrenal responses to stress. Science, 277(5332), 1659–1662.
Liu, L., Wong, T. P., Pozza, M. F., Lingenhoehl, K., Wang, Y., Sheng, M., . . . Wang, Y. T. (2004). Role of NMDA receptor subtypes in
governing the direction of hippocampal synaptic plasticity. Science, 304(5673), 1021–1024.
Livingston, R. B. (1967). Reinforcement. In G. C. Quarton, T. Melnick, & F. O. Schmitt (Eds.), The neurosciences (pp. 568–576). New
York: Rockefeller University Press
Loftus, E. (1988). Memory. New York: Ardsley House.
Loftus, E. F., Milo, E. M., & Paddock, J. R. (1995). The accidental executioner: Why psychotherapy must be informed by science.
Counseling Psychologist, 23, 300–309.
Lombroso, P. J., & Sapolsky, R. (1998). Development of the cerebral cortex: Stress and brain development. Journal of the Academy of Child
and Adolescent Psychiatry, 37, 1337–1339.
Lonstein, J. S., Simmons, D. A., Swann, J. M., & Stern, J. M. (1998). Forebrain expression of c-fos due to active maternal behaviour in
lactating rats. Neuroscience, 82(1), 267–281.
Lorberbaum, J. P., Newman, J. D., Dubno, J. R., Horwitz, A. R., Nahas, Z., Teneback, C. C., . . . George, M. S. (1999). Feasability of
using fMRI to study mothers responding to infant cries. Depression and Anxiety, 10(3), 99–104.
Lord, C. G., Ross, L., & Lepper, M. (1979). Biased assimilation and attitude polarization: The effects of prior theories on subsequently
considered evidence. Journal of Personality and Social Psychology, 37(11), 1231–1247.
Lorenz, K. (1991). Here am I: where are you: The behavior of the Greylag Goose. New York: Brace Jovanovich.
Lou, H. C., Henriksen, L., & Bruhn, P. (1984). Focal cerebral hypoperfusion in children with dysphasia and/or attention deficit disorder.
Archives of Neurology, 41(8), 825–829.
Lou, H. C., Luber, B., Crupain, M., Keenan, J. P., Nowak, M., Kjaer, T. W., . . . Liansby, S. H. (2004). Parietal cortex and representation
of the mental self. Proceedings of the National Academy of Sciences, USA, 101(17), 6827–6832.
Lou, H., Nowak, M., & Kajaer, T. W. (2005). The mental self. Progress in Brain Resources, 150, 197–204.
Love, T. (2014). Oxytocin, motivation and the role of dopamine. Pharmacology Biochemistry and Behavior, 119, 49–60.
doi:10.1016/j.pbb.2013.06.011
Lovell, J., & Kluger, J. (1994). Lost moon: The perilous voyage of Apollo 13. New York: Simon and Schuster.
Lu, S. T., Hamalainen, M. S., Hari, R., Ilmoniemi, R. J., Lounasmaa, O. V., Sams, M., & Vilkman, V. (1991). Seeing faces activates three
separate areas outside the occipital visual cortex in man. Neuroscience, 43(2–3), 287–290.
Lueken, U., Straube, B., Konrad, C., Wittchen, H., Ströhle, A., Wittmann, A., . . . Kircher, T. (2013). Neural substrates of treatment
response to cognitive-behavioral therapy in panic disorder with agoraphobia. American Journal of Psychiatry, 170(11), 1345–1355.
doi:10.1176/appi.ajp.2013.12111484
Luna, B. (2004). Algebra and the adolescent brain. Trends in Cognitive Sciences, 8(10), 437–439.
Luo, S., Li, B., Ma, Y., Zhang, W., Rao, Y., & Han, S. (2015). Oxytocin receptor gene and racial ingroup bias in empathy-related brain
activity. NeuroImage, 110, 22–31. doi:10.1016/j.neuroimage.2015.01.042
Lupien, S., de Leon, M., de Santi, S., Convit, A., Tarshish, C., Nair, N., . . . Meaney, M. J. (1998). Cortisol levels during human aging
predict hippocampal atrophy and memory deficits. Nature Neuroscience, 1(1), 69–73.
Lupien, S. J., & McEwen, B. S. (1997). The acute effects of corticosteroids on cognition: Integration of animal and human model studies.
Brain Research Reviews, 24(1), 1–27.
Lupo, M., Troisi, E., Chiricozzi, F. R., Clausi, S., Molinari, M., & Leggio, M. (2015). Inability to process negative emotions in cerebellar
damage: A functional transcranial doppler sonographic study. Cerebellum, 14(6), 663–669. doi:10.1007/s12311-015-0662-z
Lutz, A., Brefczynski-Lewis, J., Johnstone T., & Davidson R. J. (2008). Regulation of the neural circuitry of emotion by compassion
meditation: Effects of meditative expertise. PLoS One, 3(3), 1–10.
Lutz, A., Greischar, L. L., Perlman, D. M., & Davidson, R. J. (2009). BOLD signal in insula is differentially related to cardiac function
during compassion meditation in experts vs. novices. NeuroImage, 47(3), 1038-1046. doi:10.1016/j.neuroimage.2009.04.081
Lynch, C. J., Uddin, L. Q., Supekar, K., Khouzam, A., Phillips, J., & Menon, V. (2013). Default Mode Network in Childhood Autism:
Posteromedial Cortex Heterogeneity and Relationship with Social Deficits. Biological Psychiatry, 74(3), 212-219.
doi:10.1016/j.biopsych.2012.12.013

267
Lyons-Ruth, K., Bureau, J., Riley, C. D., & Atlas-Corbett, A. F. (2009). Socially indiscriminate attachment behavior in the strange
situation: Convergent and discriminant validity in relation to caregiving risk, later behavior problems, and attachment insecurity. Developmental
Psychopathology, 21(2), 355. doi:10.1017/s0954579409000376
Lyons-Ruth, K., Choi-Kain, L., Pechtel, P., Bertha, E., & Gunderson, J. (2011). Perceived parental protection and cortisol responses
among young females with borderline personality disorder and controls. Psychiatry Research, 189, 426–432.
Lyoo, K., Han, M. H., & Cho, D. Y. (1998). A brain MRI study in subjects with borderline personality disorder. Journal of Affective
Disorders, 50, 235–243.
Maccari, S., Piazza, P. V., Kabbaj, M., Barbazanges, A., Simon, H., & Le Moal, M. (1995). Adoption reverses the long-term impairment
in glucocorticoid feedback induced by prenatal stress. Journal of Neuroscience, 15(1), 110–116.
Machizawa-Summers, S. (2007). Childhood trauma and parental bonding among Japanese female patients with borderline personality
disorder. Journal of Psychology, 42(4), 265–273. doi: 10.1080/00207590601109276
Mackie, S., Shaw, P., Lenroot, R., Pierson, R., Greenstein, D. K., Nugent, T. F., & Rapoport, J. L. (2007). Cerebellar development and
clinical outcome in attention deficit hyperactivity disorder. American Journal of Psychiatry, 164(4), 647–655.
MacLean, P. D. (1985). Brain evolution relating to family, play, and the separation call. Archives of General Psychiatry, 42(4), 405–417.
MacLean, P. D. (1990). The triune brain in evolution: Role of paleocerebral functions. New York: Plenum.
Macrae, C. N., Moran, J. M., Heatherton, T. F., Banfield, J. F., & Kelley, W. M. (2004). Medial prefrontal activity predicts memory for
self. Cerebral Cortex, 14(6), 647–654.
Maddock, R., Garrett, A., & Buonocore, M. (2001). Remembering familiar people: The posterior cingulate cortex and autobiographical
memory retrieval. Neuroscience, 104(3), 667-676. doi:10.1016/s0306-4522(01)00108-7
Maguire, E. A., & Mummery, C. J. (1999). Differential modulation of a common memory retrieval network revealed by positron emission
tomography. Hippocampus, 9(1), 54– 61.
Maguire, E. A., Woollett, K., & Spiers, H. J. (2006). London taxi drivers and bus drivers: A structural MRI and neuropsychological
analysis. Hippocampus, 16(12), 1091–1101.
Maher, B. A. (1974). Delusional thinking and perceptual disorder. Journal of Individual Psychology, 30(1), 98–113.
Maier, S. F., Amat, J., Baratta, M. V., Paul, E., & Watkins, L. R. (2006). Behavioral control, the medial prefrontal cortex, and resilience.
Dialogues in Clinical Neuroscience, 8(4), 397–406.
Main, M. (1993). Discourse, prediction, and the recent studies in attachment: Implications for psychoanalysis. Journal of the American
Psychoanalytic Association, 41, 209–244.
Main, M., & Goldwyn, R. (1998). Adult attachment scoring and classification system. Unpublished manuscript, University of California at
Berkeley.
Main, M., Kaplan, N., & 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), 66–104.
Malenka, R. C., & Siegelbaum, S. A. (2001). Synaptic plasticity: Diverse targets and mechanisms for regulating synaptic efficacy. In W. M.
Cowan, T. C. Sudhof, & C. F. Stevens (Eds.), Synapses (pp. 393–453). Baltimore: Johns Hopkins University Press.
Maletic-Savatic, M., Malinow, R., & Svoboda, K. (1999). Rapid dendritic morphogenesis in CA1 hippocampal dendrites induced by
synaptic activity. Science, 283(5409), 1923–1927.
Malloy, P., Bihrle, A., Duffy, J., & Cimino, C. (1993). The orbitomedial frontal syndrome. Archives of Clinical Neuropsychology, 8(3), 185–
201.
Manoliu, A., Riedl, V., Zherdin, A., Muhlau, M., Schwerthoffer, D., Scherr, M., . . . Sorg, C. (2013). Aberrant dependence of default
mode/central executive network interactions on anterior insular salience network activity in schizophrenia. Schizophrenia Bulletin, 40(2), 428-
437. doi:10.1093/schbul/sbt037
Mantini, D., Gerits, A., Nelissen, K., Durand, J., Joly, O., Simone, L., . . . Vanduffel, W. (2011). Default Mode of Brain Function in
Monkeys. Journal of Neuroscience, 31(36), 12954-12962. doi:10.1523/jneurosci.2318-11.2011
Marais, L., van Rensburg, S. J., van Zyl, J. M., Stein, D. J., & Daniels, W. M. U. (2008). Maternal separation of rat pups increases the risk
of developing depressive-like behavior after subsequent chronic stress by altering corticosterone and neurotrophin levels in the hippocampus.
Neuroscience Research, 61(1), 106–112.
Marci, C. D., Ham, J., Moran, E., & Orr, S. P. (2007). Physiologic correlates of perceived therapist empathy and social-emotional process
during psychotherapy. Journal of Nervous and Mental Disease, 195(2), 103–111.
Marien, H., Custers, R., Hassin, R., & Aarts, H. (2012). Unconscious goal activation and the hijacking of the executive function. Journal of
Personality and Social Psychology, 103(3), 399–415. doi:10.1037/a0028955
Markis, N., Biederman, J., Vatera, E., Bush, G., Kaiser, J., Kennedy, D. N., . . . Seidman, L. (2007). Cortical thinning of the attention and
executive function networks in adults with attention-deficit/hyperactivity disorder. Cerebral Cortex, 17(6), 1364–1375.

268
Marr, D. (1971). A theory of archicortex. Philosophical Transactions of the Royal Society, 262, 23–81.
Mars, R. B., Neubert, F., Noonan, M. P., Sallet, J., Toni, I., & Rushworth, M. F. (2012). On the relationship between the “default mode
network” and the “social brain”. Frontiers in Human Neuroscience, 6, 1-9. doi:10.3389/fnhum.2012.00189
Marsh, Abigail A., Sarah A. Stoycos, Kristin M. Brethel-Haurwitz, Paul Robinson, John W. Vanmeter, and Elise M. Cardinale. “Neural
and Cognitive Characteristics of Extraordinary Altruists.” Proceedings of the National Academy of Sciences Proc Natl Acad Sci USA 111.42 (2014):
15036-5041. Web.
Marshall, R. E., Stratton, W. C., Moore, J., & Boxerman, S. B. (1980). Circumcision I: Effects upon newborn behavior. Infant Behavioral
Development, 3, 1–14.
Marshuetz, C., Smith, E., Jonides, J., DeGutis, J., & Chenevert, T. (2000). Order information in working memory: fMRI evidence for
parietal and pre-frontal mechanisms. Journal of Cognitive Neuroscience, 12(2), 130–144.
Martin, A., Wiggs, C., Ungerleider, L., & Haxby, J. (1996). Neural correlates of category-specific knowledge. Nature, 379(6566), 649–652.
Martin, S. D., Martin, E., Rai, S. S., Richardson, M. A., & Royall, R. (2001). Brain blood flow changes in depressed patients treated with
interpersonal psychotherapy or venlafaxine hydrochloride. Archives of General Psychiatry, 58, 641–648.
Mason, M. F., Norton, M. I., Horn, J. D., Wegner, D. M., Grafton, S. T., & Macrae, C. N. (2007). Wandering minds: The default
network and stimulus-independent thought. Science, 315(5810), 393-395. doi:10.1126/science.1131295
Massey, P. V., Johnson, B. E., Moult, P. R., Auberson, Y. P., Brown, M. W., Molnar, E., . . . Bashir, Z. (2004). Differential roles of
NR2A and NR2B-containing NMDA receptors in cortical long-term potentiation and long-term depression. Journal of Neuroscience, 24(36),
7821–7828.
Mateer, C. A., & Kerns, K. A. (2000). Capitalizing on neuroplasticity. Brain and Cognition, 42(1), 106–109.
Mather, M., Canli, T., English, T., Whitfield, S., Wais, P., Ochsner, K., . . . Carstensen, L. L. (2004). Amygdala responses to emotionally
valenced stimuli in older and younger adults. Psychological Science, 15(4), 259-263. doi:10.1111/j.0956-7976.2004.00662.x
Mathew, R. J., Meyer, J. S., Francis, D. J., Semchuk, K. M., & Claghorn, J. L. (1980). Cerebral blood flow in depression. American Journal
of Psychiatry, 137, 1449–1450.
Matos, M., & Pinto-Gouveia, J. (2014). Shamed by a parent or by others: The role of attachment in shame memories relation to depression.
International Journal of Psychology and Psychological Therapy, 14(2), 217–244.
Matos, M., Pinto-Gouveia, J., & Costa, V. (2011). Understanding the importance of attachment in shame traumatic memory relation to
depression: The impact of emotion regulation processes. Clinical Psychology and Psychotherapy, 20(2), 149–165.
Matsumoto, K., & Tanaka, K. (2004). The role of the medial prefrontal cortex in achieving goals. Current Opinion in Neurobiology, 14(2),
178–185.
Maurex, L., Lekander, M., Nilsonne, A., Andersson, E. E., Åsberg, M., & Öhman, A. (2010). Social problem solving, autobiographical
memory, trauma, and depression in women with borderline personality disorder and a history of suicide attempts. British Journal of Clinical
Psychology, 49, 327–342.
Mayberg, H. S. (1997). Limbic-cortical dysregulation: A proposed model of depression. Journal of Neuropsychiatry, 9(3), 471–481.
Mayberg, H. S., Liotti, M., Brannan, S. K., McGinnis, S., Mahurin, R. K., Jerabek, P. A., . . . Fox, P. T. (1999). Reciprocal limbic-cortical
function and negative mood: Converging PET findings in depression and normal sadness. American Journal of Psychiatry, 156(5), 675–682.
Mayberg, H. S., Silva, J. A., Brannan, S. K., Tekell, J. L., Mahurin, R. K., McGinnis, S., & Jerabek, P. A. (2002). The functional
neuroanatomy of the placebo effect. American Journal of Psychiatry, 159(5), 728–737.
Mazza, J. J., & Reynolds, W. M. (1998). A longitudinal investigation of depression, homelessness, social support, and major and minor life
events and their relation to suicide ideation in adolescents. Suicide and Life Threatening Behavior, 28, 358–374.
McCarthy, G. (1995). Functional neuroimaging of memory. Neuroscientist, 1(3), 155–163.
McCormick, J. A., Lyons, V., Jacobson, M. D., Noble, J., Diorio, J., Nyirenda, M., & Chapman, K. E. (2000). 5’-Heterogeneity of
glucocorticoid receptor messenger RNA is tissue specific: Differential regulation of variant transcripts by early-life events. Molecular
Endocrinology, 14(4), 506–517.
McDonald, A. J., Shammah-Lagnado, S. J., Shi, C., & Davis, M. (1999). Cortical afferents to the extended amygdala. Annals of the New
York Academy of Sciences, 877, 309–338.
McFarlane, A. C., & Bookless, C. (2010). The effect of PTSD on interpersonal relationships: Issues for emergency service workers. Sexual
and Relationship Therapy, 16(3), 261–267. DOI: 10.1080/14681990124457
McFarlane, A. C., & Yehuda, R. (1996). Resilience, vulnerability, and the course of posttraumatic reactions. In B. A. van der Kolk, A. C.
McFarlane, & L. Weisaeth (Eds.), Traumatic stress: The effects of overwhelming experience on mind, body, and society (pp. 129–154). New York:
Guilford.
McGaugh, J. L. (1990). Significance and remembrance: The role of neuromodulatory systems. Psychological Science, 1(1), 15–25.
McGaugh, J. L. (2004). The amygdala modulates the consolidation of memories of emotionally arousing experiences. Annual Review of

269
Neuroscience, 27, 1–28.
McGaugh, J. L., Introini-Collison, I. B., Cahill, L. F., Castellano, C., Dalmaz, C., Parent, M. B., & Williams, C. L. (1993).
Neuromodulatory systems and memory storage: Role of the amygdala. Behavioral Brain Research, 58(1–2), 81–90.
McGowan, P. O., Sasaki, A., D’Alessio, A. C., Dymov, S., Labonte, B., Szyf, M., . . . Meaney, M. J. (2009). Epigenetic regulation of the
glucocorticoid receptor in human brain associates with childhood abuse. Nature Neuroscience, 12(3), 342–348.
McGuire, P. K., Paulesu, E., Frackowiak, R. S. J., & Frith, C. D. (1996). Brain activity during stimulus independent thought. NeuroReport,
7(13), 2095–2099.
McKiernan, K. A., D’Angelo, B. R., Kaufman, J. N., & Binder, J. R. (2006). Interrupting the “stream of consciousness”: An fMRI
investigation. NeuroImage, 29(4), 1185–1191. http://doi.org/10.1016/j.neuroimage.2005.09.030
Mckiernan, K. A., Kaufman, J. N., Kucera-Thompson, J., & Binder, J. R. (2003). A parametric manipulation of factors affecting task-
induced deactivation in functional neuroimaging. Journal of Cognitive Neuroscience, 15(3), 394-408. doi:10.1162/089892903321593117
McLaughlin, K. A., Sheridan, M. A., Tibu, F., Fox, N. A., Zeanah, C. H., & Nelson, C. A. (2015). Causal effects of the early caregiving
environment on development of stress response systems in children. Proceedings of the National Academy of Sciences USA, 112(18), 5637–5642.
doi:10.1073/pnas.1423363112
Meaney, M. J., Aitken, D. H., van Berkel, C., Bhatnagar, S., & Sapolsky, R. M. (1988). Effect of neonatal handling on age-related
impairments associated with the hippocampus. Science, 239(4841 Pt. 1), 766–768.
Meaney, M. J., Aitken, D. H., Viau, V., Sharma, S., & Sarrieau, A. (1989). Neonatal handling alters adrenocortical negative feedback
sensitivity and hippocampal type II glucocorticoid receptor binding in the rat. Neuroendocrinology, 50(5), 597–604.
Meaney, M. J., Mitchell, J. B., Aitken, D. H., Bhatnagar, S., Bodnoff, S. R., Iny, L. J., & Sarrieau, A. (1991). The effects of neonatal
handling on the development of the adrenocortical response to stress: Implications for neuropathology and cognitive deficits in later life.
Psychoneuroendocrinology, 16(1–3), 85–103.
Meaney, M. J., & Szyf, M. (2005). Maternal care as a model for experience-dependent chromatin plasticity? Trends in Neurosciences, 28(9),
456–463.
Medendorp, W. P., Goltz, H. C., Crawford, D., & Vilis, T. (2005). Integration of target and effector information in human posterior
parietal cortex for the planning of action. Journal of Neurophysiology, 93(2), 945–962.
Medford, N., & Critchley, H. D. (2010). Conjoint activity of anterior insular and anterior cingulate cortex: Awareness and response. Brain
Structure and Function Brain Struct Funct, 214(5-6), 535-549. doi:10.1007/s00429-010-0265-x
Mehta, M. A., Golembo, N. I., Nosarti, C., Colvert, E., Mota, A., Williams, S. C., . . . Sonuga-Barke, E. J. (2009). Amygdala,
hippocampal and corpus callosum size following severe early institutional deprivation: The English and Romanian Adoptees Study Pilot.
Journal of Child Psychology and Psychiatry, 50(8), 943–951. doi:10.1111/j.1469-7610.2009.02084.x
Menard, J. L., Champagne, D. L., & Meaney, M. J. P. (2004). Variations of maternal care differentially influence “fear” reactivity and
regional patterns of cFos immunoreactivity in response to the shock-probe burying test. Neuroscience, 129(2), 297–308.
Menkes, D. L., Bodnar, P., Ballesteros, R. A., & Swenson, M. R. (1999). Right frontal lobe slow frequency transcranial magnetic
stimulation (SF-r-TMS) is an effective treatment for depression: A case-control pilot study of safety and efficacy. Journal of Neurology,
Neurosurgery, and Psychiatry, 67, 113–115.
Menon, V., & Uddin, L. (2010). Saliency, switching, attention and control: A network model of insula function. Brain Structure and
Function, 214(5–6), 655–667. doi:10.1007/s00429-010-0262-0
Merkl, A., Ammelburg, N., Aust, S., Roepke, S., Reinecker, H., Trahms, L., . . . Sander, T. (2010). Processing of visual stimuli in
borderline personality disorder: A combined behavioural and magnetoencephalographic study. International Journal of Psychophysiology, 78, 257–
264.
Merrin, E. L., & Silberfarb, P. M. (1979). The Capgras phenomenon. Archives of General Psychiatry, 33(8), 965–968.
Messina, I., Sambin, M., Palmieri, A., & Viviani, R. (2013). Neural correlates of psychotherapy in anxiety and depression: A meta-analysis.
PLoS One, 8(9), e74657. doi:10.1371/journal.pone.0074657
Mesulam, M. M. (1981). A cortical network for directed attention and unilateral neglect. Annals of Neurology, 10(4), 309–325.
Meyers, C. A., Berman, S. A., Scheibel, R. S., & Hayman, A. (1992). Case report: Acquired antisocial personality disorder associated with
unilateral left orbital frontal lobe damage. Journal of Psychiatry and Neuroscience, 17(3), 121–125.
Michael, N., & Erfurth, A. (2002). Treatment of bipolar mania with right prefrontal rapid transcranial magnetic stimulation. Journal of
Affective Disorders, 78(3), 253–257.
Michel, G. F., & Moore, C. L. (1995). Developmental psychobiology: An interdisciplinary science. Cambridge, MA: MIT Press.
Milad, M. R., Orr, S. P., Pitman, R. K., & Rauch, S. L. (2005). Context modulation of memory for fear extinction in humans.
Psychophysiology, 42(4), 456–464.
Milad, M. R., Quinn, B. T., Pitman, R. K., Orr, S. P., Fiscal, B., & Rauch, S. L. (2005). Thickness of ventromedial prefrontal cortex in

270
humans is correlated with extinction in memory. Proceedings of the National Academy of Sciences, USA, 102, 10706–10711.
Milad, M. R., & Quirk, G. J. (2002). Neurons in medial prefrontal cortex signal memory for fear extinction. Nature, 420(6911), 70–74.
Milad, M. R., Vidal-Gonzalez, I., & Quirk, G. J. (2004). Electrical stimulation of medial prefrontal cortex reduces conditioned fear in a
temporally specific manner. Behavioral Neuroscience, 118(2), 389–394.
Miller, A. (1981). Prisoners of childhood: The drama of the gifted child and the search for the true self. New York: Basic Books.
Miller, A. (1983). For your own good: Hidden cruelty in child-rearing and the roots of violence. New York: Farrar, Straus, and Giroux.
Miller, H., Alvarez, V., & Miller, A. (1990). The psychopathology and psychoanalytic psychotherapy of compulsive caretaking. Unpublished
manuscript.
Minagawa-Kawai, Y., Matsuoka, S., Dan, I., Naoi, N., Nakamura, K., & Kojima, S. (2008). Prefrontal activation associated with social
attachment: Facial emotion recognition in mothers and infants. Cerebral Cortex, 19(2), 284–292.
Ming, G., & Song, H. (2011). Adult neurogenesis in the mammalian brain: Significant answers and significant questions. Neuron, 70(4),
687–702. doi:10.1016/j.neuron.2011.05.001
Minnis, H., Macmillan, S., Pritchett, R., Young, D., Wallace, B., Butcher, J., . . . Gillberg, C. (2013). Prevalence of reactive attachment
disorder in a deprived population. British Journal of Psychiatry, 202(5), 342–346. doi:10.1192/bjp.bp.112.114074
Minzenberg, M. J., Fan, R., New, A. S., Tang, C. Y., & Siever, L. J. (2007). Frontolimbic dysfunction in response to facial emotion in
borderline personality disorder: An event-related fMRI study. Psychiatry Research, 155(3), 231–243.
Minzenberg, M. J., Fan, J., New, A. S., Tang, C. Y., & Siever, L. J. (2008). Frontolimbic structural changes in borderline personality
disorder. Journal of Psychiatric Research, 42(9), 727–733.
Minzenberg, M. J., Poole, J. H., & Vinogradov, S. (2006). Social-emotion recognition in borderline personality disorder. Comprehensive
Psychiatry, 47, 468–474.
Mirescu, C., Peters, J. D., & Gould, E. (2004). Early life experience alters response of adult neurogenesis to stress. Nature Neuroscience, 7(8),
841–846.
Mitchell, J. P., Banaji, M. R., & Macrae, C. N. (2005). The link between social cognition and self-referential thought in the medial
prefrontal cortex. Journal of Cognitive Neuroscience, 17(8), 1306–1315.
Mitchell, J. P., Macrae, C. N., & Banaji, M. R. (2006). Dissociable medial prefrontal contributions to judgments of similar and dissimilar
others. Neuron, 50(4), 655–663.
Mitchell, R. L. C., & Crow, T. J. (2005). Right hemispheric language functions and schizophrenia: The forgotten hemisphere. Brain, 128,
963–978.
Mitra, R., & Sapolsky, R. M. (2008). Acute corticosterone treatment is sufficient to induce anxiety and amygdaloid dendritic hypertrophy.
Proceedings of the National Academy of Sciences, USA, 105(14), 5573–5578.
Modney, B. K., & Hatton, G. I. (1994). Maternal behaviors: Evidence that they feed back to alter brain morphology and function. Acta
Paediatrica, 83(s397), 29–32.
Modney, B., Yang, Q., & Hatton, G. (1990). Activation of excitatory amino acid inputs to supraoptic neurons. II. Increased dye-coupling in
maternally behaving virgin rats. Brain Research, 513(2), 270–273.
Moerman, D. E., & Jonas, W. B. (2002). Deconstructing the placebo effect and finding the meaning response. Annals of Internal Medicine,
136(6), 471–476.
Molko, N., Cachia, A., Riviere, D., Mangin, J. F., Brauandet, M., Bihan, D. L., . . . Dehaene, S. (2003). Functional and structural
alteration of the intrapietal sulcus in a developmental dyscalculia of genetic origin. Neuron, 40, 847–858.
Moll, J., Krueger, F., Zahn, R., Pardini, M., Oliviera-Souza, R., & Grafman, J. (2006). Human fronto-mesolimbic networks guide
decisions about charitable donation. PLoS One, 103(42), 15623–15628.
Moll, J., Oliveira-Souza, R. D., Garrido, G. J., Bramati, I. E., Caparelli-Daquer, E. M., Paiva, M. L., . . . Grafman, J. (2007). The self as a
moral agent: Linking the neural bases of social agency and moral sensitivity. Social Neuroscience, 2(3-4), 336-352.
doi:10.1080/17470910701392024
Monfils, M., Cowansage, K. K., & LeDoux, J. E. (2007). Brain-derived neurotrophic factor: Linking fear learning to memory
consolidation. Molecular Pharmacology, 72(2), 235–237.
Moore, G. T. (1986). Effects of the spatial definition of behavior settings on children’s behavior: A quasi-experimental field study. Journal of
Environmental Psychology, 6(3), 205–231.
Morcom, A. M., & Fletcher, P. C. (2007). Cognitive neuroscience: The case for design rather than default. NeuroImage, 37, 1097–1099.
Morcom, A. M., & Fletcher, P. C. (2007). Does the brain have a baseline? Why we should be resisting a rest. NeuroImage, 37(4), 1073-
1082. doi:10.1016/j.neuroimage.2006.09.013
Morgan, C. A., Wang, S., Southwick, S. M., Rasmusson, A., Hazlett, G., Hauger, R. L., & Charney, D. S. (2000). Plasma neuropeptide-Y
concentrations in humans exposed to military survival training. Biological Psychiatry, 47(10), 902–909.

271
Morgan, M. A., Romanski, L. M., & LeDoux, J. E. (1993). Extinction of emotional learning: Contribution of medial prefrontal cortex.
Neuroscience Letters, 163(1), 109–113.
Moriceau, S., & Sullivan, R. M. (2004). Corticosterone influences on mammalian neonatal sensitive-period learning. Behavioral
Neuroscience, 118(2), 274–281.
Morishima, Y., Schunk, D., Bruhin, A., Ruff, C. C., & Fehr, E. (2012). Linking brain structure and activation in temporoparietal junction
to explain the neurobiology of human altruism. Neuron, 75, 73–79. doi:10.1016/j.neuron.2012.05.021
Morley-Fletcher, S., Rea, M., Maccari, S., & Laviola, G. (2003). Environmental enrichment during adolescence reverses the effects of
prenatal stress on play behaviour and HPA axis reactivity in rats. European Journal of Neuroscience, 18(12), 3367–3374.
Morrison, J. H., & Hof, P. R. (2003). Changes in cortical circuits during aging. Clinical Neuroscience Research, 2(5–6), 294–304.
Mountcastle, V. B. (1995). The parietal system and some higher brain functions. Cerebral Cortex, 5(5), 377–390.
Murray, E., O’doherty, J., & Schoenbaum, G. (2007). What we know and do not know about the functions of the orbitofrontal cortex after
20 years of cross-species studies. Journal of Neuroscience, 27(31), 8166–8169. doi:10.1523/jneurosci.1556-07.2007
Myers, J. J., & Sperry, R. W. (1985). Interhemispheric communication after section of the forebrain commissures. Cortex, 21(2), 249–260.
Myers, K. M., & Davis, M. (2007). Mechanisms of fear extinction. Molecular Psychiatry, 12(2), 120–150.
Myers, W. A., Churchill, J. D., Muja, N., & Garraghty, P. E. (2000). Role of NMDA receptors in adult primates cortical somatosensory
plasticity. Journal of Comparative Neurology, 418, 373–382.
Nachev, P., Mah, Y. H., & Husain, M. (2009). Functional neuroanatomy: The locus of human intelligence. Current Biology, 19(10), R418–
R420.
Nachmias, M., Gunnar, M. R., Mangelsdorf, S., Parritz, R. H., & Buss, K. (1996). Behavioral inhibition and stress reactivity: The
moderating role of attachment security. Child Development, 67(2), 508–522.
Nagahama, Y., Okada, T., Katsumi, Y., Hayashi, T., Yamauchi, H., Oyanagi, C., . . . Shibasaki, H. (2001). Dissociable mechanisms of
attentional control within the human prefrontal cortex. Cerebral Cortex, 11(1), 85–92.
Nair, H., Berndt, J., Barrett, D., & Gonzalez-Lima, F. (2001). Maturation of extinction behavior in infant rats: Large-scale regional
interactions with medial prefrontal cortex, orbitofrontal cortex, and anterior cingulate cortex. Journal of Neuroscience, 21(12), 4400–4407.
Najib, A., Lorberbaum, J. P., Kose, S., Bohning, D. E., & George, M. S. (2004). Regional brain activity in women grieving a romantic
relationship breakup. American Journal of Psychiatry, 161, 2245–2256.
Nakamarua, M., & Iwasab, Y. (2006). The coevolution of altruism and punishment: Role of the selfish punisher. Journal of Theoretical
Biology, 240, 475–488.
Nakao, T., Nakagawa, A., Yoshiura, T., Nakatani, E., Nabeyama, M., Yoshizato, C., . . . Kanba, S. (2005). Brain activation of patients with
obsessive-compulsive disorder during neuropsychological and symptom provocation tasks before and after symptom improvement: A functional
magnetic resonance imaging study. Biological Psychiatry, 57(8), 901–910. doi:10.1016/j.biopsych.2004.12.039
Nakatani, E., Nakgawa, A., Ohara, Y., Goto, S., Uozumi, N., Iwakiri, M., . . . Yamagami, T. (2003). Effects of behavior therapy on
regional cerebral blood flow in obsessive- compulsive disorder. Psychiatry Resources, 124(2), 113–120.
Nasrallah, H. A. (1985). The unintegrated right cerebral hemispheric consciousness as alien intruder: A possible mechanism for
Schneiderian delusions in schizophrenia. Comprehensive Psychiatry, 26(3), 273–282.
Nauta, W. J. H. (1971). The problem of the frontal lobe: A reinterpretation. Journal of Psychiatric Research, 8(3), 167–187.
Navalta, C. P., Polcari, A., Webster, D. M., Boghossian, A., & Teicher, M. H. (2004). Effects of childhood sexual abuse on
neuropsychological cognitive function in college women. Journal of Neuropsychiatry and Clinical Neuroscience, 18(1), 45–53.
Nebes, R. D. (1971). Superiority of the minor hemisphere in commissurotomized man for the perception of part-whole relationships.
Cortex, 7, 333–349.
Nedergaard, M., Ransom, B., & Goldman, S. A. (2003). New roles for astrocytes: Redefining the functional architecture of the brain.
Trends in Neurosciences, 26(10), 523–530.
Nelson, C. A., & Carver, L. J. (1998). The effects of stress and trauma on brain and memory: A view from developmental cognitive
neuroscience. Development and Psychopathology, 10(4), 793–809.
Nelson, E. E., & Panksepp, J. (1998). Brain substrates of infant-mother attachment: Contributions of opioids, oxytocin, and
norepinephrine. Neuroscience and Biobehavioral Reviews, 22(3), 437–452.
Nelson, E. M., & Spieker, S. J. (2013). Intervention effects on morning and stimulated cortisol responses among toddlers in foster care.
Infant Mental Health Journal, 34(3), 211-221. doi:10.1002/imhj.21382
Nelson, K. (1993). The psychological and social origins of autobiographical memory. Psychological Science, 4(1), 7–14.
Nelson, M. D., Saykin, A. J., Flashman, L. A., & Riordan, H. J. (1998). Hippocampal volume reduction in schizophrenia as assessed by
magnetic resonance imaging: A meta-analytic study. Archives of General Psychiatry, 55(5), 433–440.
Nemeroff, C., Heim, C. M., Thase, M. E., Klein, D. N., Rush, A., & Schatzberg, A. (2003). Differential responses to psychotherapy versus

272
pharmacotherapy in patients with chronic forms of major depression and childhood trauma. Proceedings of the National Academy of Sciences, USA,
25(35), 14293–14296.
Nesse, R. M., & Lloyd, A. T. (1992). The evolution of psychodynamic mechanisms. In J. H. Barkow, L. Cosmides, & J. Tooby (Eds.), The
adapted mind: Evolutionary psychology and the generation of culture (pp. 601–626). New York: Oxford University Press.
Neugebauer, V., Li, W., Bird, G., & Han, J. (2004). The amygdala and persistent pain. Neuroscientist, 10(3), 221–234.
Neumann, I. D. (2008). Brain oxytocin: A key regulator of emotional and social behaviors in both females and males. Journal of
Neuroendocrinology, 20(6), 858–865.
New, A. S., & Stanley, B. (2010). An opioid deficit in borderline personality disorder: Self-cutting, substance abuse, and social dysfunction.
American Journal of Psychiatry, 167(8), 882–885.
Newberg, A., Alavi, A., Baime, M., Pourdehnad, M., Santanna, J., & Aquili, E. (2001). The measurement of cerebral blood flow during the
complex cognitive task of meditation: A preliminary SPECT study. Psychiatric Research: Neuroimaging Section, 106, 113–122.
Newcomer, J. W., Craft, S., Hershey, T., Askins, K., & Bardgett, M. E. (1994). Glucocorticoid-induced impairment in declarative memory
performance in adult humans. Journal of Neuroscience, 14(4), 2047–2053.
Newcomer, J. W., Selke, G., Melson, A. K., Hershey, T., Craft, S., Richards, K., & Alderson, A. L. (1999). Decreased memory
performance in healthy humans induced by stress-level cortisol treatment. Archives of General Psychiatry, 56(6), 527–533.
Newman, D. (1982). Perspective-taking versus context in understanding lies. Quarterly Newsletter of the Laboratory of Comparative Human
Cognition, 4, 26–29.
Newman-Norlund, R. D., Ganesh, S., Schie, H. T., Bruijn, E. R., & Bekkering, H. (2008). Self-identification and empathy modulate
error-related brain activity during the observation of penalty shots between friend and foe. Social Cognitive and Affective Neuroscience, 4(1), 10-
22. doi:10.1093/scan/nsn028
Newman, S. D., Carpenter, P. A., Varma, S., & Just, M. A. (2003). Frontal and parietal participation in problem solving in the Tower of
London: fMRI and computational modeling of planning and high-level perception. Neuropsychologia, 41(12), 1668–1682.
Nichols, K., & Champness, B. (1971). Eye gaze and the GRS. Journal of Experimental Social Psychology, 7, 623–626.
Nielsen, M., & Dissanayake, C. (2004). Pretend play, mirror self-recognition and imitation: A longitudinal investigation through the
second year. Infant Behavior and Development, 27(3), 342-365. doi:10.1016/j.infbeh.2003.12.006
Nielson, K. A., Yee, D., & Erickson, K. I. (2005). Memory enhancement by a semantically unrelated emotional arousal source induced after
learning. Neurobiology of Learning and Memory, 84(1), 49–56.
Nietlisbach, G., & Maercker, A. (2009a). Effects of social exclusion in trauma survivors with posttraumatic stress disorder. Psychological
Trauma: Theory, Research, Practice, and Policy, 1(4), 323–331. doi:10.1037/a0017832
Nietlisbach, G., & Maercker, A. (2009b). Social cognition and interpersonal impairments in trauma survivors with PTSD. Journal of
Aggression, Maltreatment and Trauma, 18(4), 382–402. doi:10.1080/10926770902881489
Nikolaenko, N. N., Egorov, A. Y., & Freiman, E. A. (1997). Representational activity of the right and left hemispheres of the brain.
Behavioral Neurology, 10, 49–59.
Nilsson, L., Mohammed, A. K. H., Henriksson, B. G., Folkesson, R., Winblad, B., & Bergstrom, L. (1993). Environmental influence on
somatostatin levels and gene expression in the rat brain. Brain Research, 628(1–2), 93–98.
Nimchinsky, E. A., Gilissen, E., Allman, J. M., Perl, D. P., Erwin, J. M., & Hof, P. R. (1999). A neuronal morphologic type unique to
humans and great apes. Proceedings of the National Academy of Sciences, USA, 96(9), 5268–5273.
Nimchinsky, E. A., Vogt, B. A., Morrison, J. H., & Hof, P. R. (1995). Spindle neurons of the human anterior cingulate cortex. Journal of
Comparative Neurology, 355(1), 27–37.
Nisenbaum, L. K., Zigmond, M. J., Sved, A. F., & Abercrombie, E. D. (1991). Prior exposure to chronic stress results in enhanced
synthesis and release of hippocampal norepinephrine in response to novel stressors. Journal of Neuroscience, 11, 1478–1484.
Nishitani, N., & Hari, R. (2000). Temporal dynamics of cortical representation for action. Proceedings of the National Academy of Sciences,
USA, 97(2), 913–918.
Nishitani, N., Schürmann, M., Amunts, K., & Hari, R. (2004). Broca’s regions: From action to language. Physiology, 20, 60–69.
Nitschke, J. B., Nelson, E. E., Rusch, B. D., Fox, A. S., Oakes, T. R., & Davidson, R. J. (2004). Orbitofrontal cortex tracks positive mood
in mothers viewing pictures of their newborn infants. NeuroImage, 21(2), 583–592.
Nobre, A. C., Coull, J. T., Frith, C. D., & Mesulam, M. M. (1999). Orbitofrontal cortex is activated during breaches of expectation in tasks
of visual attention. Nature Neuroscience, 2(1), 11–12.
Nolte, J. (2008). The human brain: An introduction to its functional anatomy (6th ed.). St. Louis, MO: Mosby.
Nomura, M., Iidaka, T., Kakehi, K., Tsukiura, T., Hasegawa, T., Maeda, Y., & Matsue, Y. (2003). Frontal lobe networks for effective
processing of ambiguously expressed emotions in humans. Neuroscience Letters, 348(2), 113–116.
Northoff, G., & Bermpohl, F. (2004). Cortical midline structures and the self. Trends in Cognitive Sciences, 8(3), 102-107.

273
doi:10.1016/j.tics.2004.01.004
Northoff, G., Heinzel, A., Bermpohl, F., Niese, R., Pfennig, A., Pascual-Leone, A., & Schlaug, G. (2004). Reciprocal modulation and
attenuation in the prefrontal cortex: An fMRI study on emotional-cognitive interaction. Human Brain Mapping, 21(3), 202–212.
Northoff, G., Heinzel, A., Greck, M. D., Bermpohl, F., Dobrowolny, H., & Panksepp, J. (2006). Self-referential processing in our brain—
A meta-analysis of imaging studies on the self. NeuroImage, 31(1), 440-457. doi:10.1016/j.neuroimage.2005.12.002
Nottebohm, F. (1981). A brain for all seasons: Cyclical anatomical changes in song-control nuclei of the canary brain. Science, 214(4527),
1368–1370.
Oatley, K. (1992). Integrative action of narrative. In D. J. Stein & J. E. Young (Eds.), Cognitive science and clinical disorders (pp. 151–172).
New York: Academic Press.
Oberheim, N. A., Wang, X., Goldman, S., & Nedergaard, M. (2006). Astrocytic complexity distinguishes the human brain. Trends in
Neurosciences, 29(10), 547–553.
O’Brien, J. T. (1997). The “glucocorticoid cascade” hypothesis in man. British Journal of Psychiatry, 170, 199–201.
Ochs, E., & Capps, L. (2001). Living narrative: Creating lives in everyday storytelling. Cambridge, MA: Harvard University Press.
Ochsner, K. N., Beer, J. S., Robertson, E. R., Cooper, J. C., Gabrieli, J. D. E., Kihlstrom, J. F., & D’Esposito, M. (2005). The neural
correlates of direct and reflected self-knowledge. NeuroImage, 28, 797–814.
Ochsner, K. N., Bunge, S. A., Gross, J. J., & Gabrieli, J. D. E. (2002). Rethinking feelings: An fMRI study of the cognitive regulation of
emotion. Journal of Cognitive Neuroscience, 14(8), 1215–1229.
Ochsner, K. N., & Gross, J. J. (2008). Cognitive emotion regulation. Insights from social cognitive and affective neuroscience. Current
Directions in Psychological Science, 17(2), 153–158.
Ochsner, K. N., Ray, R. D., Cooper, J. C., Robertson, E. R., Chopra, S., Gabrieli, J. D., & Gross, J. J. (2004). For better or for worse:
Neural systems supporting the cognitive down- and up-regulation of negative emotion. NeuroImage, 23(2), 483–499.
O’Doherty, J. (2004). Reward representation and reward-related learning in the human brain: Insights from neuroimaging. Current Opinion
in Neurobiology, 14, 769–776.
O’Doherty, J. P., Deichmann, R., Critchley, H. D., & Dolan, R. J. (2002). Neural responses during anticipation of a primary taste reward.
Neuron, 33(5), 815–826.
O’Doherty, J., Kringelbach, M. L., Rolls, E. T., Hornak, J., & Andrews, C. (2001). Abstract reward and punishment representations in the
human orbitofrontal cortex. Nature Neuroscience, 4(1), 95–102.
O’Donnell, D., Larocque, S., Seckl, J. R., & Meaney, M. J. (1994). Postnatal handling alters glucocorticoid, but not mineralocorticoid
messenger RNA expression in the hippocampus of adult rats. Brain Research. Molecular Brain Research, 26(1–2), 242–248.
Ohman, A., Carlsson, K., Lundqvist, D., & Ingvar, M. (2007). On the unconscious subcortical origin of human fear. Physiology and
Behavior, 92(1–2), 180–185.
O’Keefe, J., & Nadel, L. (1978). The hippocampus as a cognitive map. Oxford: Clarendon.
Olesen, P. J., Nagy, Z., Westerberg, H., & Klingberg, T. (2003). Combined analysis of DTI and fMRI data reveals a joint maturation of
white and grey matter in a fronto-parietal network. Cognitive Brain Research, 18(1), 48–57. doi:10.1016/j.cogbrainres.2003.09.003
Olsson, A., Ebert, J. P., Banaji, M. R., & Phelps, E. A. (2005). The role of social groups in the persistence of learned fear. Science,
309(5735), 785–787.
Olsson, A., & Phelps, E. A. (2007). Social learning of fear. Nature Neuroscience, 10(9), 1095–1102.
O’Neill, J., Gorbis, E., Feusner, J. D., Yip, J. C., Chang, S., Maidment, K. M., . . . Saxena, S. (2013). Effects of intensive cognitive-
behavioral therapy on cingulate neurochemistry in obsessive-compulsive disorder. Journal of Psychiatric Research, 47(4), 494–504.
doi:10.1016/j.jpsychires.2012.11.010
Öngür, D., & Price, J. L. (2000). The organization of networks within the orbital and medial prefrontal cortex of rats, monkeys, and
humans. Cerebral Cortex, 10(3), 206–219.
Ono, T., Nishijo, H., & Uwano, T. (1995). Amygdala role in conditioned associative learning. Progress in Neurobiology, 46(4), 401–422.
Orban, G., Claeys, K., Nelissen, K., Smans, R., Sunaert, S., Todd, J., . . . Vanduffel, W. (2006). Mapping the parietal cortex of human and
non-human primates. Neuropsychologia, 44(13), 2647–2667.
Orban, G. A., Sunaert, S., Todd, J. T., Van Hecke, P., & Marchal, G. (1999). Human cortical regions involved in extracting depth from
motion. Neuron, 24(4), 929–940.
Orlinsky, D. E., & Howard, K. J. (1986). Process and outcome in psychotherapy. In S. L. Garfield & A. E. Bergin (Eds.), Handbook of
psychotherapy and behavior change (pp. 311–381). New York: Wiley.
Ornitz, E. M., & Pynoos, R. S. (1989). Startle modulation in children with posttraumatic stress disorder. American Journal of Psychiatry,
146(7), 866–870.
Ostby, Y., Walhovd, K. B., Tamnes, C. K., Grydeland, H., Westlye, L. T., & Fjell, A. M. (2012). Mental time travel and default-mode

274
network functional connectivity in the developing brain. Proceedings of the National Academy of Sciences, 109(42), 16800-16804.
doi:10.1073/pnas.1210627109
Ouspensky, P. D. (1954). The psychology of man’s possible evolution. New York: Alfred Knopf.
Ovtscharoff, W., Jr., & Braun, K. (2001). Maternal separation and social isolation modulate the postnatal development of synaptic
composition in the infralimbic cortex of Octodon degus. Neuroscience, 104(1), 33–40.
Ovtscharoff, W., Helmeke, C., & Braun, K. (2006). Lack of paternal care affects synaptic development in the anterior cingulate cortex.
Brain Research, 1116(1), 58–63.
Pagnoni, G., Zink, C. F., Montague, R., & Berns, G. S. (2002). Activity in human ventral striatum locked to errors of reward prediction.
Nature Neuroscience, 5(2), 97–98.
Palaniyappan, L. (2012). Does the salience network play a cardinal role in psychosis? An emerging hypothesis of insular dysfunction. Journal
of Psychiatry and Neuroscience, 37(1), 17–27. doi:10.1503/jpn.100176
Panksepp, J. (1998). Affective neuroscience: The foundation of human and animal emotions. New York: Oxford University Press.
Panksepp, J., Nelson, E., & Siviy, S. (1994). Brain opioids and mother-infant social motivation. Acta Paediatrica, 83(397), 40–46.
Paquette, V., Lévesque, J., Mensour, B., Leroux, J. M., Beaudoin, G., Bourgouin, P., & Beauregard, M. (2003). “Change the mind and you
change the brain”: Effects of cognitive-behavioral therapy on the neural correlates of spider phobia. NeuroImage, 18(2), 401–409.
Pardo, J. V., Pardo, P. J., & Raichle, M. E. (1993). Neural correlates of self-induced dysphoria. American Journal of Psychiatry, 150(5), 713-
719. doi:10.1176/ajp.150.5.713
Parente, R., & Herrmann, D. (1996). Retraining cognition. Gaithersburg, MD: Aspen.
Pariente, J., White, P., Frackowiak, R. S. J., & Lewith, G. (2005). Expectancy and belief modulate the neuronal substrates of pain treated
by acupuncture. NeuroImage, 25(4), 1161–1167.
Paris, J., Zelkowitz, P., Guzder, J., Joseph, S., & Feldman, R. (1999). Neuropsychological factors associated with borderline pathology in
children. Journal of the Academy of Child and Adolescent Psychiatry, 38(6), 770–774.
Park, C. R., Campbell, A. M., Woodson, J. C., Smith, T. P., Fleshner, M., & Diamond, D. M. (2006). Permissive influence of stress in the
expression of a u-shaped relationship between serum corticosterone levels and spatial memory errors in rats. Dose-Response, 4(1), 55–74.
Park, J., Kim, Y., Chang, W. H., Park, C., Shin, Y., Kim, S. T., & Pascual-Leone, A. (2014). Significance of longitudinal changes in the
default-mode network for cognitive recovery after stroke. European Journal of Neuroscience, 40(4), 2715-2722. doi:10.1111/ejn.12640
Park, M. K., Hoang, T. A., Belluzzi, J. D., & Leslie, F. M. (2003). Gender specific effect of neonatal handling on stress reactivity of
adolescent rats. Journal of Neuroendocrinology, 15(3), 289–295.
Parsons, C. G., Stöffler, A., & Danysz, W. (2007). Memantine: A NMDA receptor antagonist that improves memory by restoration of
homeostasis in the glutamatergic system—too little activation is bad, too much is even worse. Neuropharmacology, 53(6), 699–723.
Pascual-Leone, A., Rubio, B., Pallardo, F., & Catala, M. D. (1996). Rapid-rate transcranial magnetic stimulation of left dorsolateral
prefrontal cortex in drug-resistant depression. Lancet, 348(9022), 233–237.
Patel, R., Spreng, R. N., Shin, L. M., & Girard, T. A. (2012). Neurocircuitry models of posttraumatic stress disorder and beyond: A meta-
analysis of functional neuroimaging studies. Neuroscience and Biobehavioral Reviews, 36(9), 2130–2142. doi:10.1016/j.neubiorev.2012.06.003
Patihis, L., Frenda, S. J., LePort, A. K., Petersen, N., Nichols, R. M., Stark, C. E., . . . Loftus, E. F. (2013). False memories in highly
superior autobiographical memory individuals. Proceedings of the National Academy of Sciences, 110(52), 20947–20952.
doi:10.1073/pnas.1314373110
Paus, T., Petrides, M., Evans, A. C., & Meyer, E. (1993). Role of the human anterior cingulate cortex in the control of oculomotor,
manual, and speech responses: A positron emission tomography study. Journal of Neurophysiology, 70(2), 453–469.
Pavlides, C., Watanabe, Y., Magarinos, A. M., & McEwen, B. S. (1995). Opposing roles of the type I and type II adrenal steroid receptors
in hippocampal long-term potentiation. Neuroscience, 68(2), 387–394.
Pawluski, J., & Galea, L. (2006). Hippocampal morphology is differentially affected by reproductive experience in the mother. Journal of
Neurobiology, 66(1), 71–81.
Paz-Alonso, P. M., & Goodman, G. S. (2008). Trauma and memory: Effects of post-event misinformation, retrieval order, and retention
interval. Memory, 16(1), 58–75.
Pears, K. C., Bruce, J., Fisher, P. A., & Kim, H. K. (2009). Indiscriminate friendliness in maltreated foster children. Child Maltreatment,
15(1), 64–75. doi:10.1177/1077559509337891
Peers, P. V., Ludwig, C. J. H., Rorden, C., Cusack, R., Bonfiglioli, C., Bundesen, C., . . . Duncan, J. (2005). Attentional functions of
parietal and frontal cortex. Cerebral Cortex, 15(10), 1469–1484.
Pelphrey, K. A., Singerman, J. D., Allison, T., & McCarthy, G. (2003). Brain activation evoked by perception of gaze shifts: The influence
of context. Neuropsychologia, 41(2), 156–170.
Penades, R., Boget, T., Lomena, F., Mateos, J., Catalan, R., & Gasto, C. (2002). Could the hypofrontality pattern in schizophrenia be

275
modified through neuropsychological rehabilitation? Acta Psychiatrica Scandinavica, 105(3), 202–208.
Pencea, V., Bingaman, K. D., Wiegland, S. J., & Luskin, M. B. (2001). Infusion of brain-derived neurotrophic factor into the lateral
ventricle of the adult rat leads to new neurons in the parenchyma of the striatum, septum, thalamus, and hypothalamus. Journal of Neuroscience,
21(17), 6706–6717.
Penfield, W., & Perot, P. (1963). The brain’s record of auditory and visual experience. Brain, 86, 595–696.
Pennebaker, J. W. (1997). Writing about emotional experiences as a therapeutic process. Psychological Science, 8(3), 162–166.
Pennebaker, J. W., & Beall, S. K. (1986). Confronting a traumatic event: Toward an understanding of inhibition and disease. Journal of
Abnormal Psychology, 95(3), 274–281.
Pennebaker, J. W., Kiecolt-Glaser, J. K., & Glaser, R. (1988). Disclosure of traumas and immune function: Health implications for
psychotherapy. Journal of Consulting and Clinical Psychology, 56(2), 239–245.
Pepperberg, I. M. (2008). Alex and me: How a scientist and a parrot uncovered a hidden world of animal intelligence—and formed a deep bond in
the process. New York: HarperCollins.
Pérez-Jaranay, J. M., & Vives, F. (1991). Electrophysiological study of the response of medial prefrontal cortex neurons to stimulation of the
basolateral nucleus of the amygdala in the rat. Brain Research, 564(1), 97–101.
Perlman, G., Simmons, A. N., Wu, J., Hahn, K. S., Tapert, S. F., Max, J. E., . . . Yang, T. T. (2012). Amygdala response and functional
connectivity during emotion regulation: A study of 14 depressed adolescents. Journal of Affective Disorders, 139(1), 75-84.
doi:10.1016/j.jad.2012.01.044
Perls, F., Hefferline, R., & Goodman, P. (1951). Gestalt therapy: Excitement and growth in human personality. New York: Dell.
Perls, T. T., Alpert, L., & Fretts, R. C. (1997). Middle-aged mothers live longer. Nature, 389(6647), 133.
Perrett, D. I., Rolls, E. T., & Caan, W. (1982). Visual neurons responsive to faces in the monkey temporal cortex. Experimental Brain
Research, 47(3), 329–342.
Perrett, D. I., Smith, A. J., Potter, D. D., Mistlin, A. J., Head, A. D., Milner, A. D., & Jeeves, M. A. (1984). Neurons responsive to faces
in the temporal cortex: Studies of functional organization, sensitivity to identity and relation to perception. Human Neurobiology, 3(4), 197–208.
Perry, B. D., Pollard, R. A., Blakley, T. I., Baker, W. L., & Vigilante, D. (1995). Childhood trauma, the neurobiology of adaptation, and
“use dependent” development of the brain: How “states” become “traits.” Infant Mental Health Journal, 16(4), 271–291.
Persinger, M. A., & Makarec, K. (1991). Greater right hemisphericity is associated with lower self-esteem in adults. Perceptual and Motor
Skills, 73(3 Pt. 2), 1244–1246.
Pessoa, L. (2008). On the relationship between emotion and cognition. Nature Reviews Neuroscience, 9(2), 148–158.
Peterson, C. (2012). Children’s autobiographical memories across the years: Forensic implications of childhood amnesia and eyewitness
memory for stressful events. Developmental Review, 32(3), 287-306. doi:10.1016/j.dr.2012.06.002
Petrides, M., Alivisatos, B., & Frey, S. (2002). Differential activation of the human orbital, mid-ventrolateral, and mid-dorsolateral
prefrontal cortex during the processing of visual stimuli. Proceedings of the National Academy of Sciences, USA, 99(8), 5649–5654.
Petrie, K. J., Booth, R. J., & Pennebaker, J. W. (1998). The immunological effects of thought suppression. Journal of Personality and Social
Psychology, 75(5), 1264–1272.
Petrie, K. J., Booth, R. J., Pennebaker, J. W., Davison, K. P., & Thomas, M. G. (1995). Disclosure of trauma and immune response to a
hepatitis B vaccination program. Journal of Consulting and Clinical Psychology, 63(5), 787–792.
Petrovic, P., Kelso, E., Petersson, K. M., & Ingvar, M. (2002). Placebo and opioid analgesia—imaging a shared neuronal network. Science,
295(5560), 1737–1740.
Petty, F., Chae, Y., Kramer, G., Jordan, S., & Wilson, L. (1994). Learned helplessness sensitizes hippocampal norepinephrine to mild
stress. Biological Psychiatry, 35, 903–908.
Pezawas, L., Meyer-Lindenberg, A., Drabant, E. M., Verchinski, B. A., Munoz, K. E., Kolachana, B. S., . . . Hariri, A. R. (2005). 5-
HTTLPR polymorphism impacts human cingulate-amygdala interactions: A genetic susceptibility mechanism for depression. Nature
Neuroscience, 8(6), 828–834.
Pfeffer, C. R., Martins, P., Mann, J., Sunkenberg, M., Ice, A., Damore, J. P., Jr., . . . Jiang, H. (1997). Child survivors of suicide:
Psychological characteristics. Journal of the American Academy of Childhood and Adolescent Psychiatry, 36, 65–74.
Pfrieger, F. W., & Barres, B. A. (1996). New views on synapse-glia interactions. Current Opinions in Neurobiology, 6(5), 615–621.
Pham, K., Nacher, J., Hof, P. R., & McEwen, B. (2003). Repeated restraint stress suppresses neurogenesis and induces biphasic PSA-
NCAM expression in the adult rat dentate gyrus. European Journal of Neuroscience, 17(4), 879–886.
Pham, T. M., Soderstrom, S., Henriksson, B. G., & Mohammed, A. H. (1997). Effects of neonatal stimulation on later cognitive function
and hippocampal nerve growth factor. Behavioral Brain Research, 86(1), 113–120.
Phan, K. L., Britton, J. C., Taylor, S. F., Fig, L. M., & Liberzon, I. (2006). Corticolimbic blood flow during nontraumatic emotional
processing in post-traumatic stress disorder. Archives of General Psychiatry, 63, 184–192.

276
Phan, K. L., Fitzegerald, D., Nathan, P., Moore, G., Uhde, T., & Tancer, M. (2005). Neural substrates for voluntary suppression of
negative affect: A functional magnetic resonance imaging study. Biological Psychiatry, 57(3), 210–219.
Phan, K. L., Wager, T., Taylor, S. F., & Liberzon, I. (2002). Functional neuroanatomy of emotion: A meta-analysis of emotion activation
studies in PET and fMRI. NeuroImage, 16(2), 331–348.
Phelps, E. A. (2006). Emotion and cognition: Insights from studies of the human amygdala. Annual Review of Psychology, 57, 27–53.
Phelps, E. A., & Anderson, A. K. (1997). Emotional memory: What does the amygdala do? Current Biology, 7(5), R311–R314.
Phelps, E. A., Delgado, M. R., Nearing, K. I., & LeDoux, J. E. (2004). Extinction learning in humans: Role of the amygdala and vmPFC.
Neuron, 43(6), 897–905.
Phelps, E., Ling, S., & Carrasco, M. (2006). Emotion facilitates perception and potentiates the perceptual benefits of attention. Psychological
Science, 17(4), 292–299.
Phillips, D. P., Ruth, T. E., & Wagner, L. M. (1993). Psychology and survival. Lancet, 342(8880), 1142–1145.
Pia, L., Neppi-Modona, M., Ricci, R., & Berti, A. (2004). Special issue: The anatomy of anosognosia for hemiplegia: A meta-analysis.
Cortex, 40(2), 367–377.
Piazza, M., Izard, V., Pinel, P., Le Bihan, D., & Dehaene, S. (2004). Tuning curves for approximate numerosity in the human intraparietal
sulcus. Neuron, 44(3), 547–555.
Pilowsky, D. J., Wickramaratne, P., Talati, A., Tang, M., Hughes, C. W., Garber, J., . . . Weissman, M. M. (2008). Children of depressed
mothers 1 year after the initiation of maternal treatment: Findings from the STAR*D-Child study. American Journal of Psychiatry, 165(9), 1136–
1147.
Pissiota, A., Frans, O., Fernandez, M., von Knorring, L., Fischer, H., & Fredrikson, M. (2002). Neurofunctional correlates of
posttraumatic stress disorder: A PET symptom provocation study. European Archives of Psychiatry and Clinical Neuroscience, 252(2), 68–75.
Pitman, R. K., Orr, S. P., van der Kolk, B. A., Greenberg, M. S., Meyerhoff, J. L., & Mougey, E. H. (1990). Analgesia: A new dependent
variable for the biological study of posttraumatic stress disorder. In M. E. Wolf & A. D. Mosnaim (Eds.), Posttraumatic stress disorder: Etiology,
phenomenology, and treatment (pp. 140–147). Washington, DC: American Psychiatric Press.
Pittenger, C., & Duman, R. S. (2008). Stress, depression, and neuroplasticity: A convergence of mechanisms. Neuropsychopharmacology,
33(1), 88–109.
Pizzagalli, D., Pascual-Marqui, R. D., Nitschke, J. B., Oakes, T. R., Larson, C. L., Abercrombie, H. C., . . . Davidson, R. J. (2001).
Anterior cingulate activity as a predictor of degree of treatment response in major depression: Evidence from brain electrical tomography
analysis. American Journal of Psychiatry, 158(3), 405–415.
Platt, M. L., & Glimcher, P. W. (1999). Neural correlates of decision variables in parietal cortex. Nature, 400, 233–238.
Ploj, K., Roman, E., Bergstrom, L., & Nylander, I. (2001). Effects of neonatal handling on nociceptin/orphanin FQ and opioid peptide
levels in female rats. Pharmacology, Biochemistry and Behavior, 69(1–2), 173–179.
Plotsky, P. M., & Meaney, M. J. (1993). Early, postnatal experience alters hypothalamic corticotropin-releasing factor (CRF) MRNA,
median eminence CRF content and stress-induced release in adult rats. Molecular Brain Research, 18(3), 195–200.
Pochon, J. B., Levy, R., Fossati, P., Lehericy, S., Poline, J. B., Pillon, B., . . . Dubois, B. (2002). The neural system that bridges reward and
cognition in humans: An fMRI study. Proceedings of the National Academy of Sciences, USA, 99(8), 5669–5674.
Pollak, S. D., Nelson, C. A., Schlaak, M. F., Roeber, B. J., Wewerka, S. S., Wiik, K. L., . . . Gunnar, M. R. (2010). Neurodevelopmental
effects of early deprivation in post-institutionalized children. Child Development, 81(1), 224–236. doi:10.1111/j.1467-8624.2009.01391.x.
Pollatos, O., Gramann, K., & Schandry, R. (2007). Neural systems connecting interoceptive awareness and feelings. Human Brain Mapping,
28, 9–18.
Polley, D. B., Chen-Bee, C. H., & Frostig, R. D. (1999). Two directions of plasticity in the sensory-deprived adult cortex. Neuron, 24(3),
623–637.
Pomarol-Clotet, E., Salvador, R., Sarró, S., Gomar, J., Vila, F., Martínez, Á, . . . Mckenna, P. (2008). Failure to deactivate in the prefrontal
cortex in schizophrenia: Dysfunction of the default mode network? Psychological Medicine Psychol. Med., 38(08).
doi:10.1017/s0033291708003565
Pope, S. K., Whiteside, L., Brooks-Gunn, J., Kelleher, K. J., Rickert, V. I., Bradley, R. H., & Casey, P. H. (1993). Low-birth-weight
infants born to adolescent mothers. Effects of coresidency with grandmother on child development. Journal of the American Medical Association,
269(11), 1396–1400.
Popescu, A. T., Saghyan, A. A., & Paré, D. (2007). NMDA-dependent facilitation of corticostriatal plasticity by the amygdala. Proceedings
of the National Academy of Sciences, USA, 104(1), 341–346.
Porges, S. W. (2007). The polyvagal perspective. Biological Psychology, 74(2), 116–143.
Porges, S. W., Doussard-Roosevelt, J. A., & Maiti, A. K. (1994). Vagal tone and the physiological regulation of emotion. Monographs of the
Society for Research in Child Development, 59(2–3), 167–186.

277
Porges, S. W., Doussard-Roosevelt, J. A., Portales, A. L., & Greenspan, S. I. (1996). Infant regulation of the vagal “brake” predicts child
behavior problems: A psychobiological model of social behavior. Developmental Psychobiology, 29(8), 697–712.
Porto, P. R., Oliveria, L., Mari, J., Volchan, E., Figueira, I., & Ventura, P. (2009). Does cognitive behavior therapy change the brain? A
systematic review of neuroimaging in anxiety disorders. Journal of Neuropsychiatry and Clinical Neuroscience, 21, 114–125.
Posner, J., Hellerstein, D. J., Gat, I., Mechling, A., Klahr, K., Wang, Z., . . . Peterson, B. S. (2013). Antidepressants normalize the default
mode network in patients with dysthymia. JAMA Psychiatry, 70(4), 373. doi:10.1001/jamapsychiatry.2013.455
Posner, M. I., Rothbart, M. K., Vizueta, N., Levy, K. N., Evans, D. E., Thomas, K. M., & Clarkin, J. F. (2002). Attentional mechanisms
of borderline personality disorder. PNAS, 99(25), 16366–16370.
Post, R. M., & Weiss, S. R. B. (1997). Emergent properties of neural systems: How focal molecular neurobiological alterations can affect
behavior. Development and Psychopathology, 9(4), 907–929.
Post, R. M., Weiss, S. R. B., Li, H., Smith, A., Zhang, L. X., Xing, G., . . . McCann, U. D. (1998). Neural plasticity and emotional
memory. Development and Psychopathology, 10(4), 829–855.
Post, S. G. (2005). Altruism, happiness, and health: It’s good to be good. International Journal of Behavioral Medicine, 12(2), 66–77.
Prasko, J., Horácek, J., Záleský, R., Kopecek, M., Novák, T., Pasková, B., . . . & Höschl, C. (2004). The change of regional brain
metabolism (18FDG PET) in panic disorder during the treatment with cognitive behavioral therapy or antidepressants. Neuroendocrinology
Letters, 25(5), 340–348.
Pribram, K. H. (1991). Brain and perception: Holonomy and structure in figural processing. Hillsdale, NJ: Erlbaum.
Pribram, K. H., & Gill, M. M. (1976). Freud’s “Project” re-assessed: Preface to contemporary cognitive theory and neuropsychology. New York:
Basic Books.
Price, B. H., Daffner, K. R., Stowe, R. M., & Mesulam, M. M. (1990). The comportmental learning disabilities of early frontal lobe
damage. Brain, 113(Pt 5), 1383–1393.
Price, J. L., Carmichael, S. T., & Drevets, W. C. (1996). Networks related to the orbital and medial prefrontal cortex: A substrate for
emotional behavior? Progress in Brain Research, 107, 523–536.
Prickaerts, J., Koopmans, G., Blokland, A., & Scheepens, A. (2004). Learning and adult neurogenesis: Survival with or without
proliferation? Neurobiology of Learning and Memory, 81(1), 1–11.
Prossin, A. R., Love, T. M., Koeppe, R. A., Zubieta, J., & Silk, K. R. (2010). Dysregulation of regional endogenous opioid function in
borderline personality disorder. American Journal of Psychiatry, 167, 925–933.
Pruessner, J. C., Baldwin, M. W., Dedovic, K., Renwick, R., Mahani, N. K., Lord, C., . . . Lupien, S. (2005). Self-esteem, locus of control,
hippocampal volume, and cortisol regulation in young and old adulthood. NeuroImage, 28(4), 815–826.
Pulver, S. E. (2003). On the astonishing clinical irrelevance of neuroscience. Journal of the American Psychoanalytic Association, 51(3), 755–
772.
Purves, D., & Lichtman, J. (1980). Elimination of synapses in the developing nervous system. Science, 210(4466), 153–157.
Purves, D., & Voyvodic, J. T. (1987). Imaging mammalian nerve cells and their connections over time in living animals. Trends in
Neurosciences, 10(10), 398–404.
Quidé, Y., Witteveen, A. B., El-Hage, W., Veltman, D. J., & Olff, M. (2012). Differences between effects of psychological versus
pharmacological treatments on functional and morphological brain alterations in anxiety disorders and major depressive disorder: A systematic
review. Neuroscience and Biobehavioral Reviews, 36(1), 626–644. doi:10.1016/j.neubiorev.2011.09.004
Quin, P., & Northoff, G. (2011). How is our self related to midline regions and the default-mode network? NeuroImage, 57(3), 1221-1233.
doi:10.1016/j.neuroimage.2011.05.028
Quintana, J., & Fuster, J. M. (1999). From perception to action: Temporal integrative functions of prefrontal and parietal neurons. Cerebral
Cortex, 9(3), 213–221.
Quirk, G. J. (2004). Learning not to fear, faster. Learning & Memory, 11(2), 125-126. doi:10.1101/lm.75404
Quirk, G. J., & Beer, J. S. (2006). Prefrontal involvement in the regulation of emotion: Convergence of rat and human studies. Current
Opinion in Neurobiology, 16(6), 723–727.
Quirk, G. J., Likhtik, E., Pelletier, J. G., & Paré, D. (2003). Stimulation of medial prefrontal cortex decreases the responsiveness of central
amygdala output neurons. Journal of Neuroscience, 23(25), 8800–8807.
Rabinak, C. A., Angstadt, M., Welsh, R. C., Kenndy, A. E., Lyubkin, M., Martis, B., & Phan, K. L. (2011). Altered amygdala resting-
state functional connectivity in post-traumatic stress disorder. Frontiers in Psychiatry, 2(62), 1–8.
Radecki, D. T., Brown, L. M., Martinez, J., & Teyler, T. J. (2005). BDNF protects against stress-induced impairments in spatial learning
and memory and LTP. Hippocampus, 15(2), 246–253.
Radley, J. J., Rocher, A. B., Miller, M., Janssen, W. G. M., Liston, C., Hof, P. R., . . . Morrison, J. H. (2006). Repeated stress induces
dendritic spine loss in the rat medial prefrontal cortex. Cerebral Cortex, 16(3), 313–320.

278
Raichle, M. E., MacLeod, A. M., Snyder, A. Z., Powers, W. J., Gusnard, D. A., & Shulman, G. A. (2001). A default mode of brain
function. Proceedings of the National Academy of Sciences, 98(2), 676–682.
Raine, A., Buchsbaum, M. S., Stanley, J., Lottenberg, S., Abel, L., & Stoddard, J. (1994). Selective reductions in prefrontal glucose
metabolism in murderers. Biological Psychiatry, 36(6), 365–373.
Rainnie, D. G., Bergeron, R., Sajdyk, T. J., Patil, M., Gehlert, D. R., & Shekhar, A. (2004). Corticotrophin releasing factor-induced
synaptic plasticity in the amygdala translates stress into emotional disorders. Journal of Neuroscience, 24, 3471–3479.
Rakic, P. (1985). Limits of neurogenesis in primates. Science, 227(4690), 154–156.
Ramachandran, V. S., Rogers-Ramachandran, D., & Stewart, M. (1992). Perceptual correlates of massive cortical reorganization. Science,
258(5085), 1159–1160.
Ramirez, S., Liu, X., Lin, P., Suh, J., Pignatelli, M., Redondo, R. L., . . . Tonegawa, S. (2013). Creating a false memory in the
hippocampus. Science, 341(6144), 387–391. doi:10.1126/science.1239073
Rampon, C., Jiang, C. H., Dong, H., Tang, Y., Lockhart, D. J., Schultz, P. G., . . . Hu, Y. (2000). Effects of environmental enrichment on
gene expression in the brain. Proceedings of the National Academy of Sciences, USA, 97(23), 12880–12884.
Ranote, S., Elliott, R., Abel, K. M., Mitchell, R., Deakin, J. F. W., & Appleby, L. (2004). The neural basis of maternal responsiveness to
infants: An fMRI study. NeuroReport, 15(11), 1825–1829.
Rao, V. R., & Finkbeiner, S. (2007). NMDA and AMPA receptors: Old channels, new tricks. Trends in Neurosciences, 30(6), 284–291.
Rau, V., & Fanselow, M. S. (2007). Neurobiological and neuroethological perspectives on fear and anxiety. In L. J. Kirmayer, R. Lemelson,
& M. Barad (Eds.), Understanding trauma: Integrating biological, clinical, and cultural perspectives (pp. 27–40). New York: Cambridge University
Press.
Rauch, S. L., Jenike, M. A., Alpert, N. M., Baer, L., Breiter, H. C. R., Savage, C.R., & Fischman, A. J. (1994). Regional cerebral blood
flow measured during symptom provocation in obsessive-compulsive disorder using oxygen 15-labeled carbon dioxide and positron emission
tomography. Archives of General Psychiatry, 51(1), 62–70.
Rauch, S. L., Shin, L. M., & Phelps, E. A. (2006). Neurocircuitry models of posttraumatic stress disorder and extinction: Human
neuroimaging research—past, present, and future. Journal of Biological Psychiatry, 60(4), 376–382.
Rauch, S. L., Shin, L. M., Segal, E., Pitman, R. K., Carson, M. A., McMullin, K., . . . Nikos, M. (2003). Selectively reduced regional
cortical volumes in post-traumatic stress disorder. NeuroReport, 14(7), 913–916.
Rauch, S. L., van der Kolk, B. A., Fisler, R. E., Alpert, N. M., Orr, S. P., Savage, C. R., . . . Pitman, R. K. (1996). A symptom provocation
study of posttraumatic stress disorder using positron emission tomography and script-driven imagery. Archive of General Psychiatry, 53(5), 380–
387.
Ray, S. L., & Vanstone, M. (2009). The impact of PTSD on veterans’ family relationships: An interpretative phenomenological inquiry.
International Journal of Nursing Studies, 46(6), 838–847. doi:10.1016/j.ijnurstu.2009.01.002
Rees, G., Kreiman, G., & Koch, C. (2002). Neural correlates of consciousness in humans. Nature Reviews Neuroscience, 3(4), 261–270.
Rees, S. L., Steiner, M., & Fleming, A. S. (2006). Early deprivation, but not maternal separation, attenuates rise in corticosterone levels
after exposure to a novel environment in both juvenile and adult female rats. Behavioral Brain Research, 175(2), 383–391.
Regard, M., Oelz, O., Brugger, P., & Landis, T. (1989). Persistent cognitive impairment in climbers after repeated exposure to extreme
altitude. Neurology, 39(2 Pt 1), 210–213.
Reich, W. (1945). Character analysis. New York: Simon and Schuster.
Reiman, E. M., Lane, R. D., Ahern, G. L., Schwartz, G. E., Davidson, R. J., Friston, K. J., . . . Chen, K. (1997). Neuroanatomical
correlates of externally and internally generated human emotion. American Journal of Psychiatry, 154, 918–925.
Reiman, E. M., Raichle, M. E., Robins, E., Mintun, M. A., Fusselman, M. J., Fox, P. T., . . . Hackman, K. A. (1989). Neuroanatomical
correlates of a lactate-induced anxiety attack. Archives of General Psychiatry, 46(6), 493–500.
Resnick, H. S., Yehuda, R., Pitman, R. K., & Foy, D. W. (1995). Effects of previous trauma on acute plasma cortisol level following rape.
American Journal of Psychiatry, 152, 1675–1677.
Ressler, K. J., Rothbaum, B. O., Tannenbaum, L., Anderson, P., Graap, K., Zimand, E., . . . Davis, M. (2004). Cognitive enhancers as
adjuncts to psychotherapy. Archives of General Psychiatry, 61(11), 1136–1144.
Reynolds, D. K. (1980). The quiet therapies: Japanese pathways to personal growth. Honolulu: University Press of Hawaii.
Rezai, K., Andreasen, N. C., Alliger, R., Cohen, G., Swayze, V., & O’Leary, D. S. (1993). The neuropsychology of the prefrontal cortex.
Archives of Neurology, 50(6), 636–642.
Rholes, W. S., Paetzold, R. L., & Kohn, J. L. (2016). Disorganized attachment mediates the link from early trauma to externalizing
behavior in adult relationships. Personality and Individual Differences, 90, 61–65.
Richards, M., & Deary, I. J. (2005). A life course approach to cognitive reserve: A model for cognitive aging and development? Annals of
Neurology, 58(4), 617–622.

279
Riggs, D. S., Byrne, C. A., Weathers, F. W., & Litz, B. T. (1998). The quality of the intimate relationships of male Vietnam veterans:
Problems associated with posttraumatic stress disorder. Journal of Traumatic Stress, 11(1), 87–101. doi:10.1023/a:1024409200155
Rilling, J. K., Gutman, D. A., Zeh, T. R., Panoni, G., Berns, G. S., & Kilts, C. D. (2002). A neural basis for social cooperation. Neuron,
35(2), 395–405.
Rilling, J. K., & Young, L. J. (2014). The biology of mammalian parenting and its effect on offspring social development. Science,
345(6198), 771–776. doi:10.1126/science.1252723
Rizzolatti, G., & Arbib, M. A. (1998). Language within our grasp. Trends in Neurosciences, 21(5), 188–194.
Rizzolatti, G., Fadiga, L., Fogassi, L., & Gallese, V. (1999). Resonance behaviors and mirror neurons. Archives Italiennes De Biologie, 137,
85-100.
Rizzolatti, G., & Sinigaglia, C. (2008). Mirrors in our brain: How our minds share actions and emotions. New York: Oxford University Press.
Roberts, A. C., & Wallis, J. D. (2000). Inhibitory control and affective processing in the prefrontal cortex: Neuropsychological studies in the
common marmoset. Cerebral Cortex, 10(3), 252–262.
Robinson, R. G., Kubos, K. L., Starr, L. B., Rao, K., & Price, T. R. (1984). Mood disorder in stroke patient: Importance of location of
lesion. Brain, 1, 91–93.
Rodrigues, E., Wenzel, A., Ribeiro, M. P., Quarantini, L. C., Miranda-Scippa, A., de Sena, E. P., & de Olíveira, I. R. (2011).
Hippocampal volume in borderline personality disorder with and without comorbid posttraumatic stress disorder: A meta-analysis. European
Psychiatry, 26, 452–456.
Roffman, J. L., Marci, C. D., Glick, D. M., Dougherty, D. D., & Rauch, S. L. (2005). Neuroimaging and the functional neuroanatomy of
psychotherapy. Psychological Medicine, 35(10), 1–14.
Rogan, M. T., & LeDoux, J. E. (1996). Emotion: Systems, cells, synaptic plasticity. Cell, 85(4), 469–475.
Rogan, M. T., Staubli, U. V., & LeDoux, J. E. (1997). Fear conditioning induces associative long-term potentiation in the amygdala.
Nature, 390(6660), 604–607.
Rogers, A. R. (1993). Why menopause? Evolutionary Ecology, 7(4), 406–420.
Rogers, C. E., Anderson, P. J., Thompson, D. K., Kidokoro, H., Wallendorf, M., Treyvaud, K., . . . Inder, T. E. (2012). Regional cerebral
development at term relates to school-age social–emotional development in very preterm children. Journal of the American Academy of Child &
Adolescent Psychiatry, 51(2), 181-191. doi:10.1016/j.jaac.2011.11.009
Rogers, C. R. (1942). Counseling and psychotherapy. Boston: Houghton Mifflin.
Rogers, R. D., Ramnani, N., Mackay, C., Wilson, J. L., Jezzard, P., Carter, C. S., & Smith, S. M. (2004). Distinct portions of anterior
cingulate cortex and medial prefrontal cortex are activated by reward processing in separable phases of decision-making cognition. Biological
Psychiatry, 55(6), 594–602.
Rokeach, M. (1964). The three Christs of Ypsilanti. New York: Columbia University Press.
Roozendaal, B. (1999). Glucocorticoids and the regulation of memory consolidation. Psychoneuroendocrinology, 25(3), 213–238.
Roozendaal, B., & McGaugh, J. L. (2011). Memory modulation. Behavioral Neuroscience, 125(6), 797-824. doi:10.1037/a0026187
Ropper, A. H., & Brown, R. H. (2005). Adams and Victor’s principles of neurology. New York: McGraw-Hill.
Rorden, C., Mattingley, J., Karnath, H., & Driver, J. (1997). Visual extinction and prior entry: Impaired perception of temporal order with
intact motion perception after unilateral parietal damage. Neuropsychologia, 35(4), 421–433.
Rosenkranz, J. A., Moore, H., & Grace, A. A. (2003). The prefrontal cortex regulates lateral amygdala neuronal plasticity and responses to
previously conditioned stimuli. Journal of Neuroscience, 23(35), 11054–11064.
Rosenkranz, M. A., Busse, W. W., Johnstone, T., Swenson, C. A., Crisafi, G. M., Jackson, M. M., . . . Davidson, R. J. (2005). Neural
circuitry underlying the interaction between emotion and asthma symptom exacerbation. Proceedings of the National Academy of Sciences of the
United States of America, 102(37), 13319–13324.
Rosenzweig, M. R. (2001). Learning and neural plasticity over the life span. In P. E. Gold & W. T. Greenough (Eds.), Memory
consolidation: Essays in honor of James L. McGauggh. Washington, DC: American Psychological Association.
Ross, E. D., Homan, R. W., & Buck, R. (1994). Differential hemispheric lateralization of primary and social emotions: Implications for
developing a comprehensive neurology for emotions, repression, and the subconscious. Neuropsychiatry, Neuropsychology, and Behavioral
Neurology, 7(1), 1–19.
Rossi, E. L. (1993). The psychobiology of mind-body healing. New York: Norton.
Roth, T., & Sweatt, J. (2011). Epigenetic mechanisms and environmental shaping of the brain during sensitive periods of development.
Journal of Child Psychology and Psychiatry, 52(4), 398–408. doi:10.1111/j.1469-7610.2010.02282.x
Rothbart, M. K., Taylor, S. B., & Tucker, D. M. (1989). Right-sided facial asymmetry in infant emotional expression. Neuropsychologia,
27(5), 675–687.
Royer, S., Martina, M., & Paré, D. (1999). An inhibitory interface gates impulse traffic between the input and output stations of the

280
amygdala. Journal of Neuroscience, 19(23), 10575–10583.
Rubens, A. B. (1985). Caloric stimulation and unilateral visual neglect. Neurology, 35(7), 1019–1024.
Rubia, K., Overmeyer, S., Taylor, E., Brammer, M., Williams, S., Simmons, A., & Bullmore, E. T. (1999). Hypofrontality in attention
deficit hyperactivity disorder during higher-order motor control: A study with functional MRI. American Journal of Psychiatry, 156(6), 891–896.
Rubia, K., Smith, A. B., Brammer, M. J., Toone, B., & Taylor, E. (2005). Abnormal brain activation during inhibition and error detection
in medication-naive adolescents with ADHD. American Journal of Psychiatry, 162(6), 1067–1075. doi:10.1176/appi.ajp.162.6.1067
Rubino, G. J., Farahani, K., McGill, D., Van de Wiele, B., Villablanca, J. P., & Wang-Maithieson, A. (2000). Magnetic resonance
imaging-guided neurosurgery in the magnetic fringe fields: The next step in neuronavigation. Neurosurgery, 46(3), 643–654.
Ruby, P., & Decety, J. (2001). Effect of subjective perspective taking during simulation of action: A PET investigation of agency. Nature
Neuroscience, 4(5), 546–550.
Rudebeck, P., Mitz, A., Chacko, R., & Murray, E. (2013). Effects of amygdala lesions on reward-value coding in orbital and medial
prefrontal cortex. Neuron, 80(6), 1519–1531. doi:10.1016/j.neuron.2013.09.036
Ruocco, A. C., Medaglia, J. D., Tinker, J. R., Ayaz, H., Forman, E. M., Newman, C. F., . . . Chute, D. L. (2010). Medial prefrontal cortex
hyperactivation during social exclusion in borderline personality disorder. Psychiatry Research: Neuroimaging, 181, 233–236.
Rüsch, N., Schulz, D., Valerius, G., Steil, R., Bohus, M., & Schmahl, C. (2011). Disgust and implicit self-concept in women with
borderline personality disorder and posttraumatic stress disorder. European Archives of Psychiatry and Clinical Neuroscience, 261(5), 369–376.
doi:10.1007/s00406-010-0174-2
Rushworth, M. F. S., & Behrens, T. E. J. (2008). Choice, uncertainty and value in prefrontal and cingulate cortex. Nature Neuroscience,
11(4), 389–397.
Rushworth, M. F. S., Krams, M., & Passingham, R. E. (2001). The attentional role of the left parietal cortex: The distinct lateralization
and localization of motor attention in the human brain. Journal of Cognitive Neuroscience, 13(5), 698–710.
Russo-Neustadt, A. A., Beard, R. C., Huang, Y. M., & Cotman, C. W. (2000). Physical activity and antidepressant treatment potentiate
the expression of specific brain-derived neurotrophic factor transcripts in the rat hippocampus. Neuroscience, 101(2), 305–312.
Rutter, M., Beckett, C., Castle, J., Colvert, E., Kreppner, J., Mehta, M., . . . Sonuga-Barke, E. (2007). Effects of profound early
institutional deprivation: An overview of findings from a UK longitudinal study of Romanian adoptees. European Journal of Developmental
Psychology, 4(3), 332–350. doi:10.1080/17405620701401846
Rutter, M., & Rutter, M. (1993). Developing minds: Challenge and continuity across the life span. New York: Basic Books.
Rutter, M. L. (1997). Nature–nurture integration: The example of antisocial behavior. American Psychologist, 52(4), 390-398.
doi:10.1037/0003-066x.52.4.390
Ryan, W. (1971). Blaming the victim. New York: Pantheon.
Saarela, M. V., Hlushchuk, Y., Williams, A. C. de C., Schürmann, M., Kalso, E., & Hari, R. (2007). The compassionate brain: Humans
detect intensity of pain from another’s face. Cerebral Cortex, 17, 230–237.
Saba, G., Rocamora, J. F., Kalalou, K., Benadhira, R., Plaze, M., & Lipski, H. (2004). Repetitive transcranial magnetic stimulation as an
add-on therapy in the treatment of mania: A case series of eight patients. Psychiatry Research, 128(2), 199–202.
Sabbagh, M. A. (2004). Understanding orbitofrontal contributions to theory-of-mind reasoning: Implications for autism. Brain and
Cognition, 55(1), 209–219.
Sackheim, H. A., Greenberg, M. S., Weiman, A. L., Gur, R. C., Hungerbuhler, J. P., & Geschwind, N. (1982). Hemispheric asymmetry in
the expression of positive and negative emotions: Neurologic evidence. Archives of Neurology, 39(4), 210–218.
Sackeim, H. A., Putz, E., Vingiano, W., Coleman, E., & McElhiney, M. (1988). Lateralization in the processing of emotionally laden
information. I. Normal functioning. Neuropsychiatry, Neuropsychology, and Behavioral Neurology, 1(2), 97–110.
Sakai, Y., Kumano, H., Nishikawa, M., Sakano, Y., Kaiya, H., Imavayashi, E., . . . Kuboki, T. (2006). Changes in cerebral glucose
utilization in patients with panic disorder treated with cognitive-behavioral therapy. NeuroImage, 33(1), 218–226.
Sakamoto, H., Fukuda, R., Okuaki, T., Rogers, M., Kasai, K., Machida, T., . . . Kato, N. (2005). Parahippocampal activation evoked by
masked traumatic images in posttraumatic stress disorder: A functional MRI study. NeuroImage, 26(3), 813–821.
Salavert, J., Gasol, M., Vieta, E., Cervantes, A., Trampal, C., & Gispert, J. D. (2011). Fronto-limbic dysfunction in borderline personality
disorder: A 18F-FDG positron emission tomography study. Journal of Affective Disorders, 131, 260–267.
Salazar, R., Dotson, N., Bressler, S., & Gray, C. (2012). Content-specific fronto-parietal synchronization during visual working memory.
Science, 338(6110), 1097–1100. doi:10.1126/science.1224000
Salomon, R., Levy, D. R., & Malach, R. (2013). Deconstructing the default: Cortical subdivision of the default mode/intrinsic system
during self-related processing. Human Brain Mapping, 35(4), 1491-1502. doi:10.1002/hbm.22268
Salm, A. K., Modney, B. K., & Hatton, G. I. (1988). Alterations in supraoptic nucleus ultrastructure of maternally behaving virgin rats.
Brain Research Bulletin, 21(4), 685–691.

281
Sambataro, F., Murty, V. P., Callicott, J. H., Tan, H., Das, S., Weinberger, D. R., & Mattay, V. S. (2010). Age-related alterations in
default mode network: Impact on working memory performance. Neurobiology of Aging, 31(5), 839-852.
doi:10.1016/j.neurobiolaging.2008.05.022
Sánchez, M. M., Noble, P. M., Lyon, C. K., Plotsky, P. M., Davis, M., Nemeroff, C. B., & Winslow, J. T. (2005). Alterations in diurnal
cortisol rhythm and acoustic startle response in nonhuman primates with adverse rearing. Biological Psychiatry, 57(4), 373–381.
doi:10.1016/j.biopsych.2004.11.032
Sanfey, A. G., Rilling, J. K., Aronson, J. A., Nystrom, L. E., & Cohen, J. D. (2003). The neural basis of economic decision-making in the
ultimatum game. Science, 300, 1755–1758.
Sansone, R. A., Sansone, L. A., & Gaither, G. A. (2004). Multiple types of childhood trauma and borderline personality symptomatology
among a sample of diabetic patients. Traumatology, 10(4), 257–266.
Santini, E., Ge, H., Ren, K., Peña de Ortiz, S., & Quirk, G. L. (2004). Consolidation of fear extinction requires protein synthesis in the
medial prefrontal cortex. Journal of Neuroscience, 24(25), 5704–5710.
Sapolsky, R. M. (1985). A mechanism for glucocorticoid toxicity in the hippocampus: Increased neuronal vulnerability to metabolic insults.
Journal of Neuroscience, 5(5), 1228–1232.
Sapolsky, R. M. (1987). Glucocorticoids and hippocampal damage. Trends in Neurosciences, 10(9), 346–349.
Sapolsky, R. M. (1990). Stress in the wild. Scientific American, 262(1), 116–123.
Sapolsky, R. M. (1996). Why stress is bad for your brain. Science, 273(5276), 749–750.
Sapolsky, R. M. (2004). Mothering style and methylation. Nature Neuroscience, 7(8), 791–792.
Sapolsky, R. M., Krey, L. C., & McEwen, B. S. (1984). Glucocorticoid-sensitive hippocampal neurons are involved in terminating the
adrenocortical stress response. Proceedings of the National Academy of Sciences, USA, 81(19), 6174–6177.
Sapolsky, R. M., Uno, H., Rebert, C. S., & Finch, C. E. (1990). Hippocampal damage associated with prolonged glucocorticoid exposure
in primates. Journal of Neuroscience, 10(9), 2897–2902.
Sar, V., Akyuz, G., Kugu, N., Ozturk, E., & Ertem-Vehid, H. (2006). Axis I dissociative disorder comorbidity in borderline personality
disorder and reports of childhood trauma. Journal of Clinical Psychiatry, 67(10), 1583–1590.
Saroglou, V. (2013). Religion, spirituality, and altruism. APA Handbook of Psychology, Religion, and Spirituality, 1, 439–457.
Saroglou, V., Pichon, I., Trompette, L., Verschueren, M., & Dernelle, R. (2005). Prosocial behavior and religion: New evidence based on
projective measures and peer ratings. Journal for the Scientific Study of Religion, 44(3), 323–348.
Sarrieau, A., Sharma, S., & Meaney, M. J. (1988). Postnatal development and environmental regulation of hippocampal glucocorticoid and
mineralocorticoid receptors. Developmental Brain Research, 43(1), 158–162.
Sarter, M., & Markowitsch, H. J. (1985). The amygdala’s role in human mnemonic processing. Cortex, 21(1), 7–24.
Satterfield, J. H., & Dawson, M. E. (1971). Electrodermal correlates of hyperactivity in children. Psychophysiology, 8(2), 191–197.
Sauseng, P., Klimesch, W., Schabus, M., & Doppelmayr, M. (2005). Frontoparietal EEG coherence in theta and upper alpha reflect central
executive functions of working memory. International Journal of Psychophysiology, 57, 97–103.
Sawamoto, N., Honda, M., Okada, T., Hanakawa, T., Kanda, M., Fukuyama, H., . . . Shibasaki, H. (2000). Expectation of pain enhances
responses to nonpainful somatosensory stimulation in the anterior cingulate cortex and parietal operculum/posterior insula: An event-related
functional magnetic resonance imaging study. The Journal of Neuroscience, 20(19), 7438-7445.
Saxe, G. N., Chinman, G., Berkowitz, R., Hall, K., Leiberg, G., Schwartz, J., & van der Kolk, B. A. (1994). Somatization in patients with
dissociative disorders. American Journal of Psychiatry, 151(9), 1329–1333.
Saxena, S. (1999). Localized orbitofrontal and subcortical metabolic changes and predictors of response to paroxetine treatment in obsessive-
compulsive disorder. Neuropsychopharmacology, 21(6), 683-693. doi:10.1016/s0893-133x(99)00082-2
Saxena, S., Brody, A. L., Ho, M. L., Zohrabi, N., Maidment, K. M., & Baxter, L. R. (2003). Differential brain metabolic predictors of
response to paroxetine in obsessive-compulsive disorder, versus major depression. American Journal Psychiatry, 160(3), 522–532.
Saxena, S., Gorbis, E., O’Neill, J., Baker, S. K., Mandelkern, M. A., Maidment, K. M., . . . London, E. D. (2008). Rapid effects of brief
intensive cognitive-behavioral therapy on brain glucose metabolism in obsessive-compulsive disorder. Molecular Psychiatry, 14(2), 197–205.
doi:10.1038/sj.mp.4002134
Scarmeas, N., Zarahn, E., Anderson, K. E., Honig, L. S., Hilton, J., Flynn, J., . . . Stern, Y. (2004). Cognitive reserve-mediated modulation
of positron emission tomographic activations during memory tasks in Alzheimer disease. Archives of Neurology, 61(1), 73–78.
Schaaf, M. J. M., de Kloet, E. R., & Vreugdenhil, E. (2000). Corticosterone effects on BDNF expression in the hippocampus: Implications
for memory formation. Stress, 3(3), 201–208.
Schaafsma, S., Pfaff, D., Spunt, R., & Adolphs, R. (2015). Deconstructing and reconstructing theory of mind. Trends in Cognitive Sciences,
19(2), 65–72. doi:10.1016/j.tics.2014.11.007
Schacter, D. L. (1976). The hypnagogic state: A critical review of the literature. Psychological Bulletin, 83(3), 452–481.

282
Schacter, D. L. (1986). Amnesia and crime. American Psychologist, 41(3), 286–295.
Schacter, D. L. (1996). Searching for memory: The brain, the mind, and the past. New York: Basic Books.
Schaefer, S. M., Jackson, D. C., Davidson, R. J., Aguirre, G. K., Kimberg, D. Y., & Thompson-Schill, S. L. (2002). Modulation of
amygdalar activity by the conscious regulation of negative emotion. Journal of Cognitive Neuroscience, 14(6), 913–921.
Schaie, K. W., & Willis, S. L. (1986). Can decline in adult intellectual functioning be reversed? Developmental Psychology, 22(2), 223–232.
Schall, J. D. (2001). Neural basis of deciding, choosing and acting. Nature Reviews Neuroscience, 2, 33–42.
Schiffer, F., Teicher, M. H., & Papanicolaou, A. C. (1995). Evoked potential evidence for right brain activity during the recall of traumatic
memories. Journal of Neuropsychiatry and Clinical Neurosciences, 7(2), 169–175.
Schilbach, L., Eickhoff, S. B., Rotarska-Jagiela, A., Fink, G. R., & Vogeley, K. (2008). Minds at rest? Social cognition as the default mode
of cognizing and its putative relationship to the “default system” of the brain. Consciousness and Cognition, 17(2), 457–467.
doi:10.1016/j.concog.2008.03.013
Schmahl, C., Berne, K., Krause, A., Kleindienst, N., Valerius, G., Vermetten, E., & Bohus, M. (2009). Hippocampus and amygdala
volumes in patients with borderline personality disorder with or without posttraumatic stress disorder. Journal of Psychiatry and Neuroscience,
34(4), 289–295.
Schmahl, C. G., Elzinga, B. M., Vermetten, E., Sanislow, C., McGlashan, T. H., & Bremner, J. D. (2003). Neural correlates of memories
of abandonment in women with and without borderline personality disorder. Biological Psychiatry, 54, 142–151.
Schmahmann, J. D. (1997). The cerebellum and cognition. New York: Academic Press.
Schmand, B., Smit, J. H., Geerlings, M. I., & Lindeboom, J. (1997). The effects of intelligence and education on the development of
dementia: A test of the brain reserve hypothesis. Psychological Medicine, 27(6), 1337–1344.
Schneider, F., Bermpohl, F., Heinzel, A., Rotte, M., Walter, M., Tempelmann, C., . . . Northoff, G. (2008). The resting brain and our self:
Self-relatedness modulates resting state neural activity in cortical midline structures. Neuroscience, 157(1), 120-131.
doi:10.1016/j.neuroscience.2008.08.014
Schneider, M. L. (1992). Prenatal stress exposure alters postnatal behavioral expression under conditions of novelty challenge in rhesus
monkey infants. Developmental Psychobiology, 25(7), 529–540.
Schneiderman, I., Zagoory-Sharon, O., Leckman, J. F., & Feldman, R. (2012). Oxytocin during the initial stages of romantic attachment:
Relations to couples’ interactive reciprocity. Psychoneuroendocrinology, 37(8), 1277–1285.
Schnell, K., Dietrich, T., Schnitker, R., Daumann, J., & Herpertz, S. C. (2007). Processing of autobiographical memory retrieval cues in
borderline personality disorder. Journal of Affective Disorders, 97, 253–259.
Schore, A. N. (1994). Affect regulation and the origin of the self: The neurobiology of emotional development. Hillsdale, NJ: Erlbaum.
Schore, A. N. (1997a). Early organization of the nonlinear right brain and development of a predisposition to psychiatric disorders.
Development and Psychopathology, 9(4), 595–631.
Schore, A. N. (1997b). A century after Freud’s project for a scientific psychology: Is a rapprochement between psychoanalysis and
neurobiology at hand? Journal of the American Psychoanalytic Association, 45, 841–867.
Schore, A. N. (2000). Attachment and the regulation of the right brain. Attachment and Human Development, 2(1), 23–47.
Schore, A. N. (2011). The right brain implicit self lies at the core of psychoanalysis. Psychoanalytic Dialogues, 21(1), 75-100.
doi:10.1080/10481885.2011.545329
Schore, A. N. (2014). The right brain is dominant in psychotherapy. Psychotherapy, 51(3), 388-397. doi:10.1037/a0037083
Schore, J., & Schore, A. (2008). Modern attachment theory: The central role of affect regulation in development and treatment. Clinical
Social Work Journal, 36(1), 9–20.
Schreiber, D., Fonzo, G., Simmons, A. N., Dawes, C. T., Flagan, T., Fowler, J. H., & Paulus, M. P. (2013). Red brain, blue brain:
Evaluative processes differ in Democrats and Republicans. PLoS One, 8(2), 1–6.
Schrott, L. M. (1997). Effect of training and environment on brain morphology and behavior. Acta Paediatrica Scandanavia, 422(Suppl.),
45–47.
Schrott, L. M., Denenberg, V. H., Sherman, G. F., Waters, N. S., Rosen, G. D., & Galaburda, A. M. (1992). Environmental enrichment,
neocortical ectopias, and behavior in the autoimmune NZB mouse. Developmental Brain Research, 67(1), 85–93.
Schultz, R. T., Gauthier, I., Klin, A., Fulbright, R. K., Anderson, A. W., Volkmar, F., . . . Gore, J. C. (2000). Abnormal ventral temporal
cortical activity during face discrimination among individuals with autism and Asperger syndrome. Archives of General Psychiatry, 57(4), 331–
340.
Schultz, W. (1998). Predictive reward signal of dopamine neurons. Journal of Neurophysiology, 80(1), 1–27.
Schultz, W., Apicella, P., Scarnati, E., & Ljunberg, T. (1992). Neuronal activity in monkey ventral striatum related to the expectation of
reward. Journal of Neuroscience, 12(12), 4595–4610.
Schultz, W., Dayan, P., & Montague, P. R. (1997). A neural substrate of prediction and reward. Science, 275(5306), 1593–1599.

283
Schultz, W., Tremblay, L., & Hollerman, J. R. (2000). Reward processing in primate orbitofrontal cortex and basal ganglia. Cerebral Cortex,
10(3), 272–283.
Schulz, K. P., Fan, J., Tang, C. Y., Newcorn, J. H., Buchsbaum, M. S., Cheung, A. M., & Halperin, J. M. (2004). Response inhibition in
adolescents diagnosed with attention deficit hyperactivity disorder during childhood: An event-related fMRI study. American Journal of
Psychiatry, 161(9), 1650–1657.
Schutter, D. J. L. G. (2009). Antidepressant efficacy of high-frequency transcranial magnetic stimulation over the left dorsolateral prefrontal
cortex in double-blind sham-controlled designs: A meta analysis. Psychological Medicine, 39(1), 65–75.
Schuz, A. (1978). Some facts and hypotheses concerning dendritic spines and learning. In M. A. B. Braizer & H. Petsche (Eds.),
Architectonics of the cerebral cortex (pp. 129–135). New York: Raven.
Schwartz, D. A. (1979). The suicidal character. Psychiatric Quarterly, 51(1), 64–70. doi:10.1007/BF01064720
Schwartz, J. M. (1996). Brain lock: Free yourself from obsessive-compulsive behaviors. New York: ReganBooks.
Schwartz, J. M., Stoessel, P. W., Baxter, L. R., Martin, K. M., & Phelps, M. E. (1996). Systematic changes in cerebral glucose metabolic
rate after successful behavior modification treatment of obsessive-compulsive disorder. Archives of General Psychiatry, 53(2), 109–113.
Schwartz, S., Assal, F., Valenza, N., Seghier, M. L., & Vuilleumier, P. (2005). Illusory persistence of touch after right parietal damage:
Neural correlates of tactile awareness. Brain, 128(Pt. 2), 277–290.
Sear, R., Mace, R., & McGregor, I. A. (2000). Maternal grandmothers improve nutritional status and survival of children in rural Gambia.
Proceedings Biological Sciences, The Royal Society, 267(1453), 1641–1647.
Searleman, A. (1977). A review of right hemisphere linguistic capabilities. Psychological Bulletin, 84(3), 503–528.
Seeley, W. W., Allman, J. M., Carlin, D. A., Crawford, R. K., Macedo, M. N., Greicius, M. D., . . . Miller, B. L. (2007). Divergent social
functioning in behavioral variant frontotemporal dementia and Alzheimer disease: Reciprocal networks and neuronal evolution. Alzheimer
Disease and Associated Disorders, 21(4), S50–S57.
Seidman, L. J., Faracone, S. V., Goldstein, J. M., Goodman, J. M., Kremen, W. S., Toomey, R., . . . Tsuang, M. T. (1999). Thalamic and
amygdala-hippocampal volume reductions in first-degree relatives of patients with schizophrenia: An MRI-based morphometric analysis.
Biological Psychiatry, 46(7), 941–954.
Seidman, L. J., Valera, E. M., & Makris, N. (2005). Structural brain imaging of attention-deficit/hyperactivity disorder. Biological
Psychiatry, 57(11), 1263–1272.
Seitz, R. J., Nickel, J., & Azari, N. P. (2006). Functional modularity of the medial prefrontal cortex: Involvement in human empathy.
Neuropsychology, 20(6), 743–751.
Selden, N. R. W., Everitt, B. J., Jarrard, L. E., & Robbins, T. W. (1991). Complimentary roles for the amygdala and hippocampus in
aversive conditioning to explicit and contextual cues. Neuroscience, 42, 335–350.
Selye, H. (1979). The stress of my life. New York: Van Nostrand.
Semiz, U., Basoglu, C., Cetin, M., Ebrinc, S., Uzun, O., & Ergun, B. (2008). Body dysmorphic disorder in patients with borderline
personality disorder: Prevalence, clinical characteristics, and role of childhood trauma. Acta Neuropsychiatrica, 20, 33–40.
Semmes, J. (1968). Hemispheric specialization: A possible clue to mechanism. Neuropsychologia, 6(1), 11–26.
Sergent, J. (1986). Subcortical coordination of hemispheric activity in commissurotomized patients. Brain, 109, 357–369.
Sergent, J. (1990). Furtive incursions into bicameral minds. Brain, 113(Pt. 2), 537–568.
Serieux, P., & Capgras, J. (1909). Misinterpretive delusional states. In Les folies raisonnantes: Le delire d’interpretation (pp. 5–43). Paris:
Balliere.
Shao, Z., Watanabe, S., Christensen, R., Jorgensen, E., & Colón-Ramos, D. (2013). Synapse location during growth depends on glia
location. Cell, 154(2), 337–350. doi:10.1016/j.cell.2013.06.028
Shapira-Lichter, I., Oren, N., Jacob, Y., Gruberger, M., & Hendler, T. (2013). Portraying the unique contribution of the default mode
network to internally driven mnemonic processes. Proceedings of the National Academy of Sciences, 110(13), 4950-4955.
doi:10.1073/pnas.1209888110
Shapiro, D., Jamner, L. D., & Spence, S. (1997). Cerebral laterality, repressive coping, autonomic arousal, and human bonding. Acta
Scandinavica Physiologica, 640(Suppl.), 60–64.
Shapiro, F. (1995). Eye movement desensitization and reprocessing: Basic principles, protocols, and procedures. New York: Guilford.
Sharp, D. J., Beckmann, C. F., Greenwood, R., Kinnunen, K. M., Bonnelle, V., De Boissezon, X., . . . Leech, R. (2011). Default mode
network functional and structural connectivity after traumatic brain injury. Brain: A Journal of Neurology, 134(8), 2233–2247.
doi:10.1093/brain/awr175
Shatz, C. J. (1990). Impulse activity and patterning of connections during CNS development. Neuron, 5, 745–756.
Sheline, Y. I., Barch, D. M., Price, J. L., Rundle, M. M., Vaishnavi, S. N., Snyder, A. Z., . . . Raichle, M. E. (2009). The default mode
network and self-referential processes in depression. Proceedings of the National Academy of Sciences PNAS, 106(6), 1942-1947.

284
doi:10.1073/pnas.0812686106
Sheline, Y. I., Gado, M. H., & Price, J. L. (1998). Amygdala core nuclei volumes are decreased in recurrent major depression. NeuroReport,
9(9), 2023–2028.
Sheline, Y. I., Wang, P. W., Gado, M. H., Csernansky, J. G., & Vannier, M. W. (1996). Hippocampal atrophy in recurrent major
depression. Proceedings of the National Academy of Sciences, USA, 93(9), 3908–3913.
Shenton, M. E., Kikinis, R., Jolesz, F. A., Pollak, S. D., LeMay, M., Wible, . . . McCarley, R. W. (1992). Abnormalities of the left
temporal lobe and thought disorder in schizophrenia: A quantitative magnetic resonance imaging study. New England Journal of Medicine,
327(9), 604–612.
Sherry, D. F., & Schacter, D. L. (1987). The evolution of multiple memory systems. Psychological Review, 94(4), 439–454.
Sherry, D. F., Jacobs, L. F., & Gaulin, S. J. C. (1992). Spatial memory and adaptive specialization of the hippocampus. Trends in
Neurosciences, 15(8), 298–303.
Shilony, E., & Grossman, F. K. (1993). Depersonalization as a defense mechanism in survivors of trauma. Journal of Traumatic Stress, 6(1),
119–128.
Shima, K., & Tanji, J. (1998). Role for cingulate motor area cells in voluntary movement selection based on reward. Science, 282(5392),
1335–1338.
Shimada, K., Takiguchi, S., Mizushima, S., Fujisawa, T. X., Saito, D. N., Kosaka, H., . . . Tomoda, A. (2015). Reduced visual cortex grey
matter volume in children and adolescents with reactive attachment disorder. NeuroImage: Clinical, 9, 13–19. doi:10.1016/j.nicl.2015.07.001
Shin, J., Geerling, J. C., & Loewy, A. D. (2008). Inputs to the ventrolateral bed nucleus of the stria terminalis. The Journal of Comparative
Neurology, 511(5), 628-657. doi:10.1002/cne.21870
Shin, L. M., Dougherty, D. D., Orr, S. P., Pitman, R. K., Lasko, M., Macklin, M. L., . . . Rauch, S. L. (2000). Activation of anterior
paralimbic structures during guilt-related script-driven imagery. Biological Psychiatry, 48, 43–50.
Shin, L. M., Lasko, N. B., Macklin, M. L., Karpf, R. D., Milad, M. R., Orr, S. P., . . . Pitman, R. K. (2009). Resting Metabolic Activity in
the Cingulate Cortex and Vulnerability to Posttraumatic Stress Disorder. Arch Gen Psychiatry Archives of General Psychiatry, 66(10), 1099.
doi:10.1001/archgenpsychiatry.2009.138
Shin, L. M., Rauch, S. L., & Pitman, R. K. (2006). Amygdala, medial prefrontal cortex, and hippocampal function in PTSD. Annals of the
New York Academy of Science, 1071, 67–79.
Shmuelof, L., & Zohary, E. (2006). A mirror representation of others’ actions in the human anterior parietal cortex. Journal of Neuroscience,
26(38), 9736–9742.
Shobe, E. R. (2014, April 22). Independent and collaborative contributions of the cerebral hemispheres to emotional processing. Frontiers in
Human Neuroscience, 8, 1-19. doi:10.3389/fnhum.2014.00230
Shomstein, S. (2012). Cognitive functions of the posterior parietal cortex: Top-down and bottom-up attentional control. Frontiers in
Integrative Neuroscience, 6(38), 1–7. doi:10.3389/fnint.2012.00038
Shulman, G. L., Corbetta, M., Buckner, R. L., Fiez, J. A., Miezin, F. M., Raichle, M. E., & Petersen, S. E. (1997a). Common blood flow
changes across visual tasks: I. increases in subcortical structures and cerebellum but not in nonvisual cortex. Journal of Cognitive Neuroscience,
9(5), 624-647. doi:10.1162/jocn.1997.9.5.624
Shulman, G. L., Fiez, J. A., Corbetta, M., Buckner, R. L., Miezin, F. M., Raichle, M. E., & Petersen, S. E. (1997b). Common blood flow
changes across visual tasks: II. decreases in cerebral cortex. Journal of Cognitive Neuroscience, 9(5), 648-663. doi:10.1162/jocn.1997.9.5.648
Siegel, D. J. (1995). Trauma and psychotherapy: A cognitive sciences view. Journal of Psychotherapy Practice and Research, 4(5), 93–122.
Siegel, D. J. (1996). Cognition, memory, and dissociation. Child and Adolescent Clinics of North America, 5(2), 509–536.
Siegel, D. J. (1999). Developing mind: Toward a neurobiology of interpersonal experience. New York: Guilford.
Siegel, D. J. (n.d.). The developing mind: How relationships and the brain interact to shape who we are (2nd ed.). New York, NY: The Guilford
Press.
Sieswerda, S., Arntz, A., Mertens, I., & Vertommen, S. (2006). Hypervigilance in patients with borderline personality disorder: Specificity,
automaticity, and predictors. Behaviour Research and Therapy, 45, 1011–1024. doi:10.1016/j.brat.2006.07.012
Silberman, E. K., & Weingartner, H. (1986). Hemispheric lateralization of functions related to emotion. Brain and Cognition, 5(3), 322–
353.
Silbersweig, D. (2013). Default mode subnetworks, connectivity, depression and its treatment: Toward brain-based biomarker development.
Biological Psychiatry, 74(1), 5-6. doi:10.1016/j.biopsych.2013.05.011
Simon, O., Mangin, J., Cohen, L., Le Bihan, D., & Dehaene, S. (2002). Topographical layout of hand, eye, calculation, and language-
related areas in the human parietal lobe. Neuron, 33(3), 475–487.
Simpson, J. R., Drevets, W. C., Snyder, A. Z., Gusnard, D. A., & Raichle, M. E. (2001). Emotion-induced changes in human medial
prefrontal cortex: II. During anticipatory anxiety. Proceedings of the National Academy of Sciences, USA, 98(2), 688–693.

285
Simpson, J., & Overall, N. (2014). Partner buffering of attachment insecurity. Current Directions in Psychological Science, 23(1), 54–59.
doi:10.1177/0963721413510933
Simpson, J. R., Snyder, A. Z., Gusnard, D. A., & Raichle, M. E. (2001). Emotion-induced changes in human medial prefrontal cortex: I.
During cognitive task performance. Proceedings of the National Academy of Sciences, USA, 98(2), 683–687.
Singer, T., Seymour, B., O’Doherty, J. P., Stephan, K. E., Dolan, R. J., & Frith, C. D. (2006). Empathic neural responses are modulated by
the perceived fairness of others. Nature, 439, 466–469.
Sirevaag, A. M., & Greenough, W. T. (1988). A multivariate statistical summary of synaptic plasticity measures in rats exposed to complex,
social and individual environments. Brain Research, 441(1–2), 386–392.
Sirigu, A., Daprati, E., Ciancia, S., Giraux, P., Nighoghossian, N., Posada, A., & Haggard, P. (2003). Altered awareness of voluntary
action after damage to the parietal cortex. Nature Neuroscience, 7(1), 80–84.
Sirigu, A., Duhamel, J., Coehn, L., Pillon, B., Dubois, B., & Agid, Y. (1996). The mental representation of hand movements after parietal
cortex damage. Science, 273(5281), 1564–1568.
Siviy, S. M., & Harrison, K. A. (2008). Effects of neonatal handling on play behavior and fear towards a predator odor in juvenile rats
(rattus norvegicus). Journal of Comparative Psychology, 122(1), 1–8.
Skuse, D. H., & Gallagher, L. (2008). Dopaminergic-neuropeptide interactions in the social brain. Trends in Cognitive Sciences, 13(1), 27–
35. doi:10.1016/j.tics.2008.09.007.
Smith, E., Salat, D., Jeng, J., Mccreary, C., Fischl, B., Schmahmann, J., . . . Greenberg, S. (2011). Correlations between MRI white matter
lesion location and executive function and episodic memory. Neurology, 76, 1492–1499.
Smith, S. D., & Bulman-Fleming, M. B. (2004). A hemispheric asymmetry for the unconscious perception of emotion. Brain and Cognition,
55(3), 452–457.
Smith, S. M., Fox, P. T., Miller, K. L., Glahn, D. C., Fox, P. M., Mackay, C. E., . . . Beckmann, C. F. (2009). Correspondence of the
brain’s functional architecture during activation and rest. Proceedings of the National Academy of Sciences, 106(31), 13040-13045.
doi:10.1073/pnas.0905267106
Smyser, C. D., Inder, T. E., Shimony, J. S., Hill, J. E., Degnan, A. J., Snyder, A. Z., & Neil, J. J. (2010). Longitudinal analysis of neural
network development in preterm infants. Cerebral Cortex, 20(12), 2852–2862. doi:10.1093/cercor/bhq035
Smyser, C. D., Snyder, A. Z., Shimony, J. S., Blazey, T. M., Inder, T. E., & Neil, J. J. (2013). Effects of white matter injury on resting state
fMRI measures in prematurely born infants. PLoS One, 8(7), e68098. doi:10.1371/journal.pone.0068098
Smythe, J. W., Rowe, W. B., & Meaney, M. J. (1994). Neonatal handling alters serotonin (5-HT) turnover and 5-HT2 receptor binding in
selected brain regions: Relationship to the handling effect on glucocorticoid receptor expression. Developmental Brain Research, 80(1–2), 183–
189.
Snyder, J., & Chatterjee, A. (2004). Spatial-temporal anisometries following right parietal damage. Neuropsychologia, 42(12), 1703–1708.
Snyder, L. H., Batista, A. P., & Andersen, R. A. (1997). Coding of intention in the posterior parietal cortex. Nature, 386(6621), 167–170.
Soloff, P. H., Meltzer, C. C., Becker, C., Greer, P. J., Kelly, T. M., & Constantine, D. (2003). Impulsivity and prefrontal hypometabolism
in borderline personality disorder. Psychiatry Research: Neuroimaging, 123, 153–163.
Soloff, P., Nutche, J., Goradia, D., & Diwadkar, V. (2008). Structural brain abnormalities in borderline personality disorder: A voxel-based
morphometry study. Psychiatry Research, 164(3), 223.
Solomon, Z. (1990). Back to the front: Recurrent exposure to combat stress and reactivation of posttraumatic stress disorder. In M. E. Wolf
& A. D. Mosnaim (Eds.), Posttraumatic stress disorder: Etiology, phenomenology, and treatment (pp. 114–125). Washington, DC: American
Psychiatric Press.
Somerville, L. H., Jones, R. M., & Casey, B. (2010). A time of change: Behavioral and neural correlates of adolescent sensitivity to
appetitive and aversive environmental cues. Brain and Cognition, 72(1), 124-133. doi:10.1016/j.bandc.2009.07.003
Sontheimer, H. (1995). Glial influences on neuronal signaling. Neuroscientist, 1(3), 123–126.
Spangler, G., & Grossman, K. E. (1993). Biobehavioral organization in securely and insecurely attached infants. Child Development, 64(4),
1439–1450.
Spangler, G., & Schieche, M. (1998). Emotional and adrenocortical responses of infants to the strange situation: The differential function
of emotional expression. International Journal of Behavioral Development, 22(4), 681–706.
Spear, L. P. (2000). The adolescent brain and age-related behavioral manifestations. Neuroscience and Biobehavioral Reviews, 24(4), 417–463.
Specter, M. (2001, July 23). Rethinking the brain. New Yorker, 42–53.
Spence, S. A., Farrow, T., Herford, A., Wilkinson, I. D., Zheng, Y., & Woodruff, P. (2001). Behavioural and functional anatomical
correlates of deceptions in humans. NeuroReport, 12, 2849–2853.
Sperry, R. W. (1968). Hemispheric deconnection and unity in conscious awareness. American Psychologist, 23, 723–733.
Sperry, R. W., Gazzaniga, M. S., & Bogen, J. E. (1969). Interhemispheric relationships: The neocortical commissures; syndromes of

286
hemisphere disconnection. In P. J. Vinken & G. W. Bruyn (Eds.), Handbook of clinical neurology (Vol. 4, pp. 273–290). Amsterdam: North
Holland.
Spitz, R. (1946). Hospitalism: A follow-up report on investigation described in volume I, 1945. Psychoanalytic Study of the Child, 2, 113–117.
Spreng, R. N., & Grady, C. L. (2010). Patterns of brain activity supporting autobiographical memory, prospection, and theory of mind, and
their relationship to the default mode network. Journal of Cognitive Neuroscience, 22(6), 1112–1123.
Spreng, R. N., Mar, R. A., & Kim, A. S. (2009). The common neural basis of autobiographical memory, prospection, navigation, theory of
mind, and the default mode: A quantitative meta-analysis. Journal of Cognitive Neuroscience, 21(3), 489-510. doi:10.1162/jocn.2008.21029
Squire, L. R. (1987). Memory and brain. New York: Oxford University Press.
Squire, L. R., & Zola-Morgan, S. (1991). The medial temporal lobe memory system. Science, 253(5026), 2380–2386.
Sridharan, D., Levitin, D., & Menon, V. (2008). A critical role for the right fronto-insular cortex in switching between central-executive
and default-mode networks. Proceedings of the National Academy of Sciences, 105(34), 12569–12574. doi:10.1073/pnas.0800005105
Sripada, R. K., King, A. P., Welsh, R. C., Garfinkel, S. N., Wang, X., Sripada, C. S., & Liberzon, I. (2012). Neural dysregulation in
posttraumatic stress disorder: Evidence for disrupted equilibrium between salience and default mode brain networks. Psychosomatic Medicine,
74(9), 904-911. doi:10.1097/psy.0b013e318273bf33
Staff, R. T., Murray, A. D., Deary, I. J., & Whalley, L. J. (2004). What provides cerebral reserve? Brain, 127(Pt. 5), 1191–1199.
Stahl, S. M. (2008). Stahl’s essential psychopharmacology: Neuroscientific basis and practical applications. New York: Cambridge University Press.
Stamatakis, A., Pondiki, S., Kitraki, E., Diamantopoulou, A., Panagiotaropoulos, T., Raftogianni, A., & Stylianopoulou, F. (2008). Effect
of neonatal handling on adult rat spatial learning and memory following acute stress. Stress, 11(2), 148–159.
Stanley, B., Sher, L., Wilson, S., Ekman, R., Huang, Y.-Y., & Mann, J. J. (2010). Non- suicidal self-injurious behavior, endogenous
opioids and monoamine transmitters. Journal of Affective Disorders, 124, 134–140.
St. Clair, M. (1986). Object relations and self psychology. Monterey, CA: Brooks/Cole.
Stein, M. B., Koverola, C., Hanna, C., Torchia, M. G., & McClarty, B. (1997). Hippocampal volume in women victimized by childhood
sexual abuse. Psychological Medicine, 27(4), 951–959.
Stennett, R. G. (1957). The relationship of performance level to level of arousal. Journal of Experimental Psychology, 54(1), 54–61.
Stephan, H., & Andy, O. J. (1977). Quantitative comparison of the amygdala in insectivores and primates. Acta Anatomica, 98(2), 130–153.
Stern, D. N. (1985). The interpersonal world of the infant. New York: Basic Books.
Stern, D. N. (1995). The motherhood constellation. New York: Basic Books.
Stern, E. R., Fitzgerald, K. D., Welsh, R. C., Abelson, J. L., & Taylor, S. F. (2012). Resting-state functional connectivity between fronto-
parietal and default mode networks in obsessive-compulsive disorder. PLoS ONE, 7(5). doi:10.1371/journal.pone.0036356
Stern, Y. (2002). What is cognitive reserve? Theory and research application of the reserve concept. Journal of the International
Neuropsychological Society, 8, 448–460.
Stern, Y., Alexander, G. E., Prohovnik, I., & Mayeux, R. (1992). Inverse relationship between education and parietotemporal perfusion
deficit in Alzheimer’s disease. Annals of Neurology, 32(3), 371–375.
Stern, Y., Alexander, G. E., Prohovnik, I., Stricks, L., Link, B., Lennon, M. C., & Maveux, R. (1995). Relationship between lifetime
occupation and parietal flow: Implications for a reserve against Alzheimer’s disease pathology. Neurology, 45(1), 55–60.
Stern, Y., Habeck, C., Moeller, J., Scarmeas, N., Anderson, K. E., Hilton, H. J., . . . van Heertum, R. (2005). Brain networks associated
with cognitive reserve in healthy young and old adults. Cerebral Cortex, 15(4), 394–402.
Sterr, A., Muller, M. M., Elbert, T., Rockstroh, B., Pantev, C., & Taub, E. (1998a). Perceptual correlates of changes in cortical
representation of fingers in blind multifinger Braille readers. Journal of Neuroscience, 18(11), 4417–4423.
Sterr, A., Muller, M. M., Elbert, T., Rockstroh, B., Pantev, C., & Taub, E. (1998b). Changed perceptions in Braille readers. Nature,
391(6663), 134–135.
Stiles, J. (2000). Neural plasticity and cognitive development. Developmental Neuropsychology, 18(2), 237–272.
Stoessel, C., Stiller, J., Bleich, S., Boensch, D., Doerfler, A., Garcia, M., . . . Forster, C. (2011). Differences and similarities on neuronal
activities of people being happily and unhappily in love: A functional magnetic resonance imaging study. Neuropsychobiology, 64(1), 52-60.
Stolorow, R. D., & Atwood, G. E. (1979). Faces in a cloud: Subjectivity in psychoanalytic theory. New York: Jason Aronson.
Stranahan, A. M., Khalil, D., & Gould, E. (2006). Social isolation delays the positive effects of running on adult neurogenesis. Nature
Neuroscience, 9(4), 526–533.
Strange, B. A., & Dolan, R. J. (2004). Beta-adrenergic modulation of emotional memory-evoked human amygdala and hippocampal
responses. Proceeding of the National Academy of Science, USA, 101(31), 11454–11458.
Straub, J., Plener, P., Sproeber, N., Sprenger, L., Koelch, M., Groen, G., & Abler, B. (2015). Neural correlates of successful psychotherapy
of depression in adolescents. Journal of Affective Disorders, 183, 239–246. doi:10.1016/j.jad.2015.05.020
Straube, T., Glauer, M., Dilger, S., Mentzel, H., & Miltner, W. H. R. (2006). Effects of cognitive-behavioral therapy on brain activation in

287
specific phobia. NeuroImage, 29(1), 125–135.
Stuss, D. T., Gallup, G. G., & Alexander, M. P. (2001). The frontal lobes are necessary for “theory of mind.” Brain, 124(2), 279–286.
Sullivan, R. M., Wilson, D. A., & Leon, M. (1989). Norepinephrine and learning-induced plasticity in infant rat olfactory system. Journal of
Neuroscience, 9(11), 3998–4006.
Sulloway, F. J. (1979). Freud: Biologist of the mind. New York: Basic Books.
Sun, L., Jin, Z., Zang, Y. F., Zeng, Y. W., Liu, G., Li, Y., . . . Wang, Y.-F. (2005). Differences between attention-deficit disorder with and
without hyperactivity: A 1H-magnetic spectroscopy study. Brain Development, 27, 340–344.
Suor, J. H., Sturge-Apple, M. L., Davies, P. T., Cicchetti, D., & Manning, L. G. (2015). Tracing differential pathways of risk:
Associations among family adversity, cortisol, and cognitive functioning in childhood. Child Development, 86(4), 1142–1158.
doi:10.1111/cdev.12376
Supekar, K., Uddin, L. Q., Prater, K., Amin, H., Greicius, M. D., & Menon, V. (2010). Development of functional and structural
connectivity within the default mode network in young children. NeuroImage, 52(1), 290-301. doi:10.1016/j.neuroimage.2010.04.009
Svensson, T. H. (1987). Peripheral, autonomic regulation of locus coeruleus noradrenergic neurons in the brain: Putative implications for
psychiatry and psychopharmacology. Psychopharmacology, 92, 1–7.
Swirsky-Sacchetti, T., Gorton, G., Samuel, S., Sobel, R., Genetta-Wadley, A., & Burleigh, B. (1993). Neuropsychological function in
borderline personality disorder. Journal of Clinical Psychology, 49(3), 385–396.
Szyf, M., McGowan, P., & Meaney, M. J. (2008). The social environment and epigenome. Environmental and Molecular Mutagenesis, 49(1),
46–60.
Szyf, M., Weaver, I. C. G., Champagne, F. A., Dioro, J., & Meaney, M. J. (2005). Maternal programming of steroid receptor expression
and phenotype through DNA methylation in the rat. Frontiers in Neuroendocrinology, 26(3–4), 139–162.
Szyf, M., Weaver, I., & Meaney, M. (2007). Maternal care, the epigenome and phenotypic differences in behavior. Reproductive Toxicology,
24(1), 9–19.
Taber, K. H., & Hurley, R. A. (2008). Astroglia: Not just glue. Journal of Neuropsychiatry and Clinical Neurosciences, 20(2), 124–129.
Takahashi, T., Chanen, A. M., Wood, S. J., Walterfang, M., Harding, I. H., Yücel, M., . . . Pantelis, C. (2009). Midline brain structures in
teenagers with first-presentation borderline personality disorder. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 33, 842–846.
Takahashi, T., Ikeda, K., Ishikawa, M., Tsukasaki, T., Nakama, D., Tanida, S., & Kameda, T. (2004). Social stress-induced cortisol
elevation acutely impairs social memory in humans. Neuroscience Letters, 363(2), 125–130.
Takeda, Y., Kurita, T., Sakurai, K., Shiga, T., Tamaki, N., & Koyama, T. (2011). Persistent déjà vu associated with hyperperfusion in the
entorhinal cortex. Epilepsy and Behavior, 21(2), 196–199. doi:10.1016/j.yebeh.2011.03.031
Takeuchi, H., Taki, Y., Sassa, Y., Hashizume, H., Sekiguchi, A., Fukushima, A., & Kawashima, R. (2012). Brain structures associated with
executive functions during everyday events in a non-clinical sample. Brain Structure and Function, 218, 1017–1032. doi:10.1007/s00429-012-
0444-z
Takiguchi, S., Fujisawa, T. X., Mizushima, S., Saito, D. N., Okamoto, Y., Shimada, K., . . . Tomoda, A. (2015). Ventral striatum
dysfunction in children and adolescents with reactive attachment disorder: Functional MRI study. British Journal of Psychiatry Open, 1(2), 121–
128. doi:10.1192/bjpo.bp.115.001586
Tamietto, M., Geminiani, G., Genero, R., & De Gelder, B. (2007). Seeing fearful body language overcomes attentional deficits in patients
with neglect. Journal of Cognitive Neuroscience, 19(3), 445–454.
Tamm, L., Menon, V., & Reiss, A. L. (2006). Parietal attentional system aberrations during target detection in adolescents with attention
deficit hyperactivity disorder: Event-related fMRI evidence. American Journal of Psychiatry, 163(6), 1033–1043.
Tamm, L., Menon, V., Ringel, J., & Reiss, A. L. (2004). Event-related fMRI evidence of frontotemporal involvement in aberrant response
inhibition and task switching in attention-deficit/hyperactivity disorder. Journal of the American Academy of Child and Adolescent Psychiatry,
43(11), 1430–1440.
Tang, Y. P., Shimizu, E., Dube, G. R., Rampon, C., Kerchner, G. A., Zhuo, M., . . . Tsien, J. Z. (1999). Genetic enhancement of learning
and memory in mice. Nature, 401(6748), 63–69.
Tanji, J., & Hoshi, E. (2001). Behavioral planning in the prefrontal cortex. Current Opinion in Neurobiology, 11(2), 164–170.
Tankersley, D., Stowe, J., & Huettel, S. A. (2007). Altruism is associated with an increased neural response to agency. Nature Neuroscience,
1–2. doi:10.1038/nn1833
Taylor, G. J. (2000). Recent developments in alexithymia theory and research. Canadia Journal of Psychiatry, 45(2), 134–142.
Taylor, J. (2001). The central role of the parietal lobes in consciousness. Consciousness and Cognition, 10(3), 379–417.
Taylor, M. A. (1999). The fundamentals of clinical neuropsychiatry. Oxford: Oxford University Press.
Taylor, S. E., & Brown, J. D. (1988). Illusion and well-being: A social psychological perspective on mental health. Psychological Bulletin,
103(2), 193–210.

288
Teasdale, J. D., Howard, R. J., Cox, S. G., Ha, Y., Brammer, M. J., Williams, S. C. R., & Checkley, S. A. (1999). Functional MRI study of
the cognitive generation of affect. American Journal of Psychiatry, 156(2), 209–215.
Tebartz van Elst, L., Hesslinger, B., Thiel, T., Geiger, E., Haegele, K., Lemieux, L., . . . Ebert, D. (2003). Frontolimbic brain
abnormalities in patients with borderline personality disorder: A volumetric magnetic resonance imaging study. Biological Psychiatry, 54, 163–
171.
Tebartz van Elst, L., Thiel, T., Hesslinger, B., Leib, K., Bohus, M., Hennig, J., & Ebert, D. (2001). Subtle prefrontal neuropathology in a
pilot magnetic resonance spectroscopy study in patients with borderline personality disorder. Journal of Neuropsychiatry and Clinical Neurosciences,
13, 511–514.
Teicher, M. H., Anderson, C. M., & Polcari, A. (2012). Childhood maltreatment is associated with reduced volume in the hippocampal
subfields CA3, dentate gyrus, and subiculum. Proceedings of the National Academy of Sciences, 109(9), E563–E572. doi:10.1073/pnas.1115396109
Teicher, M. H., Dumont, N. L., Ito, Y., Vaituzis, C., Geidd, J. N., & Andersen, S. L. (2004). Childhood neglect is associated with reduced
corpus callosum area. Biological Psychiatry, 56(2), 80–85.
Teicher, M. H., Ito, Y., Glod, C. A., Andersen, S. L., Dumont, N., & Ackerman, E. (1997). Preliminary evidence for abnormal cortical
development in physically and sexually abused children using EEG coherence and MRI. Annals of the New York Academy of Sciences, 821, 160–
175.
Tejedor-Real, P., Costela, C., & Gibert-Rahola, J. (1998). Neonatal handling reduces emotional reactivity and susceptibility to learned
helplessness: Involvement of catecholaminergic systems. Life Sciences, 62(1), 37–50.
ten Cate, C. (1989). Behavioral development: Toward understanding processes. In P. P. G. Bateson & P. Klopfer (Eds.), Perspectives in
ethology (Vol. 8, pp. 243–269). New York: Plenum.
Teneback, C. C., Nahas, Z., Speer, A. M., Molloy, M., Stallings, L. E., Spicer, K. M., Risch, S. C., & George, M. S. (1999). Changes in
prefrontal cortex and paralimbic activity in depression following two weeks of daily left prefrontal TMS. Journal of Neuropsychiatry and Clinical
Neurosciences, 11(4), 426–435.
Thatcher, R. W., Walker, R. A., & Giudice, S. (1987). Human cerebral hemispheres develop at different rates and ages. Science, 236(4805),
1110–1113.
Thayer, J. F., & Cohen, B. H. (1985). Differential hemispheric lateralization for positive and negative emotion: An electromyographic
study. Biological Psychology, 21, 265–266.
Thomas, K. S. (1987). General practice consultation: Is there any point in being positive? British Medical Journal, 294, 1200–1202.
Thomason, M. E., Chang, C. E., Glover, G. H., Gabrieli, J. D., Greicius, M. D., & Gotlib, I. H. (2008). Default-mode function and task-
induced deactivation have overlapping brain substrates in children. NeuroImage, 41(4), 1493-1503. doi:10.1016/j.neuroimage.2008.03.029
Thome, J., Frewen, P., Daniels, J. K., Densmore, M., & Lanius, R. A. (2014). Altered connectivity within the salience network during
direct eye gaze in PTSD. Borderline Personality Disorder and Emotion Dysregulation, 1(1), 17. doi:10.1186/2051-6673-1-17
Thompson, R. A., Lamb, M. E., & Estes, D. (1982). Stability of infant-mother attachment and its relationship to changing life
circumstances in an unselected middle-class sample. Child Development, 53(1), 144–148.
Tillfors, M., Furmark, T., Marteinsdottir, I., Pissota, A., Langstrom, B., & Fredrikson, M. (2001). Cerebral blood flow in subjects with
social phobia during stressful speaking tasks: A PET study. American Journal of Psychiatry, 158(8), 1220–1226.
Tillich, P. (1974). The courage to be. New Haven: Yale University Press.
Todd, R., Ehlers, M., Müller, D., Robertson, A., Palombo, D., Freeman, N., . . . Anderson, A. (2015). Neurogenetic variations in
norepinephrine availability enhance perceptual vividness. Journal of Neuroscience, 35(16), 6506–6516. doi:10.1523/jneurosci.4489- 14.2015
Tomarken, A. J., & Davidson, R. J. (1994). Frontal brain activation in repressors and nonrepressors. Journal of Abnormal Psychology, 103(2),
339–349.
Tomarken, A. J., Davidson, R. J., Wheeler, R. E., & Doss, R. C. (1992). Individual differences in anterior brain asymmetry and
fundamental dimensions of emotion. Journal of Personality and Social Psychology, 62(4), 676–687.
Tomoda, A., Navalta, C. P., Polcari, A., Sadato, N., & Teicher, M. H. (2009). Childhood sexual abuse is associated with reduced gray
matter volume in visual cortex of young women. Biological Psychiatry, 66(7), 642–648. doi:10.1016/j.biopsych.2009.04.021
Tomoda, A., Polcari, A., Anderson, C. M., & Teicher, M. H. (2012). Reduced visual cortex gray matter volume and thickness in young
adults who witnessed domestic violence during childhood. PLoS One, 7(12), 1–11. doi:10.1371/journal.pone.0052528
Torasdotter, M., Metsis, M., Henriksson, B. G., Winblad, B., & Mohammed, A. H. (1998). Environmental enrichment results in higher
levels of nerve growth factor MRNA in the rat visual cortex and hippocampus. Behavioral Brain Research, 93(1–2), 83–90.
Tottenham, N., Hare, T. A., Quinn, B. T., Mccarry, T. W., Nurse, M., Gilhooly, T., . . . Casey, B. (2010). Prolonged institutional rearing
is associated with atypically large amygdala volume and difficulties in emotion regulation. Developmental Science, 13(1), 46–61.
doi:10.1111/j.1467-7687.2009.00852.x
Tottenham, N., Hare, T., Millner, A., Gilhooly, T., Zevin, J., & Casey, B. (2011). Elevated amygdala response to faces following early

289
deprivation. Developmental Science, 14(2), 190–204. doi:10.1111/j.1467-7687.2010.00971.x
Trapolini, T., Ungerer, J. A., & McMahon, C. A. (2008). Maternal depression: Relations with maternal caregiving representations and
emotional availability during the preschool years. Attachment and Human Development, 10(1), 73–90.
Travis, L. A., Bliwise, N. G., Binder, J. L., & Horne-Moyer, H. L. (2001). Changes in clients’ attachment styles over the course of time-
limited dynamic psychotherapy. Psychotherapy: Theory, Research, Practice, Training, 38(2), 149–159.
Tremblay, L., & Schultz, W. (1999). Relative reward preference in primate orbitofrontal cortex. Nature, 398, 704–708.
Trevarthen, C. (1993). The self born in intersubjectivity: The psychology of an infant communicating. In U. Neisser (Ed.), The perceived self:
Ecological and interpersonal sources of self-knowledge (pp. 121–173). Cambridge: Cambridge University Press.
Triggs, W. J., McCoy, K. J., Greer, R., Rossi, F., Bowers, D., Kortenkamp, S., . . . Goodman, W. K. (1999). Effects of left frontal
transcranial magnetic stimulation on depressed mood, cognition, and corticomotor threshold. Biological Psychiatry, 45(11), 1440–1446.
Trojan, S., & Pokorny, J. (1999). Theoretical aspects of neuroplasticity. Physiological Research, 48(2), 87–97.
Tsoory, M. M., Vouimba, R. M., Akirav, I., Kavushansky, A., Avital, A., & Richer-Levin, G. (2008). Amygdala modulation of memory-
related processes in the hippocampus: Potential relevance to PTSD. Progress in Brain Research, 167, 35–49.
Tucker, D. M. (1992). Developing emotions and cortical networks. In M. R. Gunnar & C. Nelson (Eds.), Minnesota symposia on child
psychology: Vol. 24. Developmental behavioral neuroscience (pp. 75–128). Hillsdale, NJ: Erlbaum.
Tucker, D. M., Luu, P., & Pribram, K. H. (1995). Social and emotional self-regulation. In J. Grafman & K. J. Hoyoak (Eds.), Structure and
functions of the human prefrontal cortex (pp. 213–239). New York: New York Academy of Sciences.
Tulving, E. (1985). How many memory systems are there? American Psychologist, 40(4), 385–398.
Turke, P. W. (1997). Hypothesis: Menopause discourages infanticide and encourages continued investment by agnates. Evolution and
Human Behavior, 18(1), 3–13.
Tyrka, A. R., Wier, L., Price, L. H., Ross, N., Anderson, G. M., Wilkinson, C. W., & Carpenter, L. (2008). Childhood parental loss and
adult hypothalamic-pituitary-adrenal function. Biological Psychiatry, 63(12), 1147–1154.
Uddin, L. (2014). Salience processing and insular cortical function and dysfunction. Nature Reviews Neuroscience, 16, 55–61.
doi:10.1038/nrn3857
Uddin, L. Q., Kaplan, J. T., Molnar-Szakacs, I., Zaidel, E., & Iacoboni, M. (2005). Self-face recognition activates a frontoparietal “mirror”
network in the right hemisphere: An event-related fMRI study. Neuroimage, 25(3), 926–935.
Ulfig, N., Setzer, M., & Bohl, J. (2003). Ontogeny of the human amygdala. Annals of the New York Academy of Sciences, 985, 22–33.
Ulrich, R. (1984). View through a window may influence recovery from surgery. Science, 224(4647), 420–421.
Umiltà, M. A., Wood, R., Loffredo, F., Ravera, R., & Gallese, V. (2013). Impact of civil war on emotion recognition: The denial of sadness
in Sierra Leone. Frontiers in Psychology, 4. doi:10.3389/fpsyg.2013.00523
Ungerleider, L. G. (1995). Functional brain imaging studies of cortical mechanisms for memory. Science, 270(5237), 769–775.
Ungerleider, L. G., & Haxby, J. V. (1994). “What” and “where” in the human brain. Current Opinion in Neurobiology, 4(2), 157–165.
Unoka, Z., Fogd, D., Füzy, M., & Csukly, G. (2011). Misreading the facial signs: Specific impairments and error patterns in recognition of
facial emotions with negative valence in borderline personality disorder. Psychiatric Research, 189, 419–425.
Urry, H. L., Nitschke, J. B., Dolski, I., Jackson, D. C., Dalton, K. M., Mueller, C. J., . . . Davidson, R. J. (2004). Making a life worth
living: Neural correlates of well-being. Psychological Science, 15(6), 367–372.
Urry, H. L., van Reekum, C. M., Johnstone, T., Kalin, N. H., Thurow, M. E., & Schaefer, H. S., . . . Davidson, R. J. (2006). Amygdala
and ventromedial prefrontal cortex are inversely coupled during regulation of negative affect and predict the diurnal pattern of cortisol secretion
among older adults. Journal of Neuroscience, 26(16), 4415–4425.
Usdin, E., Kvetnansky, R., & Kopin, I. J. (1976). Stress and catecholamines. Oxford: Pergamon.
Vaidya, C. J., Austin, G., Kirkorian, G., Ridlehuber, H. W., Desmond, J. E., Glover, G. H., & Beasel, A. F. (1998). Selective effects of
methyphenidate in attention deficit hyperactivity disorder: A functional magnetic resonance study. Proceedings of the National Academy of Sciences,
USA, 95(24), 14494–14499.
Vaidya, C. J., Bunge, S. A., Dudukovic, N. M., Zalecki, C. A., Elliott, G. R., & Gabrieli, J. D. (2005). Altered neural substrates of
cognitive control in childhood ADHD: Evidence from functional magnetic resonance imaging. American Journal of Psychiatry, 162(9), 1605–
1613. doi:10.1176/appi.ajp.162.9.1605
Vaina, L. M., Solomon, J., Chowdhury, S., Sinha, P., & Belliveau, J. W. (2001). Functional neuroanatomy of biological motion perception
in humans. Proceedings of the National Academy of Sciences, USA, 98(20), 11656–11661.
Vallar, G., Sterzi, R., Bottini, G., Cappa, S., & Rusconi, M. L. (1990). Temporary remission of left hemianesthesia after vestibular
stimulation: A sensory neglect phenomenon. Cortex, 26(1), 123–131.
Vallée, M., Maccari, S., Dellu, F., Simon, H., Le Moal, M., & Mayo, W. (1999). Long-term effects of prenatal stress and postnatal
handling on age-related glucocorticoid secretion and cognitive performance: A longitudinal study in the rat. European Journal of Neuroscience,

290
11(18), 2906–2916.
Vallée, M., Mayo, W., Dellu, F., Le Moal, M., Simon, H., & Maccari, S. (1997). Prenatal stress induces high anxiety and postnatal
handling induces low anxiety in adult offspring: Correlation with stress-induced corticosterone secretion. Journal of Neuroscience, 17(7), 2626–
2636.
van den Heuvel, M., Mandl, R., Kahn, R., & Pol, H. (2009). Functionally linked resting-state networks reflect the underlying structural
connectivity architecture of the human brain. Human Brain Mapping, 30(10), 3127–3141. doi:10.1002/hbm.20737
van der Kolk, B. A. (1988). The trauma spectrum: The interaction of biological and social events in the genesis of the trauma response.
Journal of Traumatic Stress, 1(3), 273–290.
van der Kolk, B. A. (1994). The body keeps the score: Memory and the evolving psychobiology of post traumatic stress. Harvard Review of
Psychiatry, 1(5), 253–265.
van der Kolk, B. A., Blitz, R., Burr, W., Sherry, S., & Hartmann, E. (1984). Nightmares and trauma: A comparison of nightmares after
combat with lifelong nightmares in veterans. American Journal of Psychiatry, 141(2), 187–190.
van der Kolk, B. A., & Greenberg, M. S. (1987). The psychobiology of the traumatic response: Hyperarousal, constriction, and addiction to
traumatic reexposure. In B. A. van der Kolk (Ed.), Psychological trauma (pp. 63–87). Washington, DC: American Psychiatric Press.
van der Kolk, B. A., Pelcovitz, D., Roth, S., Mandel, F. S., McFarlane, A., & Herman, J. L. (1996). Dissociation, somatization, and affect
dysregulation: The complexity of adaptation to trauma. American Journal of Psychiatry, 153(7 Suppl.), 83–95.
Vanduffel, W., Fize, D., Peuskens, H., Denys, K., Sunaert, S., Todd, J. T., & Orban, G. A. (2002). Extracting 3D from motion:
Differences in human and monkey intraparietal cortex. Science, 298(5592), 413–415.
van Harmelen, A., Van Tol, M., Van der Wee, N. J., Veltman, D. J., Aleman, A., Spinhoven, P., . . . Elzinga, B. M. (2010). Reduced
medial prefrontal cortex volume in adults reporting childhood emotional maltreatment. Biological Psychiatry, 68(9), 832–838.
doi:10.1016/j.biopsych.2010.06.011
van Hoesen, G. W. (1981). The differential distribution, diversity and sprouting of cortical projections to the amygdala in the rhesus
monkey. In Y. Ben-Ari (Ed.), The amygdaloid complex (pp. 77–90). Amsterdam: Elsevier/North Holland Biomedical Press.
Van Opstal, F., Verguts, G., & Fias, W. (2008). A hippocampal-parietal network for learning an ordered sequence. NeuroImage, 40(1), 333–
341.
van Reekum, C., Urry, H., Johnson, T., Throw, M., Frye, C., Jackson, C., . . . Davidson, R. (2007). Individual differences in amygdala and
ventromedial prefrontal cortex activity are associated with evaluation speed and psychological well-being. Journal of Cognitive Neuroscience, 19(2),
237–248.
van Reekum, R., Conway, C. A., Gansler, D., White, R., & Bachman, D. L. (1993). Neurobehavioral study of borderline personality
disorder. Journal of Psychiatry and Neuroscience, 18(3), 121–129.
Vansteenwegen, D., Hermans, D., Vervliet, B., Francken, G., Beckers, T., Baeyens, F., & Eelen, P. (2005). Return of fear in a human
differential conditioning paradigm caused by a return to the original acquisition context. Behavior Research and Therapy, 43(3), 323–326.
Vasterling, J. J., Brailey, K., Constans, J. I., & Sutker, P. B. (1998). Attention and memory dysfunction in posttraumatic stress disorder.
Neuropsychology, 12(1), 125–133.
Vaughn, B., Egeland, B., Sroufe, L. A., & Waters, E. (1979). Individual differences in infant-mother attachment at twelve and eighteen
months: Stability and change in families under stress. Child Development, 50(4), 971–975.
Vernadakis, A. (1996). Glia-neuron intercommunications and synaptic plasticity. Progressive Neurobiology, 49(3), 185–214.
Victor, M., & Ropper, A. H. (2001). Adams and Victor’s principles of neurology (7th ed.). New York: McGraw-Hill.
Villalba, R., & Harrington, C. (2003). Repetitive self-injurious behavior: The emerging potential of psychotropic intervention. Psychiatric
Times, 20, 1–11.
Villarreal, G., Hamilton, D. A., Petropoulos, H., Driscoll, I., Rowland, L. M., Griego, J. A., . . . Brooks, W. M. (2002). Reduced
hippocampal volume and total white matter volume in posttraumatic stress disorder. Biological Psychiatry, 52(2), 119–125.
Vogeley, K., May, M., Ritzl, A., Falkai, P., Zilles, K., & Fink, G. R. (2004). Neural correlates of first-person perspective as one constituent
of human self-consciousness. Journal of Cognitive Neuroscience, 16(5), 817–827.
Vogt, B. A. (2005). Pain and emotion interactions in subregions of the cingulate gyrus. Nature Reviews Neuroscience, 6(7), 533–544.
von Bonin, G. (1963). The evolution of the human brain. Chicago: University of Chicago Press.
Vondra, J. I., Shaw, D. S., Swearingen, L., Cohen, M., & Owens, E. B. (2001). Attachment stability and emotional and behavioral
regulation from infancy to preschool age. Development and Psychopathology, 13(1), 13–33.
Vyas, A., Bernal, S., & Chattarji, S. (2003). Effects of chronic stress on dendritic arborization in the central and extended amygdala. Brain
Research, 965(1–2), 290–294.
Vyas, A., & Chattarji, S. (2004). Modulation of different states of anxiety-like behavior by chronic stress. Behavioral Neuroscience, 118(6),
1450–1454.

291
Wada, J. (1961). Modification of cortically induced responses in brain stem by shift of attention in monkeys. Science, 133(3445), 40–42.
Wager, T. D., Davidson, M. L., Hughes, B. L., Lindquist, M. A., & Ochsner, K. N. (2008). Prefrontal-subcortical pathways mediating
successful emotion regulation. Neuron, 59(6), 1037–1050.
Wager, T., Rilling, J. K., Smith, E. E., Sokolik, A., Casey, K. L., Davidson, R. J., . . . Cohen, J. D. (2004). Placebo-induced changes in
fMRI in the anticipation and experience of pain. Science, 303(5661), 1162–1167.
Wagner, D. D., Haxby, J. V., & Heatherton, T. F. (2012). The representation of self and person knowledge in the medial prefrontal cortex.
Wiley Interdisciplinary Reviews: Cognitive Science, 3(4), 451–470. doi:10.1002/wcs.1183
Walach, H., & Jonas, W. B. (2004). Placebo research: The evidence base for harnessing self-healing capacities. Journal of Alternative and
Complementary Medicine, 10(1), 103–112.
Walker, D., Miles, L., & Davis, M. (2009). Selective participation of the bed nucleus of the stria terminalis and CRF in sustained anxiety-
like versus phasic fear-like responses. Progress in Neuro-Psychopharmacology and Biological Psychiatry, 33(8), 1291-1308.
doi:10.1016/j.pnpbp.2009.06.022
Walsh, R. N., Budtz-Olsen, O. E., Penny, J. E., & Cummins, R. A. (1969). The effects of environmental complexity of the histology of the
rat hippocampus. Journal of Comparative Neurology, 137, 361–366.
Walsh, V., Ashbridge, E., & Cowey, A. (1998). Cortical plasticity in perceptual learning demonstrated by transcranial magnetic stimulation.
Neuropsychologia, 36(4), 45–49.
Walterfang, M., & Velakoulis, D. (2005). Cortical release signs in psychiatry. Australian and New Zealand Journal of Psychiatry, 39(5), 317–
327.
Walton, M., Bannerman, D., Alterescu, K., & Rushworth, M. (2003). Functional specialization within medial frontal cortex of the anterior
cingulate for evaluating effort-related decisions. Journal of Neuroscience, 23(16), 6475–6479.
Wang, S., Tudusciuc, O., Mamelak, A. N., Ross, I. B., Adolphs, R., & Rutishauser, U. (2014). Neurons in the human amygdala selective
for perceived emotion. Proceedings of the National Academy of Sciences, 111(30), E3110–E3119. doi:10.1073/pnas.1323342111
Wanisch, K., Tang, J., Mederer, A., & Wotjak, C. T. (2005). Trace fear conditioning depends on NMDA receptor activation and protein
synthesis within the dorsal hippocampus of mice. Behavioral Brain Research, 157(1), 63–69.
Ward, A. M., Schultz, A. P., Huijbers, W., Dijk, K. R., Hedden, T., & Sperling, R. A. (2013). The parahippocampal gyrus links the
default-mode cortical network with the medial temporal lobe memory system. Human Brain Mapping Hum. Brain Mapp, 35(3), 1061-1073.
doi:10.1002/hbm.22234
Warneken, F., Hare, B., Melis, A. P., Hanus, D., & Tomasello, M. (2007). Spontaneous altruism by chimpanzees and young children.
PLoS Biology, 5(7), 1414–1420. doi:10.1371/journal.pbio.0050184
Warner-Schmidt, J. L., & Duman, R. S. (2006). Hippocampal neurogenesis: Opposing effects of stress and antidepressant treatment.
Hippocampus, 16(3), 239–249.
Watanabe, M. (1996). Reward expectancy in primate prefrontal neurons. Nature, 382, 629–632.
Watanabe, Y., Gould, E., Daniels, D. C., Cameron, H., & McEwen, B. S. (1992). Tianeptine attenuates stress-induced morphological
changes in the hippocampus. European Journal of Pharmacology, 222(1), 157–162.
Watanabe, Y. E., Gould, E., & McEwen, B. S. (1992). Stress induced atrophy of apical dendrites of hippocampal CA3 pyramidal neurons.
Brain Research, 588(2), 341–345.
Waters, E., Merrick, S., Treboux, D., Crowell, J., & Albersheim, L. (2000). Attachment security in infancy and early adulthood: A twenty-
year longitudinal study. Child Development, 71(2), 684–689.
Watson, K. K., Matthews, B. J., & Allman, J. M. (2007). Brain activation during sight gags and language dependent humor. Cerebral
Cortex, 17(2), 314–324.
Watson, S., Chilton, R., Fairchild, H., & Whewell, P. (2006). Association between childhood trauma and dissociation among patients with
borderline personality disorder. Australian and New Zealand Journal of Psychiatry, 40(5), 478–481.
Weaver, I. C. G., Cervoni, N., Champagne, F. A., D’Alessio, A. C., Sharma, S., Seckl, J. R., . . . Meaney, M. J. (2004). Epigenetic
programming by maternal behavior. Nature Neuroscience, 7(8), 847–854.
Weaver, I. C., Champagne, F. A., Brown, S. E., Dymov, S., Sharma, S., Meaney, M. J., & Szyf, M. (2005). Reversal of maternal
programming of stress response in adult offspring through methyl supplementation: Altering epigenetic marking later in life. Journal of
Neuroscience, 25(47), 11045–11054.
Weaver, I. C. G., D’Alessio, A. C., Brown, S. E., Hellstrom, I. C., Dymov, S., Sharma, S., . . . Meaney, M. J. (2007). The transcription
factor nerve growth factor-inducible protein a mediates epigenetic programming: Altering epigenetic marks by immediate-early genes. Journal of
Neuroscience, 27(7), 1756–1768.
Weaver, I. C. G., Grant, R. J., & Meaney, M. J. (2002). Maternal behavior regulates long-term hippocampal expression of BAX and
apoptosis in the offspring. Journal of Neurochemistry, 82(4), 998–1002.

292
Weaver, I. C. G., Meaney, M. J., & Szyf, M. (2006). Maternal care effects on the hippocampal transcriptome and anxiety-mediated
behaviors in the offspring that are reversible in adulthood. Proceedings of the National Academy of Sciences, USA, 103(9), 3480–3485.
Weaver, S. A., Aherne, F. X., Meaney, M. J., Schaefer, A. L., & Dixon W. T. (2000). Neonatal handling permanently alters hypothalamic-
pituitary-adrenal axis function, behavior, and body weight in boars. Journal of Endocrinology, 164, 349–359.
Weiner, I. (1998). Principles of psychotherapy. New York: Wiley and Sons.
Weinfield, N. S., Sroufe, L. A., & Egeland, B. (2000). Attachment from infancy to young adulthood in a high-risk sample: Continuity,
discontinuity, and their correlates. Child Development, 71(3), 695–702.
Weingarten, S. M., Cherlow, D. G., & Holmgren, E. (1977). The relationship of hallucinations to the depth structures of the temporal
lobe. Acta Neurochirurgica, 24(Suppl.), 199–216.
Weissman-Fogel, I., Moayedi, M., Taylor, K. S., Pope, G., & Davis, K. D. (2010). Cognitive and default-mode resting state networks: Do
male and female brains “rest” differently? Human Brain Mapping Hum. Brain Mapp. doi:10.1002/hbm.20968
Wellington, N., & Rieder, M. J. (1993). Attitudes and practices regarding analgesia for newborn circumcision. Pediatrics, 92(4), 541–543.
Weniger, G., Lange, C., Sachsse, U., & Irle, E. (2009). Reduced amygdala and hippocampus size in trauma-exposed women with
borderline personality disorder and without posttraumatic stress disorder. Journal of Psychiatry and Neuroscience, 34(5), 383–388.
West, M. J. (1993). Regionally specific loss of neurons in the aging human hippocampus. Neurobiology of Aging, 14(4), 287–293.
Wexler, B. E., & Heninger, G. R. (1979). Alterations in cerebral laterality during acute psychotic illness. Archives of General Psychiatry,
36(3), 278–284.
Whalen, P. J., Johnstone, T., Somerville, L. H., Nitschke, J. B., Polis, S., Alexander, A. L., . . . Kalin, N. H. (2008). A functional magnetic
resonance imaging predictor of treatment response to venlafaxine in generalized anxiety disorder. Biological Psychiatry, 63(9), 858–863.
Whalley, L. J., Deary, I. J., Appleton, C. L., & Starr, J. M. (2004). Cognitive reserve and the neurobiology of cognitive aging. Ageing
Research Reviews, 3(4), 369–382.
Whalley, M. G., Kroes, M. C., Huntley, Z., Rugg, M. D., Davis, S. W., & Brewin, C. R. (2013). An fMRI investigation of posttraumatic
flashbacks. Brain and Cognition, 81(1), 151–159. doi:10.1016/j.bandc.2012.10.002
Wheeler, R. E., Davidson, R. J., & Tomarken, A. J. (1993). Frontal brain asymmetry and emotional reactivity: A biological substrate of
affective style. Psychophysiology, 30(1), 82–89.
White, S. A. (2001). Learning to communicate. Current Opinion in Neurobiology, 11(4), 510–520.
Whitfield-Gabrieli, S., Moran, J. M., Nieto-Castañón, A., Triantafyllou, C., Saxe, R., & Gabrieli, J. D. (2011). Associations and
dissociations between default and self-reference networks in the human brain. NeuroImage, 55(1), 225-232.
doi:10.1016/j.neuroimage.2010.11.048
Whitney, C., Kirk, M., O’Sullivan, J., Ralph, M., & Jefferies, E. (2012). Executive semantic processing is underpinned by a large-scale
neural network: Revealing the contribution of left prefrontal, posterior temporal, and parietal cortex to controlled retrieval and selection using
TMS. Journal of Cognitive Neuroscience, 24(1), 133–147. doi:10.1162/jocn_a_00123
Whittle, S., Chanen, A. M., Fornito, A., McGorry, P. D., Pantelis, C., & Yucel, M. (2009). Anterior cingulate volume in adolescents with
first-presentation borderline personality disorder. Psychiatry Research: Neuroimaging, 172, 155–160.
Wicker, B., Keysers, C., Plailly, J., Royet, J.-P., Gallese, V., & Rizzolatti, G. (2003). Both of us disgusted in my insula: The common neural
basis of seeing and feeling disgust. Neuron, 40, 655–664.
Widom, C. S., DuMont, K., & Czaja, S. J. (2007). A prospective investigation of major depressive disorder and comorbidity in abused and
neglected children grown up. Archives of General Psychiatry, 64(1), 49–56.
Wiech, K., Lin, C. S., Brodersen, K. H., Bingel, U., Ploner, M., & Tracey, I. (2010). Anterior insula integrates information about salience
into perceptual decisions about pain. The Journal of Neuroscience, 30(48), 16324–16331.
Wiens, S. (2006). Subliminal emotion perception in brain imaging: Findings, issues, and recommendations. Progress in Brain Research, 156,
105–121.
Willcutt, E. G., Dole, A. E., Nigg, J. T., Faraone, S. V., & Pennington, B. F. (2005). Validity of the executive function theory of attention
deficit/hyperactivity disorder: A meta-analytic review. Biological Psychiatry, 57(11), 1336–1346.
Williams, L. M. (1994). Recall of childhood trauma: A prospective study of women’s memories of child sexual abuse. Journal of Consulting
and Clinical Psychology, 62(6), 1167–1176.
Williams, L. M., Kemp, A. H., Felmingham, K., Barton, M., Olivieri, G., Peduto, A., . . . Bryant, R. A. (2006). Trauma modulates
amygdala and medial prefrontal responses to consciously attended fear. NeuroImage, 29(2), 347–357.
Williams, L. M., Phillips, M. L., Brammer, M. J., Skerrett, D., Lagopoulos, J., Rennie, C., . . . Gordon, E. (2001). Arousal dissociates
amygdala and hippocampal fear responses: Evidence from simultaneous fMRI and skin conductance recording. NeuroImage, 14(5), 1070–1079.
Wilson, E. O. (2012). The social conquest of Earth. New York: Norton.
Wilson, F. A. W., O’Scalaidhe, S. P., & Goldman-Rakic, P. S. (1993). Dissociation of object and spatial processing domains in primate

293
prefrontal cortex. Science, 260(5116), 1955–1958.
Wilson, F. R. (1998). The hand. New York: Vintage.
Wilson, R. S., Boyle, P. A., Levine, S. R., Yu, L., Anagnos, S. E., Buchman, A. S., . . . Bennett, D. A. (2012). Emotional neglect in
childhood and cerebral infarction in older age. Neurology, 79(15), 1534-1539. doi:10.1212/wnl.0b013e31826e25bd
Wilson, S. L. (2003). Post-institutionalization: The effects of early deprivation on development of Romanian adoptees. Child and Adolescent
Social Work Journal, 20(6), 473–483. doi:10.1023/b:casw.0000003139.14144.06
Wingenfeld, K., Schaffrath, C., Rullkoetter, N., Mensebach, C., Schlosser, N., Beblo, T., . . . Meyer, B. (2011). Associations of childhood
trauma, trauma in adulthood and previous-year stress with psychopathology in patients with major depression and borderline personality
disorder. Child Abuse and Neglect, 35, 647–654.
Winick, M., Katchadurian, K., & Harris, R. C. (1975). Malnutrition and environmental enrichment by early adoption. Science, 190(4220),
1173–1175.
Winnicott, D. W. (1965a). The capacity to be alone. In Maturational processes and the facilitating environment: Studies in the theory of
emotional development (pp. 29–36). New York: International Universities Press.
Winnicott, D. W. (1965b). Ego integration in child development. In Maturational processes and the facilitating environment: Studies in the
theory of emotional development (pp. 56–63) New York: International Universities Press.
Winnicott, D. W. (1965c). From dependence to independence in the development of the individual. In Maturational processes and the
facilitating environment: Studies in the theory of emotional development (pp. 83–99). New York: International Universities Press.
Winning, A., Glymour, M. M., McCormick, M. C., Gilsanz, P., & Kubzansky, L. D. (2015). Psychological distress across the life course
and cardiometabolic risk. Journal of the American College of Cardiology, 66(14), 1577–1586. doi:10.1016/j.jacc.2015.08.021
Witelson, S. F., Kigar, D. L., & Harvey, T. (1999). The exceptional brain of Albert Einstein. Lancet, 353(9170), 2149–2153.
Wittling, W. (1997). The right hemisphere and the human stress response. Acta Physiologica Scandinavica, 640(Suppl.), 55–59.
Wittling, W., & Pfluger, M. (1990). Neuroendocrine hemisphere asymmetries: Salivary cortisol secretion during lateralized viewing of
emotion-related and neutral films. Brain and Cognition, 14(2), 243–265.
Wolf, N. S., Gales, M. E., Shane, E., & Shane, M. (2000). The developmental trajectory from amodal perception to empathy and
communication: The role of mirror neurons in this process. Psychoanalytic Inquiry, 21(1), 94–112.
Wolf, R. C., Sambataro, F., Vasic, N., Schmid, M., Thomann, P. A., Bienentreu, S. D., & Wolf, N. D. (2011). Aberrant connectivity of
resting-state networks in borderline personality disorder. Journal of Psychiatry and Neuroscience, 36(6), 402–411.
Wolpe, J. (1958). Psychotherapy by reciprocal inhibition. Stanford, CA: Stanford University Press.
Wolpert, D. M., Goodbody, S. J., & Husain, M. (1998). Maintaining internal representations: The role of the human superior parietal lobe.
Nature Neuroscience, 1(6), 529–533.
Woolley, C. S., Gould, E., & McEwen, B. S. (1990). Exposure to excess glucocorticoids alters dendritic morphology of adult hippocampal
pyramidal neurons. Brain Research, 531(1–2), 225–231.
Wu, J., Buchsbaum, M. S., Gillin, J. C., Tang, C., Cadwell, S., & Wiegand, M. . . . Bunney, W. E., Jr. (1999). Prediction of antidepressant
effects of sleep deprivation by metabolic rates in the ventral anterior cingulate and medial prefrontal cortex. American Journal of Psychiatry,
156(8), 1149–1158.
Wykes, T., Brammer, M., Mellers, J., Bray, P., Reeder, C., Williams, C., & Corner, J. (2002). Effects of the brain of psychological
treatment: Cognitive remediation therapy. British Journal of Psychiatry, 191, 144–152.
Xue, G., Lu, Z., Levin, I. P., & Bechara, A. (2010). The impact of prior risk experiences on subsequent risky decision-making: The role of
the insula. NeuroImage, 50, 709–716.
Yamanishi, T., Nakaaki, S., Omori, I. M., Hashimoto, N., Shinagawa, Y., Hongo, J., . . . FurukaÏwa, T. A. (2009). Changes after behavior
therapy among responsive and nonresponsive patients with obsessive-compulsive disorder. Psychiatry Research: Neuroimaging, 172(3), 242–250.
doi:10.1016/j.pscychresns.2008.07.004
Yang, J., Weng, X., Zang, Y., Xu, M., & Xu, X. (2010). Sustained activity within the default mode network during an implicit memory
task. Cortex, 46(3), 354–366.
Yang, Y., Kircher, T., & Straube, B. (2014). The neural correlates of cognitive behavioral therapy: Recent progress in the investigation of
patients with panic disorder. Behaviour Research and Therapy, 62, 88-96. doi:10.1016/j.brat.2014.07.011
Yau, J. L. W., Olsson, T., Morris, R. G. M., Meaney, M. J., & Seckl, J. R. (1995). Glucocorticoids, hippocampal corticosteroid receptor
gene expression and antidepressant treatment: Relationship with spatial learning in young and aged rats. Neuroscience, 6(3), 571–581.
Yehuda, R. (1999). Biological factors associated with susceptibility to post-traumatic stress disorder. Canadian Journal of Psychiatry, 44(1),
34–39.
Yehuda, R., Bierer, L. M., Schmeidler, J., Aferiat, D. H., Breslau, I., & Dolan, S. (2000). Low cortisol and risk for PTSD in adult offspring
of holocaust survivors. American Journal of Psychiatry, 157(8), 1252–1259. doi:10.1176/appi.ajp.157.8.1252

294
Yehuda, R., Kahana, B., Schmeidler, J., Southwick, S. M., Wilson, S., & Giller, E. I. (1995). Impact of cumulative lifetime trauma and
recent stress on current posttraumatic stress disorder symptoms in holocaust survivors. American Journal of Psychiatry, 152(12), 1815–1818.
Yehuda, R., & Siever, L. J. (1997). Persistent effects of stress in trauma survivors and their descendants. Biological Psychiatry, 41, 1S–120S.
Yerkes, M., & Dodson, D. (1908). The relation of strength to rapidity of habit–formation. Journal of Comparative Neurology and Psychology,
18, 459–482.
Yoshimura, S., Ueda, K., Suzuki, S., Onoda, K., Okamoto, Y., & Yamawaki, S. (2009). Self-referential processing of negative stimuli
within the ventral anterior cingulate gyrus and right amygdala. Brain and Cognition, 69(1), 218-225. doi:10.1016/j.bandc.2008.07.010
Yovell, Y. (2000). From hysteria to posttraumatic stress disorder: Psychoanalysis and the neurobiology of traumatic memories.
Neuropsychoanalysis, 2(2), 171–181.
Yuen, G., Gunning-Dixon, F., Hoptman, M., Abdelmalak, B., Mcgovern, A., Seirup, J., & Alexopoulos, G. (2014). The salience network
in the apathy of late-life depression. International Journal of Geriatric Psychiatry, 29(11), 1116–1124. doi:10.1002/gps.4171
Yu-Feng, Z., Yong, H., Chao-Zhe, Z., Qing-Jiu, C., Man-Qiu, S., Meng, L., . . . Yu-Feng, W. (2007). Altered baseline brain activity in
children with ADHD revealed by resting-state functional MRI. Brain and Development, 29(2), 83–91.
Zahm, D. S. (2006). The evolving theory of basal forebrain functional-anatomical “macrosystems.” Neuroscience and Behavioral Reviews,
30(2), 148–172.
Zak, P. J., Stanton, A. A., & Ahmadi, S. (2007). Oxytocin increases generosity in humans. PLoS One, 2(11), 1–5.
doi:10.1371/journal.pone.0001128
Zaki, J., Ochsner, K. N., Hanelin, J., Wager, T. D., & Mackey, S. C. (2007). Different circuits for different pain: Patterns of functional
connectivity reveal distinct networks for processing pain in self and others. Society for Neuroscience, 2(3–4), 276–291.
Zald, D. H., & Kim, S. W. (2001). The orbitofrontal cortex. In S. P. Salloway, P. F. Malloy, & J. D. Duffy (Eds.), The frontal lobes and
neuropsychiatric illness (pp. 33–69). Washington, DC: American Psychiatric Press.
Zang, Y., Jin, Z., Weng, X., Zhang, L., Zeng, Y., & Yang, L. (2005). Functional MRI in attention-deficit hyperactivity disorder: Evidence
for hypofrontality. Brain and Development, 27(8), 544–550.
Zeanah, C. H., & Fox, N. A. (2004). Temperament and attachment disorders. Journal of Clinical Child and Adolescent Psychology, 33(1), 32–
41. doi:10.1207/S15374424JCCP3301_4
Zeitlin, S. B., Lane, R. D., O’Leary, D. S., & Schrift, M. J. (1989). Interhemispheric transfer deficit and alexithymia. American Journal of
Psychiatry, 146(11), 1434–1439.
Zeitlin, S. B., & McNally, R. J. (1991). Implicit and explicit memory bias for threat in post traumatic stress disorder. Behavior Research and
Therapy, 29(5), 451–457.
Zeltzer, L. K., Anderson, C. T. M., & Schecter, N. L. (1990). Pediatric pain: Current status and new directions. Current Problems in
Pediatrics, 20(8), 415–486.
Zetzsche, T., Preuss, U. W., Frodl, T., Schmitt, G., Seifert, D., Munchhausen, E., . . . Meisenzahl, E. M. (2007). Hippocampal volume
reduction and history of aggressive behaviour in patients with borderline personality disorder. Psychiatry Research: Neuroimaging, 154, 157–170.
Zhang, L. X., Levine, S., Dent, G., Zhan, Y., Xing, G., Okimoto, D., . . . Smith, M. A. (2002). Maternal deprivation increases cell death in
the infant rat brain. Developmental Brain Research, 133(1), 1–11.
Zhang, T. Y., Chretien, P., Meaney, M. J., & Gratton, A. (2005). Influence of naturally occurring variations in maternal care on prepulse
inhibition of acoustic startle and the medial prefrontal cortical dopamine response to stress in adult rats. Journal of Neuroscience, 25(6), 1493–
1502.
Zhang, T., Parent, C., Weaver, I., & Meaney, M. J. (2004). Maternal programming of individual differences in defensive responses in the
rat. Annals of New York Academy of Sciences, 1032, 85–103.
Zhang, Y., Liu, F., Chen, H., Li, M., Duan, X., Xie, B., & Chen, H. (2015). Intranetwork and internetwork functional connectivity
alterations in post-traumatic stress disorder. Journal of Affective Disorders, 187, 114–121. doi:10.1016/j.jad.2015.08.043
Zhao, M., Toyoda, H., Lee, Y., Wu, L., Ko, S. W., Zhang, X., . . . Zhuo, M. (2005). Roles of NMDA NR2B subtype receptor in
prefrontal long-term potentiation and contextual fear memory. Neuron, 47(6), 859–872.
Zhou, Y., Wang, Z., Qin, L., Wan, J., Sun, Y., Su, S., . . . Xu, J. (2012). Early altered resting-state functional connectivity predicts the
severity of post-traumatic stress disorder symptoms in acutely traumatized subjects. PLoS One, 7(10). doi:10.1371/journal.pone.0046833
Zhu, X. O., & Waite, P. M. E. (1998). Cholinergic depletion reduces plasticity of barrel field cortex. Cerebral Cortex, 8(1), 63–72.
Zhu, X., Wang, X., Xiao, J., Liao, J., Zhong, M., Wang, W., & Yao, S. (2012). Evidence of a Dissociation Pattern in Resting-State Default
Mode Network Connectivity in First-Episode, Treatment-Naive Major Depression Patients. Biological Psychiatry, 71(7), 611-617. doi:10.101
s6/j.biopsych.2011.10.035
Zola-Morgan, S. M., & Squire, L. R. (1990). The primate hippocampal formation: Evidence for a time-limited role in memory storage.
Science, 250(4978), 288–290.

295
Zoroglu, S. S., Tuzun, U., Sar, V., Tutkun, H., Savaçs, H., Ozturk, M., . . . Kora, M. E. (2003). Suicide attempt and self-mutilation among
Turkish high school students in relation with abuse, neglect, and dissociation. Psychiatry and Clinical Neurosciences, 57, 119–126.
Zubieta, J. K., Bueller, J. A., Jackson, L. R., Scott, D. J., Xu, Y., Koeppe, R. A., . . . Stohler, C. S. (2005). Placebo effects mediated by
endogenous opioid activity on μ- opioid receptors. Journal of Neuroscience, 25(34), 7754–7762.
Zuckerman, B., Bauchner, H., Parker, S., & Cabral, H. (1990). Maternal depressive symptoms during pregnancy, and newborn irritability.
Developmental and Behavioral Pediatrics, 11(4), 190–194.
Zysset, S., Huber, O., Ferstl, E., & Cramon, D. V. (2002). The anterior frontomedian cortex and evaluative judgment: An fMRI study.
NeuroImage, 15(4), 983-991. doi:10.1006/nimg.2001.1008

296
Acknowledgments

I would like to thank Deborah Malmud and her team at W. W. Norton for their assistance with this and the
many projects we have worked on together over the last two decades. Special thanks go to Erin Santos and
Vanessa Davis for their unfailing support, encouragement, and good humor through the writing and
reorganization of this third edition. Finally, to all of you who have been on the journey with me to create a
scientific psychotherapy—we have very far to go, but we have come very far.

297
Index

Page numbers listed correspond to the print edition of this book. You can use your device’s search function to locate particular terms in the text.

In this index, f denotes figure and t denotes table.

9/11 attacks, memories of, 375–76

AA. See Alcoholics Anonymous (AA)


AAI. See Adult Attachment Interview (AAI)
abuse. See childhood abuse; early traumatic stress
adaptation, 247, 284, 420–21
See also evolution; exaptation
addiction to stress and self-harm, 339–340
the adolescent brain, 130, 134
adolescents, 390
Adult Attachment Interview (AAI), 225–26, 227
affect and cognition dissociation, 26
affective memory. See amygdala; amygdaloid memory networks
affect labeling, 191–92
affect regulation. See emotional tolerance and affect regulation
affordance, 151–54
aging, 86, 291, 385
agoraphobia, 291
Ainsworth, M., 221, 222
Alcoholics Anonymous (AA), 274, 428
alcoholism case study, 372–77
alexithymia, 30, 110–11
alone, capacity to be, 218–19
altruism, 263–271, 270t, 273–75
Alzheimer’s disease, 385–86
amygdala
about, 62, 81f, 254–55, 366
ACC and, 179
altruism and, 267
anxiety and fear and role of, 288–292, 290f, 297
conservatives and, 272
effects of abnormal functioning of, 84–85
fast fear network and, 286, 288
functions of, 83–84, 87
institutionalized children and, 312
ompfc and, 132, 178, 298
oxytocin and, 265
taxon system, 300
See also cortical-amygdala circuitry; fear processing; hippocampus-amygdala activation
“amygdala hijack,” 178
amygdaloid-hippocampal interaction, 87–89
amygdaloid memory networks, 83–84, 88, 133, 333

298
anaclitic depression, 203
Anderson, N., 147
anosognosia, 254
anterior cingulate cortex (ACC)
about, 11, 134, 135t, 253
amygdala and, 179
conservatives and liberals, 272
functions of, 131–32
treatment outcomes to, 407
anterior insula (AI), 134, 135t
See also insula cortex
anxiety and anxiety disorders
adaptiveness and, 284
ADHD misdiagnosis and, 139
amygdala’s role in, 288–292, 290f, 297
avoidance and, 284–85
BPD and, 321
cognitive therapies and, 41
depression and, 109–10
differentiation and, 43
dlpfc and, 133
DMN and, 163, 185
evolution and, 283
frontal-parietal executive network and, 149
imagination and, 284
narratives and, 189
ompfc-amygdala activation and, 179
psychotherapy goals for, 23–24
TPN and, 185
treatment outcomes for, 406
unconscious self-deception and, 371
Apollo 13 crisis, 9
appraisals, 100, 132, 286, 288
arousal, moderate states of, 386–88, 387f, 389t, 398–99
association areas, 20
attachment, 204–5, 206–9, 237–261
See also stress, fear, and attachment system
attachment disorders, 311–12, 339, 390
attachment plasticity, 389–392, 391t
attachment schemas
about, 89–90
affordance and, 153
appraisals and, 288
fluctuations in security of, 390–92
imprinting and, 12
navigation and, 151
relationship status’ effects on, 251
See also implicit memory; pathological caretaking; secure attachment
attention. See salience network and focus of attention
attention-deficit/hyperactivity disorder (ADHD), 135–39, 138t
attunement and reciprocity, 48, 125, 205–6, 422–24
autism, 211

299
autobiographical memory, 49, 81, 159, 354–55
See also narratives
autonomic nervous system, 60
avoidance, 284–85, 325
awareness. See other-awareness; self-awareness

basal forebrain, 253


BDNF. See brain-derived neurotrophic factor (BDNF)
Bechara, A., 289
bed nucleus of the stria terminalis (BNST), 291–92
behavior. See cognition and motor behavior; goal-directed behavior
behavioral therapy (BT), 407
See also dialectical behavioral therapy
belief perseverance, 169
Berger, H., 156
biases, 100, 169, 422
Big Book (AA), 274
bilateral stimulation. See EMDR
birds, 40, 60, 398
Bisiach, E., 104
BNST. See bed nucleus of the stria terminalis (BNST)
body experiences, 145
Bogen, J., 111
Bollas, C., 369
bonding, 204, 355
See also attachment; mother-infant dyad
borderline personality disorder (BPD)
about, 313–16
amygdaloid memory networks and, 88
DMN and, 320–21
early traumatic stress and, 307
interpretations and, 351
treatment rationales and combinations, 414–15
Bowen, M., 43, 44, 48
Bowlby, J., 221, 222
BPD. See borderline personality disorder (BPD)
Bradshaw, J., 217
the brain
attachment and, 248–250, 250t
conservativeness of, 418–19
exaptation within, 8–10
as social organ of adaptation, 420–21
using the mind to change, 428
See also the adolescent brain; the mind; neural network integration; neural networks; social brain
brain-derived neurotrophic factor (BDNF), 242, 244, 248, 382
brain development, 69–71, 97–98, 238–240
See also evolution
brain functioning measurement tools, 68
brain metabolism, 401, 405
brain stem, 61
British Medical Journal, 395
Broca, P., 5, 96, 119
Broca’s area

300
fine motor control and, 123
functions of, 82, 99, 363
mirror neurons and, 212
traumatic flashbacks and speechless terror, 335–36
BT. See behavioral therapy (BT)
Buddhism, 164–67, 275, 433
Buddhist psychology, 166–67

Cahill, L., 333


caloric tests, 410
Capgras syndrome, 174–75, 177
caretaking, 202, 268, 306
See also child care and longevity; mother-infant dyad; mothers; parenting; pathological caretaking
case conceptualization and treatment, 166–67
Cassidy, J., 232
caudate nucleus, 42
CBT. See cognitive-behavioral therapy (CBT)
cerebellum, 9
cerebral cortex
disadvantages of, 12
explicit memory and, 82
factors affecting the development of, 10–11
inhibition and, 123–25
roles of, 24, 61–62, 62f
transcription genetics and, 20
See also specific cortices and lobes
change, 420–22, 433
See also adaptation; evolution; exaptation
character armor, 45–46
Charcot, J.-M., 3, 4, 7
child care and longevity, 249–250, 250t
childhood abuse, 296
See also early traumatic stress
childhood amnesia, 80
childhood sexual abuse, 329–330
See also early traumatic stress
childhood trauma. See early traumatic stress
children, gifted, 355
children, institutionalized, 312
children, Korean, 385
children, orphaned, 202, 311, 312
children of Holocaust survivors, 329
cingulate cortex, 62, 134, 253
See also anterior cingulate cortex (ACC)
client-centered therapy. See Rogerian or client-centered therapy
cognition and affect dissociation. See affect and cognition dissociation
cognition and motor behavior, 122–23
cognitive-behavioral therapy (CBT)
about, 41, 42, 48, 170
predictors of success in, 180, 414
See also treatment outcomes
cognitive rehabilitation, 414
cognitive reserve, 385–86

301
cognitive therapies, 40–42, 189–190, 301, 406
See also cognitive-behavioral therapy (CBT)
complex PTSD (C-PTSD), 309–11
compulsive disorders, 330
See also obsessive-compulsive disorder (OCD)
computerized tomography (CT), 67
confabulation, 106
confirmation bias, 169
conscious control. See cortical inhibition and conscious control
conscious memory, 75, 80
consciousness, illusions of, 168–170
conservatives versus liberals, 272–73
constructive introspection, 428–431
control, sense of, 192
corpus callosum, 61, 107–8
See also split-brain research
cortex. See cerebral cortex
cortical-amygdala circuitry, 298–99
cortical inhibition and conscious control, 71–72
cortisol
damaging effects of, 367–68
effects of, 64–65, 295
learning and, 388
maternal attention and, 242
maternal deprivation and, 244
memory deficits and, 78
See also glucocorticoids (GCs)
CT. See computerized tomography (CT)

Davis, M., 285


Dawson, M., 136
daydreaming, 159
D-cycloserine, 383
default-mode network (DMN), 156–164, 178, 184–85, 269–271, 320–21
defenses
about, 25
ADHD misdiagnosis and, 139
capacity to be alone and, 218–19
confabulation and, 106
functions of, 36–37, 167–68
impingements and, 215
interpretations and, 351
in Reichian therapy, 46
See also unconscious self-deception
déjà vu experiences, 175
depression
amygdaloid-hippocampal interaction and, 88–89
anaclitic, 203
anxiety and, 109–10
cognitive therapies and, 41
DMN and, 162–63
hemispheric laterality and, 101, 430
hippocampus volume and, 244

302
left-right hemisphere balance and, 180–81
left-right integration and, 30
memory deficits and, 78
ompfc-amygdala activation and, 179
success of psychotherapy for, 42
TPN and DMN and, 185
treatment options for, 409
treatment outcomes for, 406, 407
See also maternal depression
developmental trauma disorder. See complex PTSD (C-PTSD)
diagnosis and treatment, 400–402, 405–6
dialectical behavioral therapy, 414–15
DID. See dissociative identity disorder (DID)
differentiation, 43
directed action and inhibition, 119–139, 127t, 128t, 135t, 138t
dissociation, 22, 310–11, 330, 340–41
See also affect and cognition dissociation
dissociative identity disorder (DID), 358
divergent hemispheric processing, 363–64
dlpfc. See dorsolateral prefrontal cortex (dlpfc)
dlpfc–ompfc activation, 182–83
DMN. See default-mode network (DMN)
DNA, methylation of, 241
Dodson, D., 333
Dodson, J., 387
dogs, 208
dolphins, 265
dopamine, 63, 327, 394
See also oxytocin and dopamine
dorsolateral prefrontal cortex (dlpfc), 126, 133
double amnesia, 354
Douglas, R., 87
dukkha, 165–66

early learning, bias towards, 365–66


early traumatic stress
about, 305–9, 308t, 328–330
BPD and, 313–16
case study, 330–32
C-PTSD, 309–11
divergent hemispheric processing and, 364
DMN and, 160–61, 164, 320–21
hippocampus and, 296
self-fragmentation, 316–19
social fragmentation, 319–322
See also alcoholism case study
educational level, 385–86
egocentric bias, 169
ego strength, 36
Eightfold Path, 166–67
Einstein, A., 145–46
elephants, 265
EMDR, 410, 411–12

303
emotional tolerance and affect regulation
about, 24–25
alexithymia, 110–11
amygdala-anterior cingulate activation and, 179
dlpfc–ompfc activation and, 182
effect of early experiences on, 204–5
left-right integration and, 30
narratives and, 189–190, 191–92
ompfc-amygdala activation and, 178–79
prefrontal cortex and, 126–27
as psychotherapy goal, 28–29, 49
shame and, 217
stress and, 23
See also sensory, motor, and affective systems
emotion and hemispheric laterality, 101–3, 102t, 109–10
empathy and empathic attunement, 315
See also attunement and reciprocity
endogenous opioids, 208, 284, 327, 330, 339–340
endorphins, 64, 208
enriched environments, 21–22, 384–86, 384t
environmental programming, 240–42, 243t, 244–48, 245t, 246t
environments. See enriched environments
epigenetics. See genetics and epigenetics
epilepsy, 85
evolution
anxiety and, 283
brain complexity and, 238–240
disruption in neural network integration and, 27
hemispheric laterality and, 97–98, 429–430
of language, 212
maternal dependence and, 11
memory distribution and, 82
narratives and, 186
and necessity of psychotherapy, 361–378
social brain and, 201–2
speechless terror and, 338
triune brain and, 7–8
See also exaptation
exaptation, 8–10
executive functioning
about, 120–21
affordance, 151–54
attention-deficit/hyperactivity disorder (ADHD), 135–39, 138t
Buddhism and, 164–67
directed action and inhibition, 119–139, 127t, 128t, 135t, 138t
DMN and, 156–164, 158t, 160t
dorsolateral prefrontal cortex (dlpfc), 133
emotional tolerance and affect regulation, 178–79
frontal lobes, 122–23
frontal-parietal executive network, 146–49
illusions of consciousness, 167–170
mirror neurons and, 149–150, 150t

304
navigation and attachment schema, 151
orbitomedial prefrontal cortex (ompfc), 131–32
parietal lobes and, 140–155
prefrontal cortex, 126–131, 127t, 128t
salience network and focus of attention, 134, 135t
sense of self, 154–55
experience-dependent plasticity, 251
See also environmental programming
explicit memory, 79, 80, 80t, 81, 82
exposure and response prevention, 41
extinction learning, 297, 300, 301
eyes and eye contact, 208, 209–10
See also side-to-side eye movements

facial recognition and facial expressions, 209–11, 256


false memories, 91–92
false self, 343, 352
See also true versus false self
family therapy. See systemic family therapy
fast fear network, 286, 287f, 288
fear, overcoming, 297–300
fear, tenacity of, 366–67
fear networks, fast and slow, 285–88, 287f
fear processing, 283–293, 287f, 290f
See also amygdala; amygdaloid memory networks; stress, fear, and attachment system
fears and phobias, 300–301
feedforward and feedback neural network, 17f, 18
feedforward neural network, 16f
Field, T., 329
fight-or-flight response, 132, 288–89, 292, 294
Fisher, H., 260
flashbacks, 88
See also traumatic flashbacks and speechless terror
flexibility. See habits and flexibility
Fonagy, P., 232
formless quiescence, 218
Four Noble Truths, 166
fragmentation. See self-fragmentation; social fragmentation
Freud, S., 3–6, 8, 35–37, 76, 94
frontal cortex and frontal lobes, 62, 122–23, 146
See also Broca’s area
frontal lobe disorders, 6
frontal-parietal executive network, 146–49
functional brain imaging, 400–401
fundamental attribution error, 169

Gage, P., 6, 125


Gazzaniga, M., 105
genetic inheritance and gene expression, 20–21
genetics and epigenetics, 65–67, 241–42, 246, 247, 249
Gestalt therapy, 46–47
gestures, 98–99, 105
gifted children, 355

305
glia, 58–59
glove anesthesia, 4
glucocorticoids (GCs), 64–65, 86, 294, 295, 327
See also cortisol
goal-directed behavior, 143, 147, 149
Goeldi monkeys, 249
good-enough mothers, 213
good-enough parenting, 154–55, 218
grandmother gene hypothesis, 249
grasping reflex, 24, 123–25
Grice’s maxims, 226
guided altruism, 273–75
guilt, 218, 356
habits and flexibility, 421–22
Haier, R., 147
HALT technique, 428
handedness, 98–99, 100
handling. See human handling; infant massage; touch, soothing
Hanna, C., 296
Harlow, H., 234, 235, 311
hemineglect, 103–4
hemispheric laterality, 12, 30, 96–115, 409–10, 429–430
See also left-right hemisphere balance; left-right integration
hemispheric processing, divergent, 363–64
heroic narratives, 188
See also The Little Engine That Could
heuristics, 34
high arousal, 42
See also hyperarousal
hippocampal memory networks, 86–87, 133
hippocampal volume, 244, 246
hippocampus
about, 62, 81f, 86, 255
amygdaloid memory networks and, 84
cortisol and, 388
early traumatic stress and, 296
effects of abnormal functioning of, 86–87
explicit memory and, 82
fear processing and, 286, 288, 289
functions of, 86, 87, 141
locale system, 300
in PTSD, 327
stress and, 293–97, 295t
See also amygdaloid-hippocampal interaction
hippocampus-amygdala activation, 183–84
histories of neurology and psychology, 3–13
history taking, 296, 310
Holocaust survivors, children of, 329
Hoppe, K., 111
Howard, K., 32
HPA axis, 67, 183, 257
Huang, Z., 382
human handling, 242, 245, 246t

306
hyperarousal, 324
hypergraphia, 85
hypothalamic-pituitary-adrenal (HPA) activation, 183
hypothalamus, 255–56
hysteria, 4, 8

id, 5
illness and stress, 294
illusions of consciousness, 167–170
imagination, 145–46, 159, 284
imaging. See scanning techniques
impingements, 215, 218
implicit memory, 79, 80, 80t, 89–91
imposters, delusion of. See Capgras syndrome
imprinting, 11–12
infant massage, 408
infants, 291, 329
See also mother-infant dyad
inhibition (brain function), 24, 123–25
See also directed action and inhibition; executive functioning
inner objects, 43
insecure attachment, 390
See also pathological caretaking
instantiations, 18
instinctual drives, 5
institutional autism, 311
insula cortex, 62, 134, 254, 272
See also anterior insula (AI)
intelligence, 146–48
internalization of others, 143–44
internalized mother, 237–38
interpretations, 37, 350–52
intimacy, 213
See also attunement and reciprocity
introspection. See constructive introspection; self-reflection
intrusion, 324–25
See also traumatic flashbacks and speechless terror
intuition, 132
inverted-U learning curve, 387, 387f, 388, 389t
irritation of life, 165–66

Jackson, J., 96
Jacobs, W., 300, 301
Jason, G., 107
Jaynes, J., 112
Johnson, M., 145
Jonas, W., 392
journaling, 192
Journal of Family Practice, 395
judgment. See fundamental attribution error
Jung, R., 147

Kandel, E., 60

307
Kang, H., 382
Kaplan, N., 232
Kling, A., 202
Korean children, 385
Koverola, C., 296

language, 98–99, 189, 192–95, 212, 399–400


See also Broca’s area; Wernicke’s area
language network, 105–7
language suppression, 362–63
See also traumatic flashbacks and speechless terror
lateral asymmetry, 98–100
laterality. See hemispheric laterality
learning
bias towards early, 365–66
enriched environments and, 21–22
extinction, 297, 298, 300, 301
mirror neurons and, 212
moderate states of arousal and, 387–88
neural network modification and, 18
stress and, 22–24
trial-and-error, 71
LeDoux, J., 105, 285f
left hemisphere
CBT and bias toward, 180
children and development of, 98
functions of, 30, 99, 100, 101
language network and, 105–7
See also language
left-hemisphere interpreter, 105–7, 114
left-right hemisphere balance, 180–81, 214
left-right integration, 30, 31, 87
left-side neglect. See hemineglect
Les Miserables (Hugo), 276–78
liberals. See conservatives versus liberals
limbic integration cortex, 254
See also insula cortex
limbic system, 62
The Little Engine That Could, 191
locale system, 300, 301
locus coeruleus (LC), 292–93
longevity. See child care and longevity
long-term potentiation (LTP), 69, 289, 382, 383
Lorenz, K., 11–12
LTP. See long-term potentiation (LTP)
Luzzatti, C., 104

Mace, R., 249


MacLean, P., 7, 10
magic tricycle, 92–94
magnetic resonance imaging (MRI), 67
Magritte, R., 356–58
Main, M., 221, 232

308
mania, 30, 101, 181
manic defenses, 218–19
maternal attention, 242, 243t, 247
maternal behavior in rats and offspring, 240–42, 243t, 244–45, 245t, 246t
maternal depression, 329, 390, 408–9
maternal deprivation, 244, 245t, 246, 247, 409
maternal preoccupation, 206, 214
maternal stress, 328
mathematical concepts, 145
McClarty, B., 296
McGaugh, J., 333
McGregor, I., 249
Meaney, M., 67, 240–41
meaning response, 392, 393
medications, 136, 137, 405, 406–7, 413
meditation, 155, 167
memory, 91–94, 187
See also specific types of memory
memory deficits, 76–78, 86, 87
See also childhood amnesia; double amnesia
memory distribution, 82
memory encoding and memory integration, 333–34
memory systems. See amygdaloid memory networks; hippocampal memory networks; multiple memory systems
Mendel, A., 65–66
menopause, 249
mental imagery, 145–46
metabolism. See brain metabolism
metacognition, 127
See also the mind
metaphors, 145
methylation of DNA, 241
mice, 382, 387
mild to moderate stress (MMS), 22–23
Miller, A., 352, 353–56
the mind, 6, 419, 426, 428
mirroring process, 214, 356–57, 357f
See also attunement and reciprocity
mirror neurons
about, 142, 211–13, 257
altruism and, 269
executive functioning and, 149–150, 150t
misattunement, 215, 217
MMS. See mild to moderate stress (MMS)
Moerman, D., 392
monkeys
effects of isolation in, 234–35, 311–12
effects of stress on, 293
frontal-parietal executive network, 148
longevity in, 249
maternal deprivation in, 246
mirror neurons in, 149, 211–12
Moore, G., 152
Moran, G., 232

309
mother-infant dyad, 205, 213–15
See also attachment
mothers, 248–49
See also good-enough mothers; internalized mother; maternal entries; mother-infant dyad
motivation. See social motivation system
motor behavior and cognition, 122–23, 383
See also sensory, motor, and affective systems
MRI. See magnetic resonance imaging (MRI)
multiple memory systems, 78–83, 80t, 81f
See also amygdaloid-hippocampal interaction; amygdaloid memory networks; hippocampal memory networks
multiple sclerosis, 58

Nachtraglichkeit, 94
Nadel, L., 300, 301
Naikan therapy, 275
narcissism, 343, 352–53
narratives
coconstruction of, 399
emotional regulation and, 189–190, 191–92
evolution and, 186
functions of, 180, 186, 324
heroic, 188, 191
influence of parents and therapists through, 195
neural networks and, 187
speechless terror and, 336
See also autobiographical memory; magic tricycle; stories
navigation, 141, 151
negative self-image, 89
neglect, severe, 306
Nelson, E., 260
neomammalian brain, 7, 8, 79
See also cerebral cortex
neophobia, 366–67, 410–11
nervous system overview, 57–74
neural network integration, 26–31, 34–53, 340–41
neural networks
about, 15–19, 16f, 17f, 401
as curative element, 398
dlpfc–ompfc activation, 181–83
exaptation within, 9–10
growth and integration within, 19–20, 176–77, 177t
narratives and, 187
OCD and, 401–2
PTSD and, 42, 340–41, 403
reregulation of, 402–3, 403t–5t, 405–7
See also specific circuitry or networks
neural plasticity
about, 58–59
brain development and, 69–71
definitions of, 19, 58–59
enriched environments and, 21, 384–86, 384t
factors supporting, 28
interpretations and, 350–52

310
moderate states of arousal and, 386–88, 387f, 389t
stimulation of, 379–396
use-dependent, 381–82
neural systems, 60–62, 62f
neurobiological model of the mind (Freud), 6
See also triune brain
neurogenesis, 59–60
See also brain-derived neurotrophic factor (BDNF)
neuroimaging, 401
neurology history, 3–13
neurons, 15, 57–58, 203
neuropeptides, 64
neuroscience, 14, 416
The Neuroscience of Human Relationships, 2nd. ed. (Cozolino), 252
neuroscientists, psychotherapists as, 397–417
neurotransmitters and neuromodulators, 63–64
newborns, 206–7
See also infants; mother-infant dyad
9/11 attacks, memories of, 375–76
N-methyl-D-aspartate (NMDA), 382–83
norepinephrine (NE), 63–64, 292–93, 326, 333

obsessive-compulsive disorder (OCD), 27, 41–42, 401–2, 405, 407


occipital cortex and lobe, 61, 82
occupational therapy, 130, 131, 145
ompfc. See orbitomedial prefrontal cortex (ompfc)
opioids, 393
See also endogenous opioids
optimal stress, 48
See also mild to moderate stress (MMS)
orbitomedial prefrontal cortex (ompfc)
about, 126, 131–32, 157, 253
altruism and, 267–68
amygdala and, 132, 178, 298
dlpfc and, 133, 182
OCD and, 402, 405
See also Gage, P.
orienting reflex, 293, 411–12
Orlinsky, D., 32
orphaned children, 202
other-awareness, 161
out-of-body experiences, 330, 342
owl monkeys, 249
oxytocin and dopamine, 265–67

pain versus suffering, 433


Pajurkova, E., 107
paleomammalian brain, 7, 8, 10, 79
See also limbic system
panic attacks, 290–91, 302, 407
Panksepp, J., 260
parenting, 31–33, 154–55, 240, 259
See also caretaking; mother-infant dyad; mothers; pathological caretaking

311
parietal cortex and lobes, 62, 103, 140–46, 142t, 144t
See also temporal-parietal junction (TPJ)
Parkinson’s disease, 394
parrots, 246t
pathological caretaking, 343–350, 352–56, 357
Paxil, 64, 89
Pennebaker, J., 192
Perls, F., 46
pharmacotherapy, 406
See also medications
phobias, 41
See also fears and phobias
Piazza del Duomo, Milan, 104
pigs, 246t
placebo effects, 392–93, 394–95, 394t, 395t
plasticity, 242, 244
See also attachment plasticity; experience-dependent plasticity; neural plasticity; use-dependent plasticity
politics, altruism in, 272–73
posttraumatic stress disorder (PTSD)
addiction to stress and self-harm, 339–340
amygdaloid-hippocampal interaction and, 88
BPD and, 315
brain metabolism and, 401
case studies, 372–77
children of Holocaust survivors and, 329
cognitive therapies and, 41
cortical-amygdala circuitry and, 299
divergent hemispheric processing and, 364
DMN and, 158, 163–64, 320
extinction learning in, 300
factors predicting development of, 26
irritation of life and, 166
neural networks and, 42, 340–41, 403
neurochemistry of, 326–28
orienting reflex in, 293
symptoms of, 323–26
traumatic flashbacks and speechless terror, 334–36
traumatic memory, 332–34
treatment options for, 408, 410, 411
See also complex PTSD (C-PTSD)
preconscious memory, 75
predictive capacity, suppression of, 362–63
prefrontal cortex
about, 62, 123, 126–131, 127t, 128t
depression and mania and, 101
empathic attunement and, 125
imagined memory, 82
See also cortical-amygdala circuitry; dorsolateral prefrontal cortex (dlpfc); orbitomedial prefrontal cortex (ompfc)
prejudice, 288
Pribram, K., 87
primates
altruism and, 265

312
amygdala damage in, 312
child care and longevity across species, 250t
distress calls in, 340
DMN and, 161
endorphins and, 208
eye gaze and, 209
facial recognition and facial expressions, 210–11
grasping reflex and, 123
longevity in, 249
mental imagery and, 146
mirror neurons and, 149, 212
neurogenesis in, 59, 60
reflexive social language and, 193
right frontal activity and, 109
social brain and, 201, 202
See also monkeys
primitive reflexes. See grasping reflex
Prins, B., 333
problem solving, 127–28, 133
processing, 370–71
See also divergent hemispheric processing; fear processing
“Project for a Scientific Psychology”(Freud), 5–6
projection, primacy of, 370–71
projective hypothesis, 36–37
protoconversation, 208–9
Prozac, 64, 89
psychoanalytic and psychodynamic therapies, 6, 35–38, 76
See also psychotherapy
psychology history, 3–13
psychopathology, 26–27, 108–9, 162–64, 217, 267
See also specific disorders
psychosis, 111–13
psychotherapists, as neuroscientists, 397–403, 403t–5t, 405–17
psychotherapy
brain functioning that necessitates, 362–378
Buddhism and, 165
common factors among schools of, 47–49
as an enriched environment, 22
goals of, 23–24, 28–29, 49, 76
guided altruism as adjunct to, 273–75
lessons learned from, 202–3
levers of change in, 421–22
parenting and, 31–33
predictors of success in, 180, 369
reasons for seeking, 184, 420
as reparenting, 251
See also interpretations; therapeutic relationship
public speaking, fear of, 178

RAD. See reactive attachment disorder (RAD)


rats
BNST in, 292
effects of stress on, 293, 326–27

313
electrical stimulation in, 298
enriched environments and, 384, 384t
hippocampus in, 255
humans versus, 418
maternal behavior and effects on, 67, 141, 240–47, 248, 328
NE in, 333
neurogenesis in, 60, 89
Rauch, S., 335
RCBF. See regional cerebral blood flow (RCBF)
reactive attachment disorder (RAD), 311–12
reality testing, 114
reciprocity. See attunement and reciprocity
reflective internal language, 194, 400
See also self-reflection
reflexes, 207, 208
See also grasping reflex; orienting reflex
reflexive internal narrator, 193, 399–400
reflexive social language, 193, 194, 399
regional cerebral blood flow (RCBF), 335
regulatory systems, 257
See also vagal system
Reich, W., 19, 45–46
Reichian, Gestalt, and somatic therapy, 45–47
relationships, 40, 203, 251, 319–322
See also mother-infant dyad; social brain; social synapses; therapeutic relationship
relaxation, poor tolerance of, 90
religion, 271–72, 275
See also Buddhism
repression, 179
reptilian brain, 7, 8, 10, 79
resistance, 37, 76–78
resonance behaviors, 257
reward system. See ventral striatum
rhesus monkeys, 246, 311–12
right amygdala, 267
right hemisphere
anxiety and depression and, 109–10
children and development of, 97–98
functions of, 30, 99–100
integration of the body in, 103–4
social brain and, 206
See also hemispheric laterality; left-right hemisphere balance
right-left integration, 108–9
See also left-right integration
rodents, 208
See also mice; rats
Rogerian or client-centered therapy, 38–40
Rogers, C., 32, 38–40, 278, 393, 423
Romanian orphans, 311, 312

safe emergency, 46
salience network and focus of attention, 134, 135t, 147, 159
Sapolsky, R., 293

314
Satterfield, J., 136
scanning techniques, 67–68, 400–401, 405
schizophrenia, 111–12, 162, 175, 185
Schuman, E., 382
Sear, R., 249
secure attachment, 25, 268
See also attunement and reciprocity; mother-infant dyad
seizure activity, 85
See also temporal lobe epilepsy (TLE)
self. See sense of self
self-awareness, 141, 142, 143, 192–95, 370
self-deception. See illusions of consciousness; unconscious self-deception
self-disturbance. See narcissism; out-of-body experiences
self-esteem, 246
self-fragmentation, 316–19
elf-harm. See addiction to stress and self-harm; suicide and suicide attempts
self-reflection, 194–95, 196–97, 370, 400
sense of self
about, 154–55
default-mode network (DMN) and, 158, 161–62
early traumatic stress and, 306
frontal-parietal executive network and, 148
good-enough parenting and, 154
parietal lobes and, 144
See also true versus false self
sensitive periods, 72–73
sensory, motor, and affective systems, 256–57
serotonin, 64, 327
sexual abuse, 329–330
shame, 90, 216–18, 431–33
Sheline, Y., 88
short-term memory deficits, 86, 87
Siddhartha, 164–65
side-to-side eye movements, 410, 411–12
Siegel, D., 226
silence, poor tolerance of, 90
Simon Says, 137
sinus infections, 170–72
Skinner, B., 421
slow fear network, 286, 287f, 288, 297, 299
smart vagus, 61
social brain
about, 10–12, 14–15, 201–3, 251–52, 252t
attunement and, 422–24
evolution and, 201–2
right hemisphere and, 206
structures and systems of, 252–260, 252t, 259t
social engagement system, 258–59, 259t
social fragmentation, 319–322
social motivation system, 260
social synapses, 203–5
Socrates, 433

315
somatic therapy. See Reichian, Gestalt, and somatic therapy
somatization disorder, 30
somatosensory cortex, 254
soothing touch, 393–94
speechless terror. See traumatic flashbacks and speechless terror
The Spirit of Geometry (Magritte), 357f
split-brain research, 98, 105
See also corpus callosum
stage fright, 178
state-dependent memory, 78
Stecklis, H., 202
Steele, H., 232
Steele, M., 232
Stein, M., 296
stimulant medications, 136, 137
stories, 187–88, 425
See also narratives
stress
addiction to, 339
centrality of, 407–12
hippocampus and, 293–97, 295t
HPA axis and, 67
learning and, 22–24
memory deficits and, 78, 86
neural network integration and effects of, 26, 27
recovery of fears and phobias under, 300–301
suppression of language and predictive capacity under, 362–63
treatment options for, 408
See also addiction to stress and self-harm; arousal, moderate states of; early traumatic stress; optimal stress
stress, fear, and attachment system, 257–58
stress hormones, 367–68
See also cortisol; glucocorticoids (GCs)
stress reactivity, 242, 244, 247, 284
stress regulation. See arousal, moderate states of
STZ. See superior temporal zone (STZ)
sublimation, 36
suffering categories, 166
See also pain versus suffering
suicide and suicide attempts, 339–340, 357
suicide victims, brains of, 248
sukha, 165
Suomi S., 234, 235
superego, 432
See also reflexive internal narrator
superior temporal zone (STZ), 269
swimming, learning to, 71
systemic family therapy, 43–45

task-positive networks (TPNs), 184–85


taxon system, 300, 301
temporal cortex and lobes, 61–62, 82, 256–57
See also superior temporal zone (STZ); Wernicke’s area
temporal lobe epilepsy (TLE), 85

316
temporal-parietal junction (TPJ), 268–69
terror. See traumatic flashbacks and speechless terror
theory of mind. See projection, primacy of
therapeutic relationship
in CBT, 42
client self-reflection and, 197
as curative element, 38–39, 423–24
moderate states of arousal and, 387
parenting as compared to, 32
as placebo effect, 392, 395, 395t, 396
as professionalized altruism, 263–64
See also social brain
therapists, 31, 195, 420, 422
See also psychotherapists, as neuroscientists; therapeutic relationship
titi monkeys, 249
TLE. See temporal lobe epilepsy (TLE)
TMS. See transcranial magnetic stimulation (TMS)
top-down integration, 29–30, 31, 71, 87, 178–79
Torchia, M., 296
touch, soothing, 393–94
Tourette’s syndrome, 406
TPJ. See temporal-parietal junction (TPJ)
TPNs. See task-positive networks (TPNs)
transcranial magnetic stimulation (TMS), 180–81
transcription genetics, 20–21
transference, 37, 89
trauma, early life. See early traumatic stress
traumatic brain injuries, 128–131, 413–14
traumatic flashbacks and speechless terror, 334–38
traumatic memory, 332–34
treatment outcomes, 402–3, 403t–5t, 405–7
treatment rationales and combinations, 409–12, 413–15
trial-and-error learning, 71
triune brain, 7–8, 61, 79
true versus false self, 215–16, 352

the unconscious (Freud), 4–5


unconscious memory, 75, 78
unconscious processing, 368–69
unconscious self-deception, 371–72
unknown thought, 369
use-dependent plasticity, 381–82

vagal system, 258–59, 259t


ventral striatum, 260
vervet monkeys, 293

Waddington, C., 66
WAIS. See Wechsler Adult Intelligence Scale (WAIS)
Weber, M., 333
Wechsler Adult Intelligence Scale (WAIS), 147, 148
Wernicke’s area, 99, 105
Williams, L., 325

317
Wilson, D., 105
Winnicott, D., 154, 213–16, 218, 343, 352
women, 160
working memory, 187
working through, 351–52
World Trade Center attacks, memories of, 375–76

Yerkes, M., 333


Yerkes, R., 387

318
Also available from
The Norton Series on Interpersonal Neurobiology
The Birth of Intersubjectivity: Psychodynamics, Neurobiology, and the Self
Massimo Ammaniti, Vittorio Gallese

Neurobiology for Clinical Social Work: Theory and Practice


Jeffrey S. Applegate, Janet R. Shapiro

Being a Brain-Wise Therapist: A Practical Guide


to Interpersonal Neurobiology
Bonnie Badenoch

The Brain-Savvy Therapist’s Workbook


Bonnie Badenoch

Coping with Trauma-Related Dissociation: Skills


Training for Patients and Therapists
Suzette Boon, Kathy Steele, and Onno van der Hart

Neurobiologically Informed Trauma Therapy with Children


and Adolescents: Understanding Mechanisms of Change
Linda Chapman

Intensive Psychotherapy for Persistent Dissociative


Processes: The Fear of Feeling Real
Richard A. Chefetz

The Healthy Aging Brain: Sustaining Attachment, Attaining Wisdom


Louis Cozolino

The Neuroscience of Human Relationships: Attachment


and the Developing Social Brain (Second Edition)
Louis Cozolino

The Neuroscience of Psychotherapy:


Healing the Social Brain (Second Edition)
Louis Cozolino

Why Therapy Works: Using Our Minds to Change Our Brains


Louis Cozolino

From Axons to Identity: Neurological Explorations of the Nature of the Self


Todd E. Feinberg

Loving with the Brain in Mind: Neurobiology and Couple Therapy


Mona DeKoven Fishbane

Body Sense: The Science and Practice of Embodied Self-Awareness


Alan Fogel

The Healing Power of Emotion: Affective Neuroscience,


Development & Clinical Practice
Diana Fosha, Daniel J. Siegel, Marion Solomon

319
Healing the Traumatized Self: Consciousness, Neuroscience, Treatment
Paul Frewen, Ruth Lanius

The Neuropsychology of the Unconscious: Integrating


Brain and Mind in Psychotherapy
Efrat Ginot

10 Principles for Doing Effective Couples Therapy


Julie Schwartz Gottman and John M. Gottman

The Impact of Attachment


Susan Hart

Art Therapy and the Neuroscience of Relationships,


Creativity, and Resiliency: Skills and Practices
Noah Hass-Cohen and Joanna Clyde Findlay

Affect Regulation Theory: A Clinical Model


Daniel Hill

Brain-Based Parenting: The Neuroscience of


Caregiving for Healthy Attachment
Daniel A. Hughes, Jonathan Baylin

The Interpersonal Neurobiology of Play: Brain-Building


Interventions for Emotional Well-Being
Theresa A. Kestly

Self-Agency in Psychotherapy: Attachment, Autonomy, and Intimacy


Jean Knox

Infant/Child Mental Health, Early Intervention, and Relationship-Based


Therapies: A Neurorelational Framework for Interdisciplinary Practice
Connie Lillas, Janiece Turnbull

Clinical Intuition in Psychotherapy: The


Neurobiology of Embodied Response
Terry Marks-Tarlow

Awakening Clinical Intuition: An Experiential


Workbook for Psychotherapists
Terry Marks-Tarlow

A Dissociation Model of Borderline Personality Disorder


Russell Meares

Borderline Personality Disorder and the Conversational


Model: A Clinician’s Manual
Russell Meares

Neurobiology Essentials for Clinicians: What


Every Therapist Needs to Know
Arlene Montgomery

Neurobiology and the Development of Human


Morality: Evolution, Culture, and Wisdom
Darcia Narvaez

Brain Model & Puzzle: Anatomy & Functional Areas of the Brain
Norton Professional Books

320
Sensorimotor Psychotherapy: Interventions for Trauma and Attachment
Pat Ogden, Janina Fisher

Trauma and the Body: A Sensorimotor Approach to Psychotherapy


Pat Ogden, Kekuni Minton, Clare Pain

The Archaeology of Mind: Neuroevolutionary Origins of Human Emotions


Jaak Panksepp, Lucy Biven

The Polyvagal Theory: Neurophysiological Foundations of


Emotions, Attachment, Communication, and Self-regulation
Stephen W. Porges

Affect Dysregulation and Disorders of the Self


Allan N. Schore

Affect Regulation and the Repair of the Self


Allan N. Schore

The Science of the Art of Psychotherapy


Allan N. Schore

The Mindful Brain: Reflection and Attunement


in the Cultivation of Well-Being
Daniel J. Siegel

The Mindful Therapist: A Clinician’s Guide to


Mindsight and Neural Integration
Daniel J. Siegel

Pocket Guide to Interpersonal Neurobiology: An


Integrative Handbook of the Mind
Daniel J. Siegel

Healing Moments in Psychotherapy


Daniel J. Siegel, Marion Solomon

Healing Trauma: Attachment, Mind, Body and Brain


Daniel J. Siegel, Marion Solomon

Love and War in Intimate Relationships: Connection,


Disconnection, and Mutual Regulation in Couple Therapy
Marion Solomon, Stan Tatkin

The Present Moment in Psychotherapy and Everyday Life


Daniel N. Stern

The Neurobehavioral and Social-Emotional


Development of Infants and Children
Ed Tronick

The Haunted Self: Structural Dissociation and the


Treatment of Chronic Traumatization
Onno Van Der Hart, Ellert R. S. Nijenhuis, Kathy Steele

Changing Minds in Therapy: Emotion, Attachment,


Trauma, and Neurobiology
Margaret Wilkinson

321
For complete book details, and to order online,
please visit the Series webpage at wwnorton.com/Psych/IPNBseries

322
Note to Readers: Standards of clinical practice and protocol change over time, and no technique or recommendation is guaranteed to be safe or
effective in all circumstances. This volume is intended as a general information resource for professionals practicing in the field of psychotherapy
and mental health; it is not a substitute for appropriate training, peer review, or clinical supervision. Neither the publisher nor the author can
guarantee the complete accuracy, efficacy, or appropriateness of any particular recommendation in every respect.

Copyright © 2017, 2010, 2002 by Louis J. Cozolino

All rights reserved

For information about permission to reproduce selections from this book,


write to Permissions, W. W. Norton & Company, Inc.,
500 Fifth Avenue, New York, NY 10110

For information about special discounts for bulk purchases, please contact
W. W. Norton Special Sales at specialsales@wwnorton.com or 800-233-4830

Production manager: Christine Critelli


Cover design by Lauren Graessle
Cover image © Tim Flach/Getty Images

The Library of Congress has cataloged the printed edition as follows:

Names: Cozolino, Louis J., author.


Title: The neuroscience of psychotherapy :
healing the social brain / Louis Cozolino.
Other titles: Norton series on interpersonal neurobiology.
Description: Third edition. | New York : W.W. Norton & Company, [2017] |
Series: Norton series on interpersonal neurobiology |
Includes bibliographical references and index.
Identifiers: LCCN 2016048033 | ISBN 9780393712643 (hardcover)
Subjects: | MESH: Neuropsychology | Psychotherapy—methods |
Brain—physiology | Brain Diseases—psychology | Mental
Processes—physiology | Neurobiology—methods
Classification: LCC RC480.5 | NLM WL 103.5 | DDC 616.89/14—
dc23 LC record available at https://lccn.loc.gov/2016048033

ISBN 978-0-393-71265-0 (e-book)

W. W. Norton & Company, Inc.


500 Fifth Avenue, New York, N.Y. 10110
www.wwnorton.com

W. W. Norton & Company Ltd.


15 Carlisle Street, London W1D 3BS

323

You might also like