Sie sind auf Seite 1von 60

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/232594305

Temperament, Attention, and Developmental


Psychopathology

Article January 2006


DOI: 10.1002/9780470939390.ch11

CITATIONS READS

159 705

2 authors:

Mary K Rothbart Michael Posner


University of Oregon University of Oregon
200 PUBLICATIONS 24,601 CITATIONS 436 PUBLICATIONS 75,341 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Human attentional networks View project

An information theory account of cognitive control View project

All content following this page was uploaded by Mary K Rothbart on 08 January 2016.

The user has requested enhancement of the downloaded file.


Temp., Atten., & Dev. Psychopathology 1

Temperament, Attention, and Developmental Psychopathology

Mary K. Rothbart and Michael I. Posner


University of Oregon

Chapter to appear in 2006. In D. Cicchetti & D. J. Cohen (Eds.), Handbook of Developmental


Psychopathology. Wiley Press.
Temp., Atten., & Dev. Psychopathology 2

Cicchetti and Toth (1995) have described two major goals of a developmental approach
to psychopathology. The first is to study the development of individual capacities across the life
span. The second is to examine the relation between outcomes and earlier adaptations in
functioning, considering distal as well as proximal influences on development. A temperament
approach to development and psychopathology is concerned with the constitutionally based
dispositions that underlie the individuals social and emotional development. Temperament also
provides a view of distal influence that is evolutionary in scope, identifying variations in
affective-motivational and attentional adaptations to human life that are both inherited and
shaped by experience. Along with attention, these include affect related to threatening input that
organizes motor and autonomic circuits to support avoidant behavior, and perceptual pathways to
enhance information relevant to threat and safety (Derryberry& Rothbart, 1984; Derryberry &
Tucker, 1992; Gray, 1982; Ohman, 2000). They also include affect and emotion related to
incentives that support approach, and perceptual pathways that enhance information relevant to
reward, as well as affect related to goal blockage and loss that support attack and withdrawal.

Except in rare cases, we all exhibit these adaptations, and share a number of them with
non-human animals (Panksepp, 1998). Individuals vary in the reactivity of their affective
motivational systems and in the self-regulatory capacities, such as attention, that modulate them
(Rothbart & Derryberry, 1981). As the child develops, initially more reactive forms of
regulation such as approach and fearful inhibition are increasingly supplemented by capacities of
voluntary or effortful control (Posner and Rothbart, 1998; Rothbart & Bates, 1998). These
patterns of regulation may, promote a progressive stabilization of synapses within the brain,
contributing to the structural organizations central to personality (Derryberry & Reed, 1994, p.
633). The self-organization of social and emotional development based on temperament and
experience provides a further proximal basis for thinking about psychopathology and its
development (Cicchetti & Tucker, 1994).

In this chapter, we consider relations between temperament and the development of


psychopathology. We define temperament as constitutionally based individual differences in
reactivity and self-regulation, as observed within the domains of emotionality, motor activity,
and attention (Rothbart, 1989c; Rothbart & Derryberry, 1981). By reactivity, we mean
characteristics of the individual's responsivity to changes in stimulation, as reflected in somatic,
autonomic, and endocrine nervous systems. By self-regulation, we mean processes modulating
this reactivity, including behavioral approach, avoidance, inhibition, and attentional self-
regulation. In our view, individual differences in temperament constitute the earliest expression
of personality and the substrate from which later personality develops. By tracing the
development of temperament and the adaptations it supports across different periods of the life
span, we can illuminate our understanding of the development of risk for psychopathology.

In the previous edition of this chapter (Rothbart, Posner, & Hershey, 1995), we identified
a number of ways in which temperament might shape individual adaptation and vice versa.
These include (1) individual differences that in the extreme may constitute the psychopathology
or dispose the person toward it; (2) characteristics evoking reactions in others that can promote
the risk of psychopathology or buffer it; (3) characteristics influencing the persons niche
picking, putting him/her at greater or lesser risk for psychopathology; (4) temperamental
influences on the form of a disorder, its course, and the likelihood of its recurrence, (5)
Temp., Atten., & Dev. Psychopathology 3

temperamental biases in information processing about the self and the world, increasing or
decreasing the likelihood of psychopathology; (6) temperamental regulation or buffering against
the effects of risk factors or stress; (7) temperamental heightened responsiveness to
environmental events; (8) interaction among temperament systems, some of them later
developing. Two major temperamental control systems are fear, developing late in the first year,
inhibiting approach and expressive tendencies, and effortful control, developing during the
preschool years and beyond. In addition, (9) temperamental dispositions may shape different
developmental pathways to a given outcome, and individual dispositions may allow for multiple
outcomes (Cicchetti & Cohen, 1995). Finally, (10) temperament characteristics and caregiving
environment may make independent contributions to outcomes, or may interact in increasing or
decreasing risk of disorder; and (11) disorders themselves may change aspects of temperament.
We have been impressed at the number of studies during the past decade that speak to these
issues, and we revisit them in the course of our discussion.

This chapter has two major sections. In the first, we describe findings on individual
differences in temperament and attention and consider their links to neural networks. We then
discuss theories and research relating temperament to the development of risk conditions for
psychopathology. Individual differences in attentional capacities, a major aspect of
temperament, are critical to our understanding of these issues.

We begin the chapter with a brief historical introduction, followed by a discussion of


dimensions of temperament that have emerged from conceptual and factor-analytic studies of
temperament early in life (Bates, Freeland, & Lounsbury, 1979; Hagekull & Bohlin, 1981;
Rothbart, 1981; Sanson, Prior, Garino, Oberklaid, & Sewell, 1987). These dimensions include
fear (distress and behavioral inhibition to novelty), anger or irritability, positive affect and
approach, activity level, and attentional persistence. We note similarities between these
dimensions and the "Big Five and "Big Three" personality factors identified in research on
adults (Goldberg, 1990). We then consider temperament dimensions in connection with models
for developmental psychopathology, and review recent research on temperament, attention, and
the development of psychopathology.

HISTORICAL APPROACHES

The Greco-Roman Typology

The ancient Western typology of temperament was developed by Greek and Greco-
Roman physicians, including Theophrastus, Vindician, and Galen (Diamond, 1974). Like later
approaches to temperament, the fourfold typology related individual differences in behavior and
emotionality to variability in the human constitution as it was understood at the time. The
temperamental syndrome of melancholia (a tendency to negative and depressed mood) and its
relation to a predominance of black bile was developed in the 4th century B.C. writings of
Theophrastus, and the complete fourfold typology of temperament as linked to the Hippocratic
bodily humors was present by the time of Vindician in the 4th century A.D. (Diamond, 1974).

In addition to the melancholic type, the fourfold typology included descriptions of the
choleric person (quick-tempered and touchy, easily aroused to anger, with a predominance of
Temp., Atten., & Dev. Psychopathology 4

yellow bile), the phlegmatic (apathetic and sluggish, not easily stirred to emotion, with a
predominance of phlegm), and the sanguine person (characterized by warmth, optimism, and
lively expressiveness, with a predominance of blood). The fourfold temperament typology was
endorsed throughout the Middle Ages and Renaissance (cf. Burton, 1621/1921; Culpeper, 1657)
and discussed by Kant (1798/1978). More recently, Eysenck's (1970) work made use of the
typology, and Merenda (1987) has argued that it constitutes a four-factor model for our current
understanding of temperament and personality. Recently, Watson and Clark (1995) discussed
the melancholic temperament in connection with links between negative emotionality and
psychopathology.

If we look at woodcuts from the Middle Ages, representing the four temperament types,
the behaviors depicted often do appear pathological. The choleric man is depicted beating a
woman, the phlegmatic person is asleep in bed while others are working, and the melancholic
person is seen in depressed attitude or portrayed as a distraught lover (e.g., Carlson, 1984, p.
613). The sanguine type is more often depicted as adjusted than the other three, but even the
sanguine individual may be shown overeating or overdrinking. These images suggest that, at the
extreme, individual differences in temperament may in themselves constitute maladaptations.
One of the issues immediately raised by these historical treatments of temperament and
psychopathology is thus the relation between a temperamental extreme and a psychopathological
condition.

The Typological Approach of Carl Jung

In this century, the type construct has been modified conceptually, resulting in
dimensional and factor-analytically derived approaches to an understanding of individual
differences. Even the psychological types put forward by Jung (1923) differ in important ways
from the fourfold typology. To Jung, introverted and extraverted qualities were predominant
human attitudes, related to thinking, feeling, sensation, and intuition. Unlike the ancients, or
Kagan's (1989) view of inhibited and uninhibited types as qualitatively different classes of
individuals, Jung (1923) did not view psychological types as representing qualitative differences.
Instead, he described them as Galtonesque family-portraits, which sum up in a cumulative
image the common and therefore typical characters, stressing these disproportionately, while the
individual features are just as disproportionately effaced (Jung, 1923, p. 513).

Jung's approach to the type is similar to the idea of the prototype in cognition (Posner &
Keele, 1968; Rosch, 1973); it is also related to principles of measuring the individual trait. In
deriving a score for an individual on a temperament or personality trait, the individual's item
scores on a scale are aggregated to describe a single score that represents similarity of response
across time and situations. This trait score can be seen as the individuals "common and
therefore typical" characteristics, "stressing these disproportionately"; the influence of particular
situations or other important factors is also "just as disproportionately effaced." The type can
then be seen to refer to a set of traits that tend to occur together in groups of individuals. Factor-
analytic approaches to temperament and personality identify clusters of characteristics that tend
to covary; results of this work are described in more detail below.
Temp., Atten., & Dev. Psychopathology 5

Jung also suggested that introverted and extraverted tendencies are present in all persons,
but, for a given person, one attitude tends to become more elaborated and conscious, while the
other is less elaborated, more primitive, and, for the most part, unconscious. Differentiation
between extraversion and introversion, he wrote, can be seen early in life:

The earliest mark of extraversion in a child is his quick adaptation to the


environment, and the extraordinary attention he gives to objects, especially to his
effect upon them. Shyness in regard to objects is very slight; the child moves and
lives among them with trust. He makes quick perceptions, but in a haphazard
way. Apparently he develops more quickly than an introverted child, since he is
less cautious, and as a rule, has no fear. Apparently, too, he feels no barrier
between himself and objects, and hence he can play with them freely and learn
through them. He gladly pushes his undertakings to an extreme, and risks himself
in the attempt. Everything unknown seems alluring (Jung, 1928, p. 303).

By objects, Jung included both physical and social entities, so that the more introverted
child would show a tendency to dislike new social situations, approaching strangers with caution
or fear, and the introverted adult to dislike new situations or social gatherings, and demonstrating
hesitation and reserve. Jung suggests that the introvert would also be inclined toward pessimism
about the outcome of future events. The extraverted person would show more ready approach
and action toward social and physical objects (impulsivity), and greater sociability and optimism
about the future (Jung, 1923).

Jung also related extraversion-introversion to psychopathology. He took over a


distinction between two types of neurosis put forward by Janet, and developed the idea that the
extravert is predisposed to hysteria, characterized by an exaggerated rapport with the members
of his circle, and a frankly imitatory accommodation to surrounding conditions (Jung, 1923, p.
421), combined with a tendency toward experiencing somatic disorders. The introvert was seen
as prone to psychasthenia, a malady which is characterized on the one hand by an extreme
sensitiveness, on the other by a great liability to exhaustion and chronic fatigue (p. 479).

Factor-Analytic Studies

Extraversion-introversion also emerged from factor-analytic studies of temperament in


adults. Early factor-analytic studies were carried out in Great Britain by Webb and Burt. Webb
(1915) analyzed items assessing emotionality, activity, qualities of the self, and intelligence, and
identified two factors. One he labeled w, defined as consistency of action resulting from
deliberate volition or will (Webb, 1915, p. 34). This factor bears similarities to the higher-order
personality factor recently labeled control, constraint, or conscientiousness (Digman & Inouye,
1986) and to temperamental effortful control. A second factor assessed negative emotionality,
sometimes labeled emotional stability-instability; Eysenck would later call it neuroticism. By
1938, Burt had also identified the factor of extraversion-introversion. Later factor-analytic work
has repeatedly identified three factors similar to these: extraversion (the first factor that is usually
extracted from a large data set of personality descriptors), neuroticism, and conscientiousness
(Costa & McCrae, 1988; Goldberg, 1990; John, 1989). These three factors, with the additional
factors of agreeableness and openness to experience, as derived from factor analyses on trait
Temp., Atten., & Dev. Psychopathology 6

descriptive words and personality items, constitute what have been called the Big Five
personality factors.

In Eysenck's (1944, 1947) early factor-analytic work on male neurotic subjects, factors of
extraversion-introversion and neuroticism were extracted. He also reported support for Jung's
distinction between more extraverted (hysteric) symptoms and more introverted (dysthymic or
psychasthenic) symptoms in neurosis. Eysenck (1947) also reviewed results of a factor analysis
of Ackerson's (1942) data describing a large sample of children studied by the Illinois Institute
for Juvenile Research. These data yielded a general neuroticism factor, along with a factor that
distinguished between more introverted behavior problems (sensitive, absent-minded, seclusive,
depressed, daydreams, inefficient, queer, inferiority feelings, and nervous), and extraverted
behavior problems (such as stealing, truancy from home and school, destructive, lying, swearing,
disobedient, disturbing influence, violent, rude and egocentric) (Eysenck & Eysenck, 1985, p.
53).

There has since been extensive replication of this factor structure of psychopathology in
childhood, distinguishing between internalizing problems, including inhibition, shyness, and
anxiety related to introversion, and externalizing problems, including aggressive and acting-out
problems related to extraversion (Achenbach & Edelbrock, 1978; Cicchetti & Toth, 1991).
Results of these analyses, combined with evidence for stability of extraversion-introversion
across long periods (see reviews by Bronson, 1972; Kagan, 1998; and Rothbart, 1989a), suggest
that individual differences on extraversion-introversion may be basic to both a description of
temperament and, in their extremes, to an individual's likelihood of demonstrating particular
kinds of behavior problems or psychopathology. Below, we describe some direct and indirect
links between temperament, caregiving, and risk for psychopathology. First, however, we
discuss the structure of temperament.

TEMPERAMENT DIMENSIONS IN DEVELOPMENT

Much of the early research and thinking in the field of temperament in childhood has
relied on Thomas and Chess's pioneering work in the New York Longitudinal Study (NYLS;
Thomas & Chess, 1977; Thomas, Chess, Birch, Hertzig, & Korn, 1963). Dimensions of
temperament identified by the NYLS included activity level, threshold, mood, rhythmicity,
approach/withdrawal, intensity, adaptability, distractibility, and attention span/persistence.
Chess and Thomas (1986; Thomas, Chess, & Birch, 1968) also developed the important idea of
"goodness-of-fit" to describe situations in which there is a "match" between the temperamental
characteristics of the child and the expectations of others and/or demands of the situation. In
their model, a good fit would predict a more favorable mental health outcome; a poor fit or
incompatibility would predict impaired function and risk of development of behavior disorders.

In their case studies of children who survived an upbringing in multiproblem families


with mental illness in one or more parents, Radke-Yarrow and Sherman (1990) identified two
key protective factors. The first is very similar to Chess and Thomass (1986) goodness-of-fit
idea: One key factor is a match between a psychological or physical quality in the child and a
core need in one or both of the parents that the child fulfills (p. 112). The second, an extension
of the first, is: the child's clear conception that there is something good and special about
Temp., Atten., & Dev. Psychopathology 7

himself or herself. The child quality is then a source of positive self-regard for the child, as well
as need-satisfying to the parent (p. 112).

Child qualities providing a match or mismatch with parental need could be gender,
appearance, intellectual or physical aptitude, or temperament. Within American society
generally, a positive and outgoing temperament may represent a particularly auspicious "match"
between the child and many situational requirements; in school, effortful and/or fear control may
provide a better match (Lerner, Nitz, Talwar, & Lerner, 1989). In other cultures, reserve rather
than extraversion may be more generally valued (Chen, Rubin, & Li, 1995: Chen, Rubin, & Sun,
1992).

Difficulty in the NYLS Typology

Thomas, Chess, and associates (Thomas et al., 1963, 1968) also identified groups of
children they designated as easy, difficult, and slow to warm up, and their construct of
difficultness, has been frequently used in studies of the development of behavior problems.
This construct stemmed from results of a factor analysis of data from the original NYLS, which
identified a cluster of five temperament dimensions: approach/withdrawal, mood, intensity,
adaptability, and rhythmicity, labeled as difficult (Thomas et al., 1963). In algorithms for
assessing difficulty, developed by Carey and McDevitt (1978), Fullard, McDevitt, and Carey
(1984), and others, difficulty is derived from extreme scores on all five dimensions.

There are several problems with the use of the difficulty construct in temperament
research (For a discussion of this issue, see Plomin, 1982; Rothbart, 1982.). One is the wide
variability of operationalizations of the construct. Because factor analyses of data collected on
the NYLS dimensions have frequently failed to yield loadings for rhythmicity in the "difficulty"
factor (Bates, 1989), these studies often delete rhythmicity measurements or use varying
composites, depending on the outcome of the factor analysis or the particular measures used in
the study. Those based on the Carey and McDevitt (1978) or Fullard et al. (1984) algorithms
always include it. Inconsistency in the use of difficulty as a construct then creates serious
problems for knowing what is meant by difficulty in any given study, and greater precision will
result only when assessments of the construct are psychometrically sound and comparable across
studies.

Bates (1980, 1989) argued that the core variable of measures of difficulty is negative
emotionality, and his work used measures assessing this variable. Factor-analytic work by Bates
and others also identified two kinds of negative emotionality in infancy: the first is related to
children's responses to novelty and the second, more general distress proneness (Bates et al.,
1979). Bates's research indicated that the two varieties of negative emotionality may be
differentially related to specific behavior problems that develop later, and differentiated
temperamental contributions to behavioral problems and psychopathology have consistently
been found in recent research, as indicated below and in Rothbart and Bates (in press). This
differentiation is lost in the more general NYLS-based difficulty construct.

Another way of looking at difficulty is to say that some children are labeled by their
caregivers as problem children at an early age, and this labeling is likely influenced by the
Temp., Atten., & Dev. Psychopathology 8

childs temperament. Use of the terms easy and difficult may be further extended as a
parent has more than one child (the first was easy, but the second was difficult). The terms
may be applied in an endearing way, or used to indicate the source of problems in family
relationships. Attributions about the child may in turn contribute to developmental outcomes in
the childs adaptation. Nevertheless, social attributions and values take us to a different level of
analysis than measurement of temperamental dispositions in the child.

Revised Dimensions of Temperament in Infancy

In the original clinical application of the NYLS work (Thomas et al., 1968), no attempts
were made to avoid conceptual overlap across the nine dimensions, so that individual
questionnaire scales based on the NYLS have sometimes been highly overlapping in meaning.
Findings of relatively high intercorrelations among scale scores designed to assess the NYLS
dimensions have not been too surprising, given that a behavioral item might, in some cases,
belong as easily to any one of three different scales (Rothbart & Mauro, 1990). Lack of internal
homogeneity within individual scales based on some of the NYLS dimensions has also been
found, which, when combined with lack of independence among scales, has led researchers in
Sweden and Australia to attempt item-level factor analysis of NYLS-based parent report scales
(Hagekull & Bohlin, 1981; Sanson et al., 1987).

Results of item-level factor-analytic in turn led to identification of a smaller number of


factors that correspond well to dimensions of temperament derived from other developmentally
based theoretical approaches to temperament (Buss & Plomin, 1975, 1984; Rothbart &
Derryberry, 1981). These factors include distress and inhibition to novelty (including behavioral
inhibition), other distress proneness or irritability, positive affect and approach, activity level,
and persistence (Rothbart & Mauro, 1990).

Gartstein and Rothbart (2003) recently studied the factor structure of expanded scales
measuring parent-reported infant temperament, adapted from dimensions identified in research
on temperament in childhood. Factor analysis of a large data set describing 3- to 12-month-old
children yielded three broad dimensions: Surgency/Extraversion, defined primarily from scales
of approach, vocal reactivity, high intensity pleasure (stimulation seeking), smiling and laughter,
activity level and perceptual sensitivity; Negative Affectivity, with loadings from sadness,
frustration, fear, and loading negatively, falling reactivity; and Orienting/Regulation, with
loadings from low intensity pleasure, cuddliness, duration of orienting and soothability, and a
secondary loading for smiling and laughter. As early as infancy, there is thus evidence for three
broad temperament dimensions.

Dimensions of Temperament in Childhood

We have also developed a highly differentiated and comprehensive parent report


instrument called the Childrens Behavior Questionnaire, or CBQ (Rothbart, Ahadi, Hershey, &
Fisher, 2001). Across a number of data sets using the CBQ, three temperament systems of
Surgency/Extraversion, Negative Affectivity, and Effortful Control have emerged from research
on temperament in children 3-7 years of age (Lonigan & Phillips, 2001; Rothbart, Ahadi, &
Hershey, 1994; Rothbart et al., 2001). The Surgency factor is primarily defined by scales
Temp., Atten., & Dev. Psychopathology 9

assessing positive emotionality and approach, including positive anticipation, high intensity
pleasure (sensation-seeking), activity level, impulsivity, smiling and laughter, and a negative
loading from shyness. The Negative Affectivity factor involves positive loadings for shyness,
discomfort, fear, anger/frustration, and sadness, and a negative loading from soothability-falling
reactivity. The Effortful Control Factor is defined by positive loadings from inhibitory control,
attentional focusing, low intensity pleasure (non-risk taking pleasure), and perceptual sensitivity.
These broader factors suggest a hierarchical structure to temperament, with more narrow
constructs related to each other at a higher level in the hierarchy. Questions regarding breadth of
temperamental predictors of psychopathology will be addressed later in the chapter.

The Big Five Factors of Personality and Temperament

These childhood temperament factors conceptually and empirically map fairly well upon
the Extraversion/Positive Emotionality, Neuroticism/Negative Emotionality, and
Conscientiousness/Constraint dimensions found in Big Five studies of the adult personality
(Ahadi & Rothbart, 1994; Rothbart, Ahadi, & Evans, 2000). These broad temperament constructs
further suggest that temperament dimensions go beyond lists of unrelated traits and generalized
characteristics of positive and negative emotionality. Particularly important are temperament-
based interactions between the childs motivational impulses and his or her efforts to constrain
them.

We (Ahadi & Rothbart, 1994; Rothbart, 1989d; Rothbart & Ahadi, 1994) and Martin and
Presley (1991) have indicated possible relationships between temperament factors identified for
childhood and personality factors identified for adults, and in our adult research program, we
have found empirical links between Negative Affectivity and big five Neuroticism,
Surgency/Extraversion, and Extraversion, and Effortful Control and Conscientiousness (Evans &
Rothbart, 2004; Rothbart et al., 2000).

In the next section, we describe conceptual and neural models developed to enhance our
understanding of these dimensions. The first two dimensions are positive affectivity and
approach (Surgency/Extraversion) and fear (distress and latency to approach novel or
challenging stimuli). The third dimension, anger/irritability, is seen, along with fear, discomfort,
and sadness to represent part of a general susceptibility to negative affect. These dimensions are
particularly interesting because they are related to differences observed in non-human animals
(Gosling & John, 1999; Panksepp, 1998), and to factor structures extracted from studies of
personality in adults. The fourth dimension, effortful control, will be further discussed in
connection with a neural model for individual differences in alerting, orienting, and executive
attention (Posner & Fan, in press). We begin our more specific discussion with a consideration
of positive affectivity/approach and fear.

NEURAL MODELS

Cloninger (1986, 1987), Gray (1979, 1982), LeDoux (1987, 1989), Panksepp (1982,
1986, 1998), and Zuckerman (1984) have all made contributions to the development of neural
models for temperament (Rothbart, Derryberry, & Posner, 1994). LeDoux identifies emotions as
broadly integrative systems ordering feeling, thought, and action. In his analysis, emotions are
Temp., Atten., & Dev. Psychopathology 10

seen to be the output of information processing networks assessing the meaning or affective
significance of events for the individual (LeDoux, 1989). Whereas object recognition systems
and spatial processing systems address the questions, What is it? and Where is it?,
respectively, neural emotion processing networks give answers to the questions, Is it good for
me? and Is it bad for me?, in turn influencing the organism's behavioral answers to the
questions, What shall I do about it? or simply, What shall I do?

In the neural processing of emotion, thalamic connections route information about object
qualities of a stimulus through sensory pathways, while simultaneously routing information for
evaluative analysis to the limbic system and the amygdala, where memories of the affective
meaning of the stimulus further influence the process (LeDoux, 1989). Later stages of object
processing update the emotional analysis based on early sensory information, but, in the
meantime, back projections from the amygdala influence the subsequent sensory processing of
the stimulus. Output of the amygdala to organized autonomic reactions via the hypothalamus
and to motor activation via the corpus striatum constitutes the motivational aspect of the
emotions.

In this view, emotional processing can be seen as information processing of a special sort.
Attentional neural networks can then act on emotional information similarly to their action on
output of other data processing systems, including those for object recognition, language, and
motor control, each of which in turn has its own underlying neural networks (Posner & Fan, in
press; Posner & Petersen, 1990). Neuroimaging studies demonstrate connections between
emotional processing networks and the anterior attention system, including the anterior
cingulate, that allow attentional influence on the selection of emotional information for
conscious processing, a consequence of which is that we may or may not be aware of our
emotional evaluations (Bush, Luu, & Posner, 2000; Posner & Rothbart, 1992). However, the
attentional controls on emotion may be somewhat limited and difficult to establish. Panksepp
(1998) lays out anatomical reasons why the regulation of emotion may pose a difficult problem
for the child as follows:

One can ask whether the downward cognitive controls or the upward emotional controls
are stronger. If one looks at the question anatomically and neurochemically, the evidence
seems overwhelming. The upward controls are more abundant and
electrophysiologically more insistent: hence one might expect they would prevail if push
came to shove. Of course, with the increasing influence of cortical functions as humans
develop, along with the pressures for social conformity, the influences of the cognitive
forces increase steadily during maturation. We can eventually experience emotions
without sharing them with others. We can easily put on false faces, which can make the
facial analysis of emotions in real-life situations a remarkably troublesome business.
(Panksepp, 1998, p. 319)

Attentional systems can act on and influence conscious aspects of emotional analyses,
and emotion also influences the focusing and shifting of attention (Derryberry & Reed, 1994,
1996; Gray, 1982). An important aspect of social adaptation involves the appropriateness of the
persons social interaction and related acceptance by others (Parker & Asher, 1987). Information
about the state of others is an important contributor to appropriate social action, and failure of
Temp., Atten., & Dev. Psychopathology 11

access of this information to action and consciousness can be a critical element in the
development of disordered functioning (Blair, Jones, Clark, & Smith, 1997). Focusing attention
on threatening stimuli or on the self also may make access to information about others less
accessible. These are important examples of information processing aspects of temperament,
with implications for social development and psychopathology.

Positive Affectivity and Approach

We now briefly consider neural models developed to describe a physical substrate for
approach and inhibition. Based on animal research, Gray (1979, 1982) described the Behavioral
Activation System (BAS), involving sensitivity to rewards, and the Behavioral Inhibition System
(BIS), involving sensitivity to punishment, nonreward, novelty, and innate fear stimuli. These
two systems are seen as mutually inhibitory, with their balance determining degrees of
extraversion-introversion. Gray also posited a fight-flight system moderating unconditioned
punishment (Gray, 1971, 1975). According to Grays model of the BAS, reward-related
projections from the amygdala to the nucleus accumbens activate the next step in a motor
program that maximally increases proximity to the desired stimulus. Converging dopaminergic
projections to the accumbens enable the switch to this neuronal set, facilitating goal-oriented
behavior (Gray, 1994; Gray & McNaughton, 1996).

In a broader Behavioral Facilitation System (BFS), Depue and Collins (1999) propose
a circuit involving the nucleus accumbens, ventral pallidum, and dopaminergic neurons that
codes the intensity of the rewarding stimuli, while related circuits involving the medial orbital
cortex, amygdala, and hippocampus integrate the salient incentive context. Individual
differences in the functioning of this network are thought to arise from functional variation in
dopaminergic projections that encode the intensity of the incentive motivation and facilitate
contextual processing across multiple limbic and frontal sites. As development proceeds,
dopaminergic facilitation helps to stabilize synaptic contacts within the medial orbital and limbic
circuits, and thus to enhance responsivity to positive incentive stimuli (Depue & Collins, 1999).

Depue and Iacono (1989) originally posited the BFS to help account for bipolar affective
disorders. The BFS involves the individual's initiation of locomotor activity, incentive-reward
motivation, exploration of environmental novelty (if the stimulus and its context do not induce
opposing fear reactions), and irritable aggression. In Depue and Iaconos model, nucleus
accumbens DA activity appears to modulate the flow of motivational information from the
limbic system to the motor system, and thereby contributes to the process of initiating locomotor
activity, incentive, and goal-directed behavior (1989, p. 470). Panksepp (1982, 1986a, 1998)
also reviewed the literature on DA effects, concluding that the general function of DA activity
in appetitive behavior is to promote the expression of motivational excitement and anticipatory
eagernessthe heightened energization of animals searching for and expecting rewards (1986,
p. 91). Cloninger (1986, 1987) specified a novelty-seeking dimension related to DA functioning,
as did Zuckerman (1984) in his dimension of sensation seeking (see review by Rothbart, 1989b).

Tellegens (1985) factor-analytic research on personality in adults identified positive


emotionality, a broad factor that includes positive affect and anticipation, and Watson and Clark
(1997) have argued that there is a positive affect core to individual differences in Extraversion.
Temp., Atten., & Dev. Psychopathology 12

In research on infants (Rothbart, 1988; Rothbart, Derryberry, & Hershey, 2000), we have found
expressions of positive affect (smiling and laughter) to be positively related to infants' rapid
latency to approach objects, and to predict anticipatory eagerness about upcoming positive
events when the same children were 7 years old.

In recent fMRI research, Canli et al. (2001) found that the brains activation to positive
and negative pictures varied depending on the subjects temperament or personality. Persons
higher in extraversion showed greater brain response to positive than to negative stimuli in
widespread frontal, temporal, and limbic activation of both hemispheres. Subjects higher in
neuroticism reacted more to the negative than to the positive stimuli, and showed more
circumscribed activation (fronto-temporal on the left side) and deactivation in a right frontal
area. These findings also demonstrated a positive emotional bias for extraverts and a negative
bias for those high in neuroticism.

A quite different set of brain areas proved to be active in studies of persistence, a


dimension of personality conceptually related to effortful control in temperament (Gusnard et al.,
2003). The effects of persistence acted strongly on midline and lateral prefrontal areas that are
quite different than those found active for positive and negative affect, suggesting regulatory
aspects of persistence. An increasingly popular view is that effortful control (persistence) is
represented in midline frontal areas acts to regulate brain areas like the amygdala that are more
clearly related to reactive aspects of negative affect (Posner & Rothbart, 2000).

Fear and Behavioral Inhibition

Factor analyses of infant temperament questionnaires have reliably yielded two distress
factors: one involving distress to novelty, and the other, irritability, including distress to
limitations or frustration (Rothbart & Mauro, 1990). In our recent revision of the Infant
Behavior Questionnaire, we have also found a higher order negative emotionality factor that
includes fear and frustration (Gartstein & Rothbart, 2003). The distress to novelty factor is
linked to an extended latency to approach new objects (Rothbart, 1988), and a combination of
behavioral inhibition and distress proneness to novelty appears to correspond to the introverted
pole of a broader extraversion-introversion dimension.

Fear is an emotional response activated in the presence of threat or signals of upcoming


danger, and its function appears to be a defensive one. Fear activation is accompanied by
inhibition of ongoing motor programs and preparation of response systems controlling coping
options such as fleeing, fighting, or hiding (see review by Rothbart, Derryberry, et al., 1994). In
Grays (1991) model of the Behavioral Inhibition System (BIS), the fear-related BIS is based
upon circuits including the orbital frontal cortex, medial septal area, and the hippocampus.
However, the amygdala has been more often identified as the critical structure in the processing
of conditioned fear (LeDoux, 1989).

Emotional networks involving the amygdala also appear to respond more strongly to
novel than to familiar stimuli (Nishijo, Ono, & Nishino, 1988). Lesions of the amygdala in
rodents disrupt autonomic and cortisol reactions, behavioral freezing and fear vocalizations, and
similar findings have been reported in primates (Lawrence & Calder, in press). In humans,
Temp., Atten., & Dev. Psychopathology 13

functional neuroimaging studies by Calder, Lawrence, & Young (2001) and others support
involvement of the amygdala in both acquiring and expressing fear, although not in the voluntary
production of facial expressions of fear (Anderson & Phelps, 2002). The amygdala also is
involved in the recognition of fear in the human face (Calder et al., 2001), and there is evidence
in humans with amygdalar damage of reduced fear experience (Adolphs et al., 1999;
Sprengelmeyer et al., 1999).

Projections from the amygdala implement autonomic and behavioral components of fear,
including startle, motor inhibition, facial expression, and cardiovascular and respiratory change
(Davis, Hitchcock, & Rosen, 1987). Individual variability in the structure and functioning of any
of these subsystems may be related to variations in behavioral expressions of fear, and multiple
components of other affective motivational systems such as approach/positive affect and
anger/irritable distress would also be expected.

The amygdala also appears to affect information processing within the cortex. For
example, the basolateral nucleus of the amygdala projects to frontal and cingulate regions
involved in the executive attention system (Posner & Petersen, 1990), as well as to ventral
occipital and temporal pathways involved in processing object information. These connections
are consistent with findings that anxious individuals show enhanced attention to threatening
sources of information (e.g., Derryberry & Reed, 1994).

A psychobiological analysis of fear suggests there may be less disagreement about


temperament variables than had been previously thought. Fear includes components of arousal,
felt emotion, motor response preparation for flight and/or attack (with responses often inhibited),
as well as attention toward the fear-inducing stimulus and/or possible escape routes (Davis et al.,
1987). When temperament researchers study the construct, they sometimes stress its
motivational aspects (e.g., Thomas & Chess's [1977] approach/withdrawal; Kagan's [1994]
behavioral inhibition), its distress proneness aspects (Buss & Plomin's [1975] emotionality;
Goldsmith & Campos' [1982] fear), its duration and susceptibility to interventions (Rothbart's
[1981] soothability), its relation to arousal (Strelau's [1983] reactivity), or multiple components
of response (Rothbart & Derryberry's [1981] fear). If we take the broader view of fear as
suggested by the neuroscience analysis, however, agreement is more evident. Intercorrelations
among scales measuring the differently named constructs further support this contention
(Goldsmith, Rieser-Danner, & Briggs, 1991).

Differences in cerebral hemispheric activation have been related to temperamental


tendencies toward approach versus inhibition-withdrawal. Evidence from electrophysiological
(EEG) and lesion studies has related higher anterior left hemisphere activation in response to
stimulation to increased positive affect and/or decreased negative affect (see reviews by
Davidson et al., 2003; Davidson & Tomarken, 1989). The reverse relationships higher anterior
right hemisphere activation related to higher negative affect and/or decreased positive affect
have also been reported. Resting EEG asymmetries have also been related to positive and
negative emotional reactivity (e.g., Davidson & Fox, 1989; Tomarken, Davidson, Wheeler, &
Doss, 1992). Harmon-Jones and Allen (1997), for example, reported greater left than right-
frontal cortical activity in women subjects with higher scores on Carver and Whites (1994) BAS
questionnaire. The BIS measure was not related to asymmetry. Buss et al. (2003) found
Temp., Atten., & Dev. Psychopathology 14

relations between extreme right EEG asymmetry and basal and reactive cortisol, replicating
primate findings from Kalin and his colleagues (Kalin, Larson, Shelton, & Davidson, 1998).

Fox, Calkins, and Bell (1994) found that infants with stable right frontal EEG asymmetry
between 9 and 24 months of age displayed more fearfulness and inhibition in the laboratory than
other children. At four years, children who showed more reticence and social withdrawal were
also more likely to show right frontal asymmetry. Calkins, Fox, and Marshall (1996) also
reported that children selected for high motor activity and negative affect to laboratory
stimulation at 4 months showed greater right frontal asymmetry at 9 months, greater mother
reports of fear at 9 months, and more inhibited behavior at 14 months. However, no concurrent
relation was found between behavior and frontal asymmetry at 9 and 14 months, and greater
activation of both right and left frontal areas was related to higher inhibition scores at 14 months.
Calkins et al. (1996) suggest a need to differentiate between fearful and angry distress, as
discussed above in the Structure of Temperament section. They also hypothesize that infant high
motor/high negative affect and high motor/high positive affect may be associated with different
kinds of problems developing later. For high motor/high positive affect, the problems would be
associated with difficulties in control.

Anger/Irritability and Aggressive Behavior

In Gray's (1991) model, a Fight/Flight system is constituted by circuits connecting the


amygdala, ventromedial nucleus of the hypothalamus, central gray region of the midbrain, and
somatic and motor effector nuclei of the lower brain stem, processing information involving
unconditioned punishment and non-reward. When there is detection of painful or frustrating
input, the brain stem effectors produce aggressive or defensive behavior. Panksepp (1982)
discusses similar neural circuitry in terms of a "rage" system (see review by Rothbart,
Derryberry, et al., 1994).

Important distinctions have more recently been made among different varieties of
aggression and anger, in relation to their underlying neural systems. Aggression as a self-
defense reaction seems to be based upon the functioning of the same amygdala circuits as
involved in the production of fear (Blanchard & Takahashi, 1988). Aggression linked to
protection of resources, competition, and offensive aggression, on the other hand, involves a
different neural system based on the monoamine Dopamine (DA; Lawrence & Calder, in press).
The DA system has been linked to both the production of offensive aggression (Smeets &
Gonzlez, 2000), and to the recognition of anger in the human face (Lawrence, Calder,
McGowan, & Grasby, 2002). In Lawrence et al.s study, DA blockade impaired the recognition
of anger, while recognition of other emotions and of facial identity was spared.

Depue and Iacono (1989) suggest the dopaminergic system may facilitate irritable
aggression aimed at removing a frustrating obstacle, consistent with findings that DA agonists
(e.g., amphetamine) can enhance aggressive behaviors. Their view suggests links between
approach and frustration/anger, and, in fact, infants and childrens activity level and anger have
been consistently positively related in parent-reported temperament (Rothbart & Derryberry,
2002). Moreover, infant activity level predicts not only positive emotionality at age seven, but
also anger/frustration and low soothability-falling reactivity (Rothbart, Derryberry, et al., 2000).
Temp., Atten., & Dev. Psychopathology 15

Together with findings relating seven-year Surgency to aggression (Rothbart, Derryberry, et al.,
2000), this suggests that strong approach tendencies may be linked to negative as well as positive
emotions (Derryberry & Reed, 1994; Rothbart, Derryberry, et al., 2000). Children who quickly
grasped objects during infancy showed seven-year high levels of positive anticipation and
impulsivity, as well as high anger-frustration and aggression, again suggesting that strong
approach tendencies can contribute to later specific negative emotions as well as to positive
emotionality.

Negative Emotionality or General Distress Proneness

Negative Emotionality or distress proneness is often viewed as a general dimension


subsuming emotions including fear, anticipatory anxiety, sadness, frustration/anger, guilt, and
discomfort. For example, the Five-Factor model of adult personality includes negative
emotions as components of the Neuroticism superfactor. Neuroticism/Negative Emotionality has
been found to be orthogonal to Extraversion/Positive Emotionality (Eysenck & Eysenck, 1985;
Tellegen, 1985, Watson & Clark, 1992). As evident in our previous discussion, however, the
positive relationship between anger/frustration and strong approach tendencies suggests that a
more differentiated model is needed.

There are several possibilities for neural systems supporting higher order negative
emotion reactions. One is the link between systems supporting fear and self-defensive
aggression (Blanchard & Takahashi, 1988). Defensive aggression in animal models seems to be
based on the same amygdalar circuits as fear, and in humans, anger in response to threat may
also be linked to fear. Negative affect systems are also regulated by more general neurochemical
systems including dopaminergic and serotonergic projections arising from the midbrain, and by
circulating gonadal and corticosteroid hormones (Rothbart, Derryberry, et al., 1994; Zuckerman,
1995). Neurochemical influences may thus also provide coherence of emotional states within an
individual, and support more general factors of temperament such as negative emotionality.

For example, serotonergic projections from the midbrain raphe nuclei appear to moderate
limbic circuits related to anxiety and aggression (Spoont, 1992). Low serotonergic activity may
thus increase an individual's vulnerability to both fear and frustration, contributing to a general
factor of negative affectivity, including depression, and molecular genetics studies related to
serotonin function are therefore of interest in identifying negative affect as a general diathesis.
Lesch et al. (1996), for example, have reported relations between a polymorphism of the
serotonin transporter gene and measures of anxiety, depression, and neuroticism.

Affiliativeness/Agreeableness

We share with other animals, including mammals, birds, and fish, systems of affiliation
that support pair bonds and care of the young (Insel, 2003). Panksepp (1986b) indicates that
affiliative and prosocial behaviors may depend in part on opiate projections from higher limbic
regions (e.g., amygdala, cingulate cortex) to the ventromedial hypothalamus, with brain opiates
promoting social comfort and bonding, and opiate withdrawal promoting irritability and
aggressiveness. Because ventromedial hypothalamic lesions dramatically increase aggression,
Panksepp (1986b) also suggests that this brain region normally inhibits aggressive behaviors
Temp., Atten., & Dev. Psychopathology 16

controlled by the midbrain's central gray area. Hypothalamic projections can allow for friendly,
trusting, and helpful behaviors between members of a species by suppressing aggressive
tendencies. Mechanisms underlying prosocial and aggressive behaviors would in this way be
reciprocally related, in keeping with the bipolar Agreeableness-Hostility dimension found in Five
Factor Models of personality.

Panksepp (1993) has also reviewed research suggesting links between social bonding and
the hypothalamic neuropeptide oxytocin (OXY), involved in maternal behavior, feelings of
social acceptance and social bonding, and reduction of separation distress. OXY is also released
during sexual activity by both females and males. In Cloninger's (1987) theory, Reward
Dependence is a dimension ranging from being emotionally dependent, warmly sympathetic,
sentimental, persistent, and sensitive to social cues to being socially detached, cool, tough-
minded and independently self-willed. He sees Reward Dependence as related to social
motivation; it is thus similar to dimensions of affiliativeness or agreeableness.

Agreeableness, including at the high end, the prosocial emotions and behaviors and
affiliative tendencies, and at the low end, aggression and manipulativeness has been increasingly
studied in childhood (Graziano, 1994; Graziano & Eisenberg, 1997; Graziano & Tobin, 2002).
Like other originally bipolar dimensions, prosocial and antagonistic dispositions have also been
studied separately (Bohart & Stipek, 2001), and Graziano and Eisenberg (1997) suggest that the
two dispositions may be separable, even though they are negatively related. In a related
argument, Shiner and Caspi (2003) point out that antisocial and prosocial behaviors have
different etiologies (Krueger, Hicks, & McGue, 2001). Any temperamental predisposition to
pro-social behavior needs to be seen as an open system, interacting with social experience for its
outcomes. In research linking parent-reported temperament to personality in early and middle
childhood, two forms of extraversion/surgency have been identified; one is linked to prosocial
behavior and the other to antisocial behavior and aggression (Rothbart & Victor, 2004), again
suggesting the importance of socialization to the development of pro- or anti-social behavior.

Attentional Networks

Functional neuroimaging has allowed many cognitive tasks to be analyzed in terms of the
brain areas they activate, and studies of attention have been among the most often examined
(Corbetta & Shulman, 2002; Driver, Eimer, & Macaluso, in press; Posner & Fan, in press).
Imaging data support the presence of three networks related to different aspects of attention.
These networks carry out the functions of alerting, orienting, and executive control (Posner &
Fan, in press). A summary of the anatomy and transmitters involved in the three networks is
shown in Table 1.

Alerting is defined as achieving and maintaining a state of high sensitivity to incoming


stimuli; orienting is the selection of information from sensory input; executive control involves
the mechanisms for monitoring and resolving conflict among thoughts, feelings, and responses.
Temp., Atten., & Dev. Psychopathology 17

Table 1
Brain Areas and Neuromodulators Involved in Attention Networks

FUNCTION STRUCTURES MODULATOR


Superior parietal
Orient Temporal parietal junction Acetylcholine
Frontal eye fields
Superior colliculus
Locus Coeruleus
Alert Right frontal and parietal Norepinephrine
cortex
Anterior cingulate
Executive Attention Lateral ventral prefrontal Dopamine
Basal ganglia

The alerting system has been associated with frontal and parietal regions of the right hemisphere.
A particularly effective way to vary alertness has been to use warning signals prior to targets.
The influence of warning signals on alertness is thought to be due to modulation of neural
activity by the neurotransmitter norepinepherine (Marrocco & Davidson, 1998).

Orienting involves aligning attention with a source of sensory signals. This may be overt,
as in eye movements, or may occur covertly, without any movement. Orienting to visual events
has been associated with posterior brain areas, including the superior parietal lobe and temporal
parietal junction and the frontal eye fields (Corbetta & Shulman, 2002). Orienting can be
manipulated by presenting a directional or symbolic cue indicating where in space a person
should attend, thereby directing attention to the cued location (Posner, 1980). Event related
functional magnetic resonance imaging (fMRI) studies have suggested that the superior parietal
lobe is associated with orienting following the presentation of a symbolic cue (Corbetta &
Shulman, 2002). The superior parietal lobe in humans is closely related to the lateral
intraparietal area (LIP) in monkeys, which is known to produce eye movements (Anderson,
1989). When a target occurs at an uncued location and attention has to be disengaged and moved
to a new location, there is activity in the temporal parietal junction (Corbetta & Shulman, 2002).
Lesions of the parietal lobe and superior temporal lobe have been consistently related to
difficulties in orienting (Karnath, Ferber, & Himmelbach, 2001).

The second major temperamental control system over reactive approach and action (the
first is reactive fear) is effortful control, supported by development of the executive attention
network. Executive control of attention is often studied by tasks that involve conflict, such as
various versions of the Stroop task. In the Stroop task, subjects must respond to the color of ink
(e.g., red) while ignoring the color word name (e.g., blue) (Bush et al., 2000). Resolving conflict
in the Stroop task activates midline frontal areas (anterior cingulate) and lateral prefrontal cortex
Temp., Atten., & Dev. Psychopathology 18

(Botwinick, Braver, Barch, Carter, & Cohen, 2001; Fan, Flombaum, McCandliss, Thomas, &
Posner, 2003). There is also evidence for the activation of this network in tasks involving
conflict between a central target and surrounding flankers that may be congruent or incongruent
with the target (Botwinick et al., 2001; Fan, McCandliss, Sommer, Raz, & Posner, 2002).
Experimental tasks may provide a means of fractionating the functional contributions of different
areas within the executive attention network (MacDonald, Cohen, Stenger, & Carter, 2000).

Regulatory functions of attention. The anterior cingulate gyrus, one of the main nodes of
the executive attention network, has been linked to a variety of specific functions in attention
(Posner & Fan, in press), working memory (Duncan et al., 2000), emotion (Bush et al., 2000),
pain (Rainville, Duncan, Price, Carrier, & Bushnell, 1997) and monitoring for conflict
(Botwinick et al., 2001) and for error (Holroyd & Coles, 2002). These functions have been well
documented, but no single rubric seems to explain all of them. In emotion studies, the cingulate
is often seen as part of a network involving orbital frontal and limbic (amygdala) structures. The
frontal areas seem to have an ability to interact with the limbic system (Davidson, Putnam, &
Larson, 2000) that could fit well with the idea that attention subserves self-regulation.

A specific test of a self-regulatory role for the cingulate involves exposure to erotic films,
with the requirement to regulate any resulting arousal. The cingulate activity shown by fMRI
was found to increase in relation to the regulatory instruction (Beauregard, Levesque, &
Bourgoulin, 2001). In a different study, cognitive reappraisal of photographs producing negative
affect showed a positive correlation between the extent of cingulate activity and the amount of
affect to be controlled (Ochsner, Bunge, Gross, & Gabrieli, 2002). Similarly, in a study in which
hypnotism was used to control the perception of pain, cingulate activity reflected the degree of
pain as rated by the subject under hypnotism rather than the physical intensity of the heat
stimulus (Rainville et al., 1997). These results show a role for the anterior cingulate in
regulating limbic activity related to emotion, and provide evidence for its role as a part of a
network involved in regulation, cognition, and affect (Bush et al., 2000).

In experimental paradigms like the Stroop and flanker tasks, conflict is introduced by the
need to respond to one aspect of the stimulus while ignoring another (Bush et al., 2000; Fan et
al., 2003). Cognitive activity that involves this kind of conflict activates the dorsal anterior
cingulate and lateral prefrontal cortex. Large lesions of the anterior cingulate either in adults
(Damasio, 1994) or children (Anderson, Damasio, Tranel, & Damasio, 2000) result in great
difficulty in regulating behavior, particularly in social situations. Smaller lesions may produce
only a temporary inability to deal with conflict in cognitive tasks (Turken & Swick, 1999;
Ochsner et al., 2002).

A major advantage of viewing attention in relation to self-regulation is that knowledge of


the development of a specific neural network can be linked to the ability of children and adults to
regulate their thoughts and feelings. Over the early years of life, regulation of emotion and
behavior is a major issue of development, and issues of dysregulation continue to play a role in
psychopathology in adults.
Temp., Atten., & Dev. Psychopathology 19

TEMPERAMENT AND ADAPTATION

Early Development

Developmental studies of non-disordered children are also consistent with models such as
Gray and McNaughtons (1996), where an anxiety-related behavioral inhibition system inhibits
an approach-related behavioral activation system. An early appearing inhibitory influence on
temperamental approach is behavioral inhibition or fear control. Fearful inhibition is well
developed by the end of the first year of life, and if a disposition to behavioral inhibition is
strong, positive and negative expressive and approach/avoidance behavior will be moderated
under novel or challenging situations. In non-disordered children, measures of infants'
fearfulness in the laboratory negatively predict mothers' reports of children's impulsivity,
activity, and aggression at age 7. Fearfulness in infancy also positively predicted susceptibility
to guilt and shame, two powerful socializing emotions (Rothbart, Derryberry, et al., 2000;
Rothbart, Ahadi, et al., 1994).

Together with the earlier finding relating seven-year Surgency to aggression (Rothbart,
Derryberry, et al., 1994), this suggests that strong approach tendencies may contribute to
negative as well as positive emotionality (Derryberry & Reed, 1994; Rothbart, Derryberry, et al.,
1994). A more complex pattern emerges with the infants approach behavior. Children who
showed short-latency grasps of objects at 6.5, 10, and 13.5 months showed high levels of
positive anticipation and impulsivity, along with high anger-frustration and aggression at seven
years. This again suggests that strong approach tendencies contribute to later negative as well as
positive emotionality.

In addition, children showing rapid approach as infants tended to be low in attentional


control and inhibitory control at age seven. This is consistent with our finding of a negative
relation between Surgency and Effortful Control factors (Rothbart, Derryberry, et al., 1994), and
suggests that strong approach tendencies may constrain the development of voluntary self
control. If approach tendencies are viewed as the accelerator toward action and inhibitory
tendencies as the brakes, it is not surprising that stronger accelerative tendencies may weaken
the braking influence of inhibitory control (Rothbart & Derryberry, 2002).

In our laboratory study, frustration at 6 and 10 months predicted seven-year


anger/frustration but not fear, as well as other components of negative affectivity, including high
discomfort, high guilt/shame, and low soothability (Rothbart, Derryberry, et al., 2000). Greater
infant frustration in the laboratory was also related to higher seven-year activity level, positive
anticipation, impulsivity, aggression, and high intensity pleasure. While infant fear is thus
related to relatively weak approach behavior and to internalizing tendencies in childhood, infant
frustration is related to stronger approach and to externalizing as well as internalizing tendencies.
This finding is consistent with Panksepp's (1998) suggestion that unsuccessful reward-related
activities may activate the anger/frustration functions of a Rage system. Strong approach
tendencies may include positive expectations and frustrated reactions under conditions when
those expectations are not met.
Temp., Atten., & Dev. Psychopathology 20

Fear appears to take on an important inhibitory role in early development, constraining


approach, and aggression, acting as a protective factor for the development of externalizing, and
contributing to the development of conscience. In addition to Kochanskas findings on fear and
the development of conscience (reviewed by Rothbart & Bates, in press), children with
concurrent ADHD and anxiety also show reduced impulsivity relative to children with ADHD
alone (Pliszka, 1989), and aggressiveness appears to decrease between kindergarten and first
grade in children who show internalizing patterns of behavior (Bates et al., 1995).

From an evolutionary point of view, fearful inhibition can serve to protect the individual
from approaching potentially harmful objects or situations. Nevertheless, the fearful form of
inhibitory control remains a relatively reactive process that can be easily elicited by situational
cues. In the course of development, this system can lead to rigid and over-controlled patterns of
behavior that can limit the individual's positive experiences with the world (Block & Block,
1980; Kremen & Block, 1998).

Later Development

Shiner and her colleagues have recently been studying the continuity of personality from
the period 8-12 years to 20 years (Shiner, Masten, & Tellegen, 2002) and 30 years (Shiner,
Masten, & Roberts, 2003). The middle childhood variables, based on parent and child interviews
and teacher questionnaires, included mastery motivation, academic conscientiousness, surgent
engagement, agreeableness, and self-assurance versus anxious insecurity. Adult measures
employed self- and parent-report questionnaires, including the Multidimensional Personality
Questionnaire (MPQ; Tellegen, 1985), as well as data on academic achievement, rule
compliance, social competence, job competence, and romantic competence.

Tellegens self-report MPQ contains three broad personality factors. Positive


Emotionality (PEM) includes scales for well-being, achievement, social potency, and social
closeness. Negative Emotionality (NEM) includes scales for stress reaction, alienation, and
aggression. Constraint (CON) includes scales for control, harm avoidance, and traditionalism.
PEM was moderately predicted at age 20 by mastery motivation, surgent engagement, and self-
assurance in middle childhood (Shiner et al., 2002). PEM was also related to concurrent social
and romantic competencies at 20 years, but not to any of the childhood measures of adaptation.
Negative Emotionality (NEM) at 20, however, was related to childhood low adaptation in all
areas, and to all concurrent adaptation measures except romantic competence. Even controlling
for childhood personality, lower academic achievement and greater conduct problems in
childhood continued to predict adult NEM. Childhood mastery motivation and surgent
engagement were inversely related to NEM in adulthood. At age 20, Constraint (CON) was
predicted by earlier lower self-assurance and higher academic competence, but when childhood
personality was controlled, it was not related to childhood adaptation.

Shiner, Masten, and Tellegen (2002) suggest on the basis of longitudinal findings that
positive emotionality might be more closely linked to current adaptation, whereas negative
emotionality shows more continuity with earlier adaptation. Below, we note strong links
between negative emotionality and psychopathology, both in childhood and adulthood. Negative
emotionality is also particularly linked to behavior problems when effortful control is low
Temp., Atten., & Dev. Psychopathology 21

(Eisenberg, Fabes, Guthrie, & Reiser, 2000). How might this occur? First, mental processing
related to distress may have involved attempts to decrease distress, possibly including calling up
repeated memories of the negative event. Positive experiences would in most instances be less
problematic and less intense, and unlikely to involve negative affect unless shame or guilt is also
present, and they are likely to have been less rehearsed. Thus, when mental processing has been
tied to difficult and painful situations in the past, one may face not only current problems, but
also representations in memory that bring forward the linked mental habits experienced in
similar situations in the past. Well-practiced associations make negative affect, cognition, and
action links stronger. Thus early failure, e.g., poor achievement in school, may create the
possibility of long term negative affect that extends to a range of achievement situations later in
life. Shiner et al. (2002) also found that adults high on negative emotionality were particularly
upset by daily problems (Suls, Martin, & David, 1998).

Caspi, Harrington, et al. (2003) linked observations of 1000 children at age 3 to their self-
reported personality at age 26 (96% of the original sample). Undercontrolled children (10% of
the sample) had been temperamentally impulsive, restless, distractible, and negativistic at age 3;
Confident children (28%) were friendly, eager, and somewhat impulsive; Inhibited children (8%)
were fearful, reticent, and easily upset; Reserved children (15%) were timid but not extreme in
shyness; Well-adjusted children (40%) appeared to be capable of self control, adequately self-
confident, and did not become upset during testing. At age 26, previously Undercontrolled
children were higher in negative emotionality, more alienated, and subject to stress reactions.
They also tended to follow a traditional morality. Formerly Inhibited and Reserved children
were high in harm avoidance, low in social potency (less vigorous, dynamic, forceful), and low
in achievement. Both previously Undercontrolled and Confident children were low in harm
avoidance. Confident children were high on social potency; whereas Inhibited children were
high in Constraint and low in Positive Emotionality.

Caspi, Harrington, et al.s (2003) findings provide evidence that the temperament of the
child truly provides a core for the developing aspects of personality. Undercontrolled children,
who combined extraversion/surgency, negative affect, and low attentional control at age 3,
showed neurotic and alienated tendencies as adults. Confident extraverted children were
confident and unfearful as adults. The more shy and fearful Inhibited and Reserved children
maintained their caution and harm avoidance and were low in social potency, whereas the more
extreme Inhibited children were also high in Constraint (a mixture of fearfulness and self-
control) and low in Positive Emotionality and social support. The most interesting aspect of the
results, however, go beyond temperament to touch upon alienation, traditional values, and social
support. Additional research on these topics is needed.

TEMPERAMENT, ATTENTION, AND PSYCHOPATHOLOGY

Methodological and Conceptual Concerns

In this section, we discuss studies that have linked temperament and attention to the
development of behavior problems and other forms of psychopathology. Two concerns must be
taken into account, however, in reviewing this literature. First, although studies have identified
links between temperament and behavior problems or psychopathology, a clear drawback lies in
Temp., Atten., & Dev. Psychopathology 22

determining the direction of influence. Although temperament extremes may predispose one to
behavior problems, an alternative interpretation is that the child's problems associated with the
disorder lead to extremes in the behavior measured as temperament. For example, punishment
from peers and teachers may lead to increased negative affect in the child. Asendorpfs (1989,
1990) model for the development of shyness follows two developmental routes: one is early
appearing and related directly to temperament (social fearfulness); the other is indirect and
operates through peer rejection and punishment. In the latter case, the childs shyness may have
been not present initially but shaped through social experience. In Asendorpf and van Akens
(1994) longitudinal research, only the latter form of shyness was related to later problems in self-
evaluation.

Another important issue raised by a number of investigators (e.g., Barron & Earls, 1984;
Prior, Sanson, Oberklaid, & Northam, 1987; Sanson, Prior, & Kyrios, 1990; Wertlieb, Wiegel,
Springer, & Feldstein, 1987), is the possible confounding of concepts of temperament and
behavior problems, especially when they are assessed by the same rater (usually the mother).
For example, items assessing temperamental dimensions such as activity, approach/withdrawal,
and adaptability may bear considerable similarity to items assessing behavior problems such as
hyperactivity, anxiety, and oppositional behavior. Bates (1990), on the other hand, argued that if
temperament is seen as contributing to the development of behavior problems, then conceptual
overlap across constructs might be theoretically expected.
Lengua, West, and Sandler (1998) addressed possible contamination of measures by
eliminating overlapping items between temperament and symptom measures; associations were
nevertheless found between negative emotionality and depression, and between impulsivity and
conduct problems in a large sample of children of divorce. In further support of Lengua et al.s
(1998) findings, Lemery, Essex, and Smider (2002) asked experts to categorize both Childrens
Behavior Questionnaire items (CBQ; Rothbart et al., 2001) and behavior problem symptoms
(Preschool Behavior Questionnaire; Behar & Stringfield, 1974) as temperament and/or behavior
problem constructs. Factor analysis of longitudinal multi-informant combined data was also
used. With the latter method, 9% of temperament and 23% of symptom items were confounded.

Removing all confounded items did not affect the extensive relation between the two
measures, including mothers reports of temperament predicting fathers and caregivers reports
of behavior problems, and temperament at 3-4 years predicted later DSM-IV symptoms at age 5
(Lemery et al., 2002). Activity level, attentional focusing, and inhibitory control predicted
ADHD better than internalizing or externalizing, but the general finding was that high negative
emotionality and low effortful control predicted childrens problems. Because the CBQ allowed
further differentiation of the negative affectivity construct, anger was found to predict
externalizing, whereas fear and shyness predicted internalizing problems.

Finally, Sheeber (1995; Sheeber & Johnson, 1994) used a treatment outcome study to
address the overlap issue. In her research, parents were trained in a temperament-focused
approach to parenting. In post-treatment assessment, differences were found for the trained
parents in behavior problem ratings, but not temperament ratings. Childrens behavior was
reported as improved on both internalizing and externalizing measures. Although the
temperament-focused content may have influenced this finding, Sheeber (1995) reviews another
Temp., Atten., & Dev. Psychopathology 23

study with similar findings that used a behavior management program (Webster-Stratton &
Eyberg, 1982).

Models and Research

Clark, Watson, and Mineka (1994), follow Tellegen (1985) in describing Negative
Affectivity (NA) or Neuroticism as a temperamental sensitivity to negative stimuli. Those high
in NA:

. . . experience a broad range of negative moods, including not only fear/anxiety and
sadness/depression but also such emotions as guilt, hostility, and self-dissatisfaction
(Watson & Clark, 1984). A wide range of non-mood variables are related to this
affective core, including negative cognitions (Clark, Beck, & Stewart, 1990), somatic
complaints (Watson & Pennebaker, 1989), negativistic appraisals of self and others (Gara
et al., 1993), diverse personality characteristics such as hardiness, pessimism, and low
self-esteem, and various indices of job, marital and life dissatisfaction (Clark et al., 1994,
p. 104).

Watson, Clark, and Harkness (1994) also identify an affective core to Positive Affectivity
or Extraversion, ranging from joyful, enthusiastic, energetic, friendly, bold, assertive, proud,
and confident to dull, flat, disinterested, and unenthusiastic (p. 106). In their review of the
adult literature, including measures of negativity predating the onset of depression, Watson et al.
(1994) concluded that temperamental negative affect is generally related to the mood disorders
of both depression and anxiety, whereas low positive affect appears to be more specifically
related to depression. In this case, two general temperamental dispositions would be seen as
predispositions to depression; one to anxiety.

As noted above, externalizing disorders include aggression, overactivity, delinquency,


ADHD, conduct, and antisocial disorder (Achenbach & Edelbrock, 1978; Cicchetti & Toth,
1991). Using factor scores derived from DSM IV items to assess these problems, scores based
on hyperactivity-impulsivity, conduct disorder, and oppositional defiant disorder are highly
intercorrelated, and all three are highly correlated with attentional problems (Hartman et al.,
2001). Internalizing problems include extreme shyness or inhibition, sensitivity, anxiety and
depression. In Hartman et al.s (2001) study of measures of child psychopathology based on the
DSM-IV, anxiety and depression scores were highly intercorrelated, and both were correlated
with attention problems. The temperamental characteristic most often linked to internalizing
problems has been negative emotionality, but links have also been made between low positive
emotionality and depression, and between the mood disorders and attentional control.
Theoretical and empirical models are increasingly using both temperamental and environmental
variables in predicting outcomes.

Differentiated negative emotionality and behavior disorders. We first consider studies in


which differentiated measures of affect and self regulation (effortful control) were used predict
both internalizing and externalizing behavior problems. In the Bloomington Longitudinal Study
(BLS), infant and toddler ICQ temperamental difficultness (frequent and intense negative affect
and attention demanding) predicted later externalizing and internalizing problems as seen in the
Temp., Atten., & Dev. Psychopathology 24

mother-child relationship, from the preschool to the middle-childhood periods (Bates & Bayles,
1988; Bates, Bayles, Bennett, Ridge, & Brown, 1991; Bates, Maslin, & Frankel, 1985; Lee &
Bates, 1985). Early negative reactivity to novel situations (unadaptability) predicted less
consistently, but when it did, it predicted internalizing problems more than externalizing
problems. Early resistance to control (perhaps akin to the manageability dimension of Hagekull
(1989), and perhaps at least partly related to the construct of effortful control) predicted
externalizing problems more than internalizing problems. This was also found in predicting
externalizing problems at school in both the BLS and in a separate longitudinal study, the Child
Development Project (CDP; Bates, Pettit, Dodge, & Ridge, 1998).

In a structural modeling analysis of CDP data, dealing with the overlap in externalizing
and internalizing symptoms, Keiley, Lofthouse, Bates, Dodge, and Pettit (2003) separated
mother and teacher reports of behavior problems across ages 5 to 14 years into pure
externalizing, pure internalizing, and covarying factors, and then considered early childhood
predictors of each of these factors. Resistant temperament (unmanageability) predicted the pure
factors of mother- and teacher-rated externalizing problems, but not the pure internalizing
factors. Unadaptable temperament (novelty distress) predicted, in a positively direction, both
mother and teacher pure internalizing factors, and to a lesser degree, and negatively, the pure
mother and teacher externalizing factors. That is, fearful temperament predicted higher levels of
internalizing problems and, less strongly, lower levels of externalizing problems. Although a
disposition to fearfulness would not necessarily constrain the possibilities of aggressive and
uncooperative behaviors, it is intuitively reasonable that children who are fearful and sensitive to
potential punishment would be likely to inhibit externalizing behavior (Bates, Pettit, & Dodge,
1995). And finally, difficult temperament (negative emotionality and demandingness) predicted,
in this multivariate context, none of the pure factors, but only the covarying externalizing plus
internalizing factor for mothers.

These predictions are all consistent with models where temperament extremes either
constitute pathology dimensions or predispose to risk for these conditions. The linkages are of
modest size, but they obtain from early in life, and are not eliminated by the inclusion of family
and parenting characteristics in prediction, so they are not simply artifacts of family functioning.
Also supporting the general pattern, Gilliom and Shaw (2004) found, in a sample of preschool-
age boys from low-income families, that high levels of negative emotionality were associated
with initial levels of both externalizing and internalizing problems, whereas high levels of
fearfulness were associated with high initial levels of internalizing problems and with increases
in internalizing problems over time, and with decreases in externalizing problems over time.

In the Dunedin longitudinal study (Caspi & Silva, 1995), ratings based on the child's
behavior during testing sessions, aggregated from ages 3 and 5 years, predicted ratings of parents
and teachers over ages 9 to 11 and in early adolescence (over ages 13 and 15). A temperament
dimension combining lack of control, irritability, and distractibility, predicted for both genders
externalizing problems more strongly than internalizing problems or positive competencies.
Early approach (positive responses to strangers and new test materials) predicted, inversely,
internalizing problems better than externalizing problems for boys. It did not predict either kind
of problem for girls. Early sluggishness (a factor combining a lack of positive affect, passivity,
and wariness/withdrawal from novelty) predicted later internalizing and externalizing problems
Temp., Atten., & Dev. Psychopathology 25

for girls, but not boys, as well as the relative absence of positive competencies for both girls and
boys. The discovery of differentiated patterns in these studies has been found despite the
tendency for externalizing and internalizing adjustment scores to be correlated with each other.
Caspi and Silva (1995) report somewhat similar patterns of linkage between early temperament
and self-reported personality patterns on Tellegen's Multidimensional Personality Questionnaire,
with undercontrolled children tending as adults to be low on harm/avoidance and high on social
alienation, and inhibited children high on harm/avoidance, low on aggression, and low on social
potency.

In a Swedish study, Hagekull (1994) found that parent-reported EASI sociability/shyness


aggregated over mothers' and fathers' questionnaires at 28 and 36 months predicted age 48-
month aggregated parent ratings of internalizing problems. Early impulsivity also predicted later
externalizing but not internalizing problems, and lower ego strength. Early activity predicted
later externalizing and, inversely, later internalizing. Negative emotionality predicted both kinds
of behavior problems as well as lowered ego strength. This pattern of findings is also consonant
with the BLS and Dunedin studies.

Another group of studies considered relations between parent ratings on the Middle
Childhood Temperament Inventory (MCTI; Hegvik, McDevitt, & Carey, 1982) and adjustment.
Wertlieb et al. (1987) found parent ratings of high negative mood were correlated with both
externalizing and internalizing disorders. Scales having to do with manageability (activity, non-
adaptability, and intensity) and self-regulation (non-persistence and non-predictability), on the
other hand, tended to have higher correlations with externalizing than with internalizing. Guerin,
Gottfried, Oliver, and Thomas (1994) found a similar pattern of correlations between parent
temperament reports and teacher adjustment reports at age 10 years, except that negative mood
correlated more strongly with teacher reports of externalizing than internalizing behavior
problems; similar relations were found between age 10 parent temperament ratings and age 11
teacher ratings. Teglasi and MacMahon (1990) also found a similar pattern, with manageability
and self-regulation temperament scales correlating more strongly with externalizing adjustment
scales than with internalizing scales. In addition, their negative reaction to novelty scale was
more closely associated with internalizing than externalizing scales, and mood was moderately to
highly correlated with both kinds of adjustment.

McClowry et al. (1994) found parent rated negative emotional reactivity to be related to
both internalizing and externalizing problems, with low task persistence (self-regulation)
correlated only with externalizing. Their approach scale did not correlate with either
internalizing or externalizing. However, in research by Rothbart, Derryberry, et al. (2000),
infant activity and smiling, related to extraversion, predicted 7-year aggressiveness and
negativism; infant fear predicted lower levels of aggressiveness, and higher levels of fear,
sadness, empathy, guilt, and shame.

Examining the link between temperament and behavior problems in older children (6-
and 9-year-olds), Wertlieb et al. (1987) found high activity, withdrawal, distractibility, and
intensity, and low adaptability, persistence, negative mood, and unpredictability to be
significantly related to higher behavior problem scores. Internalizing behavior scores (reflecting
anxiety, phobias, social withdrawal, etc.) and externalizing behavior scores (reflecting
Temp., Atten., & Dev. Psychopathology 26

hyperactivity, aggression, etc.) were also found to be differentially related to temperament.


Withdrawal was more associated with internalizing scores, and high activity, distractibility, and
low threshold more associated with externalizing scores.

Biederman et al. (1990), in a longitudinal cohort of highly inhibited and highly


uninhibited children, found that inhibited children accounted for the majority of diagnoses of
anxiety disorder (especially phobias) in their sample, and uninhibited children, the majority of
the externalizing disorders (especially oppositional disorder). Despite the extreme sample, some
extremely inhibited children later received attention deficit diagnoses, and some extremely
uninhibited children received anxiety diagnoses, such as overanxious disorder. These results are
similar to the more general pattern based on parent report studies and dimensional adjustment
scores, in which there are some, if typically smaller, cross-domain correlations.

Rothbart, Ahadi, et al. (1994) considered relations between parent reports of temperament
on the Children's Behavior Questionnaire (CBQ) and adjustment on a newly developed
questionnaire on social behavior patterns in a group of 6- and 7-year-olds. A second-order factor
composite index of temperamental negative affectivity predicted the full range of social traits,
including aggressiveness, guilt, help-seeking, and negativity. However, components of the
general negative affect factor were associated with the social traits in a more differentiated way:
fear and sadness were more related to traits such as empathy; anger and discomfort were more
related to aggression and help-seeking.

In summary, these findings indicate that fear-related negative affect is particularly related
to internalizing disorders; anger and irritability to externalizing disorders. In addition, although
self regulation is related to both internalizing and externalizing problems, it appears to be more
strongly related to externalizing disorders. Fearful negative affect also appears to predict later
lower externalizing problems, in keeping with its aspects as a control system. Thus, when
negative affect is decomposed, there appear to be greater possibilities for differentiating the
disposition to mood and behavior problems. Effortful control is also an important negative
predictor of later problems. This control may serve both emotional and behavioral regulation. In
the future, the uses of more highly differentiated measures of temperament are likely to lead to a
better understanding of the development of adjustment and psychopathology.

Generality of Links with Negative Affectivity: Genetics Studies

Behavior genetics studies have also examined the generality of negative affect links to
mood disorders, suggesting a general negative affect predisposition to anxiety and depression.
The genetic correlation between anxiety and depression symptoms in adult twins is in the
vicinity of 1.00, indicating shared genetic influence on these two disorders (Kendler, Heath,
Martin, & Eaves, 1987). Pennington (2002) suggests, given this overlap, that more specific
syndromes may be related to environmental events, with threat related to anxiety and loss to
depression (Eley & Stevenson, 2000).

Considering again the possibility that temperamental extremes may predispose to


psychopathology, we might expect individual differences in fear or behavioral inhibition to be
related to susceptibility to anxiety and a tendency toward social withdrawal. Wolfson, Fields,
Temp., Atten., & Dev. Psychopathology 27

and Rose (1987) found that preschool children with anxiety disorders were reported by their
mothers to be less adaptable, more negative in mood, and more difficult than a nonanxious
control group, and internalizing problems were related to high temperamental withdrawal.
Rosenbaum et al. (1988) studied 2- to 7-year-old children of parents with agoraphobia or panic
disorder. Compared with children of psychiatric controls, the rates of behavioral inhibition were
close to 80% for the children of the agoraphobic or panic disorder parents, and approximately
15% for non-major-depression controls. In a study of adult subjects with major depression, the
co-diagnosis of agoraphobia or panic disorder has been found to be positively related to risk for
psychiatric disorders in their children, especially risk for major depression and separation anxiety
(Weissman, 1989). Weissman concluded that there may be a common underlying process
constituting a vulnerability (diathesis) for both panic disorder and major depression. High
comorbidity of anxiety and depression further suggest that differentiating the two is difficult,
even at the level of clinical diagnosis (Kovacs & Devlin, 1998).

Imaging studies may cast further light on this question. Thomas et al. (2001) carried out
an fMRI study on amygdala functions, comparing a small sample of children with anxiety and
major depression diagnoses in comparison with controls. Differential findings were reported for
anxious and depressed children, with anxious children showing a stronger right amygdala
response to fear faces, and depressed children showed a decreased left amygdala response to all
faces.

Kagan, Snidman, Zentner, and Peterson (1999) have reported that highly reactive infants
at 4 months were likely to be behaviorally inhibited at 4 years, and to exhibit more anxious
symptoms at age 7. Behavioral inhibition in toddlers also predicted general social anxiety at age
13, particularly for girls (Schwartz, Snidman, & Kagan, 1999). Recent genetics contributions to
the fear/phobia literature suggest a continuum of diathesis, from fearfulness to phobias. In a twin
study by Kendler, Myers, Prescott, and Neale (2001), the lifetime history of multiple phobias as
well as irrational fears was assessed. For a sample of males, a common genetic factor was found
for 5 phobia subtypes, as well as genetic factors specific to each subtype, and a common familial
environmental factor. Kendler et al. (2001) also analyzed irrational fears, with findings
consistent with the idea that the irrational fears were a mild manifestation of the same genetic
risk factor that produced clinical phobias. A previous study of women subjects was in agreement
with these findings for men (Kendler, Neal, Kessler, Heath, & Eaves, 1992).

Temperament and Environment Influences

Rubin and Burgess (2001) have reviewed research on links between temperamental
behavioral inhibition, social withdrawal, and maladaption. They distinguish, as did Rubin and
Asendorpf (1993), between children who isolate themselves from the peer group (social
withdrawal) and children who have been rejected by the peer group (social isolation). Their
model posits that inhibited children withdraw from the peer group, thereby failing to develop
social interaction skills, becoming increasingly uncomfortable amongst peers, and finally facing
social isolation from the peer group. This experience in turn, is expected to encourage negative
self-perceptions with respect to social relationships and social competence. Rubin (1993)
identified a group of 11-year-old children above the clinical cutoff on depression inventory
scores. Compared with non-depressed 11-year-olds, the depressed children at age 7 had been
Temp., Atten., & Dev. Psychopathology 28

high on social withdrawal and negative self-perception of social competence. The groups did not
differ on 7-year-old aggression or popularity, suggesting that, contrary to expectation, the
negative self-perception appeared to precede social isolation.

Rubin and Burgess (2001) also suggest that relative right frontal EEG activation in
inhibited children may be related to low left hemisphere-related affective regulation. They note
Eisenberg, Shepard, Fabes, Murphy, and Guthries (1998) findings that early school age
childrens shyness was related to distress, upset and nervousness, and negatively related to
excitement and enthusiasm. Higher levels of negative emotions were also related to avoidant
coping in social problem situations.

Finally, Rubin and Burgess (2001) review research indicating that mothers of anxious
and avoidant children tend to direct, control, and protect their children. They note, however, that
the influence could well be bidirectional, with childrens apparent vulnerability leading to
parental overprotection. It is worthy of note that several studies have now reported that children
whose mothers are more demanding and less protective. For a more complete report of this
work, see Rothbart and Bates (in press).

Bidirectionality can also apply to interpretations of parent treatment of children with


externalizing disorders such as conduct disorder, where parents are found to reject or punish their
children. Studies involving conduct-disordered boys interacting with other mothers as well as
their own find no differences between the two groups of mothers commands and negative
response to the conduct-disordered boys, indicating that the child may be playing an important
role in driving the adults responses (Anderson, Lytton, & Romney, 1986). Given shared
genetic endowment, both parents and children may also share biologically based temperamental
dispositions. In the case of internalizing, for example, if parents also tend to be more fearful,
they may seek to control situations and to protect their children.

Studies have investigated the relationship of temperament to observations of naturalistic


mother-child interaction in preschool age children. Stevenson-Hinde and Simpson (1982) also
found that mothers reported effects of difficult-to-manage behavior on all relationships in the
family, including interactions with siblings and mother-father interactions. In comparison to
mothers of more manageable children, these mothers also reported higher levels of anxiety and
irritability in themselves, many commenting that they did not used to be like this. Rutter and
Quinton (1984) reported that children showing negative mood, low regularity, low malleability,
and fastidiousness were more likely to be the focus of depressed parents' criticism, hostility, and
irritability. In Rutter's (1990) terms, the interactive process reflects a pattern in which the
children's attributes make them a focus for the discord (thus increasing exposure to the risk
variable) and increase the probability that exposure will set in motion a train of adverse
interactions that will prolong the risk (p. 191).

Lee and Bates (1985), observing mothers' interaction with 2-year-olds at home, in
situations where the mothers were attempting to control their toddlers' "trouble" behavior, found
that more distress-prone ("difficult") children approached trouble more often, tended to resist the
mothers' attempts to control them, and had mothers who were more likely both to use more
aversive discipline and to give in to their children. Lee and Bates indicated that the pattern of
Temp., Atten., & Dev. Psychopathology 29

these early interactions resembles the coercive cycle of behavior described in socially aggressive
boys and their mothers (Patterson, 1977, 1980). Even at early ages, children who present their
mothers with more troublesome, resistant behavior may encourage mothers to use aversive
discipline tactics more frequently or to give in, and these in turn may be the antecedents of a
coercive cycle of parent-child interactions related to problems with aggression (Patterson, 1980).

Activation and inhibition thus appear to be important in the development of childrens


conduct problems. Patterson (1980) reported that parents of nonaggressive problem children are
effective in stopping their children's aversive behavior on three out of every four occasions of
punishment. When parents of aggressive problem children employ punishment, however, the
likelihood that the child will persist in the problem behavior increases rather than decreases
(Patterson, 1977, 1980; Snyder, 1977). In our laboratory, we have found (Hershey, 1992) that
more aggressive (by mothers' report) 7-year-old children show less caution, that is, they respond
more quickly after encountering catch trials in an experimental task, than less aggressive 7-year-
olds.

The energizing effect of punishment on extraverts response in the Nichols and Newman
(1986) study is interesting in this respect. Nichols and Newman asked subjects to view a visual
pattern and match a subsequent pattern against it. Rewards and punishments were
preprogrammed and noncontingent on subject response. On trials following punishment,
introverts responded more slowly, but extraverts responded more quickly. Subjects who were
more slowed by punishment performed better on the go-no go task (Patterson, Kosson, &
Newman, 1987). In a study with young children, Saltz, Campbell, and Skotko (1983) found that,
in support of Luria's (1961) observations, increasing the loudness of a no go command
increased 3- to 4-year-old children's likelihood of performing a prohibited act. They also found
that increasing loudness decreased the likelihood of responding for children 5 to 6 years of age,
indicating developmental change in the development of excitatory and inhibitory control and
their balance during the preschool years. Development of the executive attention system
underlying effortful control may be related to these findings.

In addition to their relation to disinhibiting psychopathologies, positive, outgoing


characteristics of the young child have been identified as a general protective factor for the
development of risk factors and psychopathology. More active and outgoing children are likely
to show fewer negative effects of institutionalization, possibly because they are more likely to
receive greater attention from adult caregivers (Schaffer, 1966). Werner's longitudinal studies in
Hawaii indicate that more positive affective and expressive characteristics in the young child
from an at-risk population are associated with fewer psychosocial problems later in life (Werner,
1986; Werner & Smith, 1982). In addition, higher positive affect appears to be a negative
predictor of problems with depression (Mineka, Watson, & Clark, 1998).

Effortful Control and Psychopathology

A second inhibitory influence on impulsivity is the attentionally influenced capacity to


withhold prohibited responses and activate prescribed ones, that is, effortful control. It has been
suggested that this function may be divided into monitoring for conflict and executing inhibitory
control (Botwinick et al., 2001). Monitoring for conflict has been related to anterior cingulate
Temp., Atten., & Dev. Psychopathology 30

function, while lateral prefrontal areas may be more related to executing the inhibitory operation.
In our 7-year-old sample, inhibitory control was predicted by infants' longer latency to approach
toys at 11 and 13 months and by their lower activity level in the laboratory at 13 months
(Rothbart, Derryberry, et al., 2000). At 7 years, inhibitory control was concurrently related to
attentional control and empathy, and negatively to aggression. Inhibitory control was also
moderately positively related to guilt/shame, although uncorrelated with behavioral inhibition as
assessed in shyness.

Several components of effortful control have been identified in the development of


behavior problems. Poor delay of gratification has been identified as a risk factor for aggressive
and delinquent behaviors, and high delay of gratification to adaptive behaviors (Krueger, Caspi,
Moffitt, White, & Stouthamer-Loeber, 1996). In clinic referred children, Huey and Weisz (1997)
found externalizing problems to be related to under-regulation of behavior, and internalizing
problems to high inhibition and low resiliency. Eisenberg et al. (1996) reported correlations
between teacher and parent ratings of low emotional and behavioral self-regulation and acting-
out behavior problems. Eisenberg, Guthrie, et al. (2000) also found a direct prediction of
internalizing problems from low behavioral regulation and high negative emotionality, as well as
moderating effects of negative emotionality on the relation between low attentional regulation
and externalizing problems in grade school children (only children high in emotionality showed
the relation).

Eisenberg et al. (2001) compared the temperament of school age children with
externalizing problems to those with internalizing problems and non-disordered children.
Children with externalizing problems showed higher anger, impulsivity, and lower regulation.
Eisenberg et al. (1998) also found, in non-disordered children, that high internalizing negative
emotion at age 4-6 years predicted shyness at ages 6-8 and 8-10, but chiefly for children low in
the ability to control attention shifting. Lack of emotional, behavioral, and attitudinal control in
the preschool period, assessed via negativism, emotional lability, restlessness, and short attention
span, has predicted externalizing behaviors in adolescents (Caspi, Henry, McGee, Moffitt, &
Silva, 1995). In a multiple regression analysis, lack of control and low approach independently
predicted boys anxiety; lack of control and sluggishness were independent predictors for girls.
Here, both lack of control and response to novelty (behavioral inhibition) predicted childrens
later problems with anxiety. Preschool lack of control has also been used to distinguish between
life-course persistent antisocial behavior and antisocial behavior limited primarily to the period
of adolescence (Moffitt, Caspi, Dickson, Silva, & Stanton, 1996). There is also strong evidence
of links between high levels of control and positive or adaptive outcomes (see review by
Eisenberg, Fabes, Guthrie, & Reiser, 2002).

Merikangas, Swendsen, Preisig, and Chazan (1998) studied adults with psychiatric
disorders and their 7- to 17-year-old offspring. Externalizing or substance use disorders were
associated with low attentional control and high activity across the lifespan. Anxiety and
depression were associated with low adaptability and approach/withdrawal, and comorbidity of
externalizing and externalizing disorders was associated with both clusters of temperament
characteristics and with high clinical severity and impairment.
Temp., Atten., & Dev. Psychopathology 31

In a study of temperament, externalizing problems, and internalizing problems in older


adolescents using the Attention Network Task (Fan et al., 2002), conflict performance on
computerized measures of executive attention was related to mother-reported adolescents' self-
regulation and negative affectivity (Ellis, 2002). In addition, childrens performance was related
to teacher assessment of risk for developing antisocial behaviors. Executive attention scores,
mother- and self-report self-regulation, and parental monitoring all contributed unique variance
to the prediction of problem behavior scores, with individual difference variables accounting for
relatively more variance than parenting variables. Mother- and self-reported self-regulation, and
self-reported affiliativeness and approach tendencies, along with gender, contributed to
prediction of depressive mood scores. Parenting variables did not add significant variance to
predicting depressive mood.

In summary, in additional to lack of inhibitory reactions to punishment (the BIS), lack of


attentional or effortful control has also been related to externalizing problems. We now consider
how temperament may be related to family interaction that, in turn can shape outcomes.

Temperament and Environment

Often, temperament and parenting show additive contributions to behavior problems.


Campbell (1991) reported results from two longitudinal samples. The first included 3-year-olds
referred by parents; the second, a 3-year-old group rated by their teachers as inattentive,
overactive, and impulsive, and a matched control group from the same class. Data were
collected in home visits, laboratory, and preschool classrooms, and children in both samples have
been followed longitudinally. Consistency of ratings of externalizing behavior problems across
parents and teachers of preschool, 1st-grade, and 4th-grade children, and relations were found
between early inattention, hyperactivity, and the development of poor impulse control, and
aggression and noncompliance. In addition, higher levels of externalizing problems in preschool
children were found to be related to both higher levels of stress in the family and the mother's
use of more negative control strategies. In support of this model, Keenan and Shaw (1994) also
found in a low-income sample that a composite of difficultness and resistance to control
predicted laboratory measures of aggression in 18-month old boys, but not girls.

By combining measures of temperament in the first three years, mothers positive


involvement (affection, teaching) and three-year behavior problems, Bates and Bayles (1988)
were also able to predict both externalizing and internalizing behavior problems at six years.
Fisher and Fagot (1992) found that toddler temperament, and parental discipline practices were
independently related to childrens antisocial and coercive behavior at 5-7 years.

Interactive effects. Interactive effects of temperament and parenting are also found.
These are reviewed in Rothbart and Bates (in press), where recent contributions to the literature
make this a particularly exciting area of research. For example, in the 1500-subject Australian
Temperament Study, infant difficult temperament had little direct relation with 7-8 year-old
hostile-aggressive problems, but when this temperament occurred in a setting reflecting poor
mother-child relationships, the level of risk for behavior problems substantially increased
(Swanson et al., 2000). Belsky, Hsieh, and Crnic (1998) also found that parenting of male
Temp., Atten., & Dev. Psychopathology 32

infants was more predictive of externalizing behavior problems at 3 years for infants high in
negativity versus those low in negativity.

Stice and Gonzales (1998) found a moderating effect for temperament; maternal control
and support were more strongly related to antisocial behavior among adolescents who were high
on negative affectivity and behavioral undercontrol. Maziade et al. (1990) reported that 7-year
child difficulty in dysfunctional families with lack of parental rule clarity, consistency, and
consensus, predicted later oppositional and attention deficit disorders at 12 and 16 years. Few
children with difficult temperament and a well-functioning family developed behavior disorders,
and few of the very easy children at 7 showed problems at 12, regardless of family functioning.

Lengua, Wolchik, Sandler, and West (2000) found that both temperament (high positive
emotionality, low negative emotionality, and impulsivity) and parenting (low rejection and
inconsistent parenting) independently predicted successful adjustment after parental divorce.
Interactive effects were found for dysfunction. Rejection was most strongly associated with
behavior problems in children low in positive affect and inconsistency with problems for highly
impulsive children. Belsky et al. (1998) found that parenting of male infants was more
predictive of externalizing behavior problems at age 3 years for infants high in negativity. Shaw
et al. (1998) regressed out main effects in predicting 24-month externalizing and found a
remaining Sex x Temperament x Parenting interaction. For boys, noncompliant temperament
and mother rejection exerted independent effects on externalizing. For girls, noncompliant
temperament predicted later behavior problems only when combined with mother rejection.

In a setting where negativity was not measured, Bates et al. (1998) found that 13 and 24
month resistance to control interacted with mothers restrictive control at 6, 13 and 24 months, to
predict externalizing in middle childhood. For mothers who were low in control, correlations
were found between earlier temperament and externalizing; when mothers were high in
restrictive control, there was a low correlation between these measures. Bates et al. (1998)
carried out a conceptual replication of this finding in a second study and suggested that
resistance to control may reflect strong approach, weak fear, and/or low effortful control.
Parental control may be more likely to lead to coercive cycles of interaction when the child is
also highly negative, but when control is the main issue, children may benefit from mothers
intervention.

Overall, these findings suggest that there are particular temperamental vulnerabilities that
may make some children more susceptible to negative environments than others. In a recent
study, Caspi et al. (2002) have added an exciting molecular genetics contribution to this
literature, looking specifically at the relation between child maltreatment and the childs
possession of the MAOA gene, which has been related to the development of aggressivity in
monkeys. Maltreatment was assessed between the ages of 3 and 11 years and antisocial
outcomes were predicted for male teenagers and adults. A significant gene by environment
interaction was found, indicating that the effect of maltreatment was stronger in males with low
MAOA activity.

This finding will require replication, but a link has also been reported between the same
gene (MAOA) and the ability of normal subjects to regulate conflict in the Attention Network
Temp., Atten., & Dev. Psychopathology 33

Test (Fossella, Posner, Fan, & Swanson, 2002). When people with two different alleles of the
MAOA gene were studied in fMRI (Fan, Fossella, Sommer, Wu, & Posner, 2003), a difference
was observed in the amount of activity in the anterior cingulate, which is an important node in
the executive attention network (Fan et al., 2002). MAOA is also linked to the regulation of NE,
DA and 5-HT and thus appears to provide protection for antisocial outcomes given stressful early
experience. The authors argue that studying other genotypes associated with stress responses
may promote an understanding of, how stressful experiences are converted into antisocial
behavior toward others in some, but not all, victims of maltreatment (p. 853).

In summary, evidence has increasingly been found for interactions between child
temperament and caregiver treatment in psychopathology outcomes. In addition, studies
continue to find additive contributions to temperament and caregiver treatment in child
outcomes. The coercion model suggests that child characteristics can influence parental
reactions, leading to increased risk of externalizing psychopathology.

Gender Influences

With the exception of activity level (Eaton, 1994), sex differences are rarely found in
early assessments of temperament (Rothbart, 1986), but behavior problems develop
disproportionately in boys. Girls are also at greater risk than boys for depression (Zahn-Waxler,
Cole, & Barrett, 1991), and antisocial behavior is generally found at higher rates for males than
for females (Eagly & Steffen, 1986). Research results currently describe interplay among
temperament, social experience, and gender in the development of depressive mood; two recent
studies suggest that antecedents for depressive mood may be somewhat gender-specific (Block,
Gjerde, & Block, 1991; Patterson & Capaldi, 1990).

Measuring depressive mood at age 18, with self-reported anxiety partialed out, Block et
al. (1991) found that girls' depressive mood was predicted from intropunitive, oversocialized,
and overcontrolling behavior at age 7 and from higher preschool IQ. Boys' depressive mood, on
the other hand, was predicted from higher aggressive, self-aggrandizing, and undercontrolled
behavior at age 7, and from lower preschool IQ. Variables assessing undercontrol for males and
overcontrol for females were also found to be concurrently related to depressive mood at age 18
for this sample (Gjerde, Block, & Block, 1988). Block et al. (1991) noted that their findings
were congruent with findings of a positive relationship between conduct disorder and depression
for male but not for female clinically depressed preadolescents (Edelbrock & Achenbach, 1980;
Puig-Antich, 1982).

Patterson and Capaldi (1990) provided further information on this point, in research with
two samples of 4th-grade boys. Their research results supported a model in which peer rejection
is a central mediating influence on depressed mood. In this model, for at least some boys,
undercontrolled behavior is seen to lead to peer rejection, which in turn is related to development
of depressive mood. For girls, on the other hand, Zahn-Waxler et al. (1991) offered a
developmental model for depression, highlighting vulnerability to depression in highly socialized
girls. Once again, gender may be a moderating variable in the relation between temperament and
the development of depressive mood.
Temp., Atten., & Dev. Psychopathology 34

In a further analysis of gender and psychopathology, Maziade, Cote, Bernier, Boutin, and
Thivierge (1989) suggested that boys difficult temperament is more stable than girls', and that
girls are prone to become less difficult over time even when there is stress in the family. In
addition, Earls and Jung (1987) reported that marital discord and maternal depression are related
to behavior problems in boys but not in girls. They speculated that boys may be under different
socialization pressures than girls, especially when the mother is under stress, although the
direction of effects might also be reversed, with boys' greater externalizing problems creating
higher levels of stress for the mother and for her relationship with the father.

The reader will now have noted the important links between failures of self-regulation
and developmental disorders. In the final section of our review, we consider what we have
learned about attention systems in relation to the development of psychopathology.

ATTENTION AND PSYCHOPATHOLOGY

The ability of attention to regulate distress can be traced to early infancy (Harman,
Rothbart, & Posner, 1997). In infants as young as three months, we have found that orienting to
a visual stimulus provided by the experimenter produces powerful if only temporary soothing of
distress. One of the major accomplishments of the first few years is for infants to develop the
means to achieve this regulation on their own. These events have been reviewed in Rothbart and
Rueda (in press) and Rothbart, Posner, and Kieras (in press).

A sign of the control of cognitive conflict is found late in the first year of life (Diamond,
1991). For example, in A-not-B tasks, children are trained to reach for a hidden object at
location A, and then tested on their ability to search for the hidden object at a new location B.
Children younger than 12 months of age tend to look in the previous location A, even though
they see the object disappear behind location B. After the first year, children develop the ability
to inhibit the prepotent response toward the trained location A, and successfully reach for the
new location B. During this period, infants develop the ability to resolve conflict between line of
sight and line of reach when retrieving an object. At nine months of age, line of sight dominates
completely. If the open side of a box is not in line with the side in view, infants will withdraw
their hand and reach directly along the line of sight, striking the closed side (Diamond, 1991). In
contrast, 12-month-old infants can simultaneously look at a closed side while reaching through
the open end to retrieve a toy.

The ability to use context to reduce conflict can be traced developmentally using the
learning of sequences of locations. Infants as young as 4 months can learn to anticipate the
location of a stimulus, provided the association in the sequence are unambiguous (Haith, Hazan,
& Goodman, 1988). In unambiguous sequences, each location is invariably associated with
another location (e.g., 123) (Clohessy, Posner, & Rothbart, 2001). Because the location of the
current target is fully determined by the preceding item, one type of information needs to be
attended, and there is therefore no conflict (e.g., location 3 always follows location 2). Adults
can learn unambiguous sequences of spatial locations implicitly even when attention is distracted
by a secondary task (Curran & Keele, 1993).
Temp., Atten., & Dev. Psychopathology 35

Ambiguous sequences (e.g., 1213) require attention to the current association in addition
to the context in which the association occurs (e.g., location 1 may be followed by location 2, or
by location 3). Ambiguous sequences pose conflict because for any association, there exists two
strong candidates that can only be disambiguated by context. When distracted, adults are unable
to learn ambiguous sequences of length six (e.g., 123213) (Curran & Keele, 1993), a finding that
demonstrates the need for higher-level attentional resources to resolve this conflict. In young
children, even simple ambiguous associations (e.g., 1213) were not performed at above chance
level until about two years of age (Clohessy et al., 2001).

Developmental changes in executive attention were found during the third year of life
using a conflict key-pressing task (Gerardi-Caulton, 2000). Because children of this age do not
read, location and identity, rather than meaning and ink color, served as the dimensions of
conflict (spatial conflict). Children sat in front of two response keys, one located to the childs
left and one to the right. Each key displayed a picture, and on every trial a picture identical to
one member of the pair appeared on either the left or right side of the screen. Children were
rewarded for responding to the identity of the stimulus regardless of its spatial compatibility with
the matching response key (Gerardi-Caulton, 2000). Reduced accuracy and slowed reaction
times for spatially incompatible relative to spatially compatible trials reflect the effort required to
resist the prepotent response and resolve conflict between these two competing dimensions.
Performance on this task produces a clear interference effect in adults and activates the anterior
cingulate (Fan, Flombaum, et al., 2003). Children 24 months of age tended to perseverate on a
single response, while 36-month-old children performed at high accuracy levels, but like adults
responded more slowly and with reduced accuracy to incompatible trials.

At 30 months, when toddlers were first able to successfully perform the spatial conflict
task, we found that performance on this task was significantly correlated with their ability to
learn the ambiguous associations in the sequence learning task described above (Rothbart, Ellis,
Rueda, & Posner, in press). This finding, together with the failure of 4-month-olds to learn
ambiguous sequences, holds out the promise of being able to trace the emergence of executive
attention during the first months and years of life.

The importance of being able to study the emergence of executive attention is enhanced
because cognitive measures of conflict resolution in these laboratory tasks have been linked to
aspects of children's temperament. Signs of the development of executive attention by cognitive
tasks relate to a temperament individual differences measure obtained from caregiver reports
called effortful control. Children relatively less affected by spatial conflict received higher
parental ratings of temperamental effortful control and higher scores on laboratory measures of
inhibitory control (Gerardi-Caulton, 2000). We regard effortful control as reflecting the
efficiency with which the executive attention network operates in naturalistic settings.

Assessing Individual Differences in Attentional Networks

The Attention Network Test (ANT) examines the efficiency of the three brain networks
we have discussed above (Fan et al., 2002). The ANT procedure uses differences in reaction
times (RT) between conditions to measure the efficiency of each network. The procedure for the
ANT is illustrated in Figure 1. Subtracting RTs obtained in the double-cue condition from RT in
Temp., Atten., & Dev. Psychopathology 36

Figure 1. The figure illustrates the Attention Network Test (ANT). Each trial begins with one of
the four cues shown in a. The three targets are shown in b. The time course of each trial is
shown in c. At the bottom, are the subtractions that provide the measurement of the efficiency
of each network (After Fan et al., 2002).

the no-cue condition gives a measure of alerting due to a warning signal. Subtracting RTs to
targets at the cued location (spatial cue) from trials using a central cue gives a measure of
orienting. Subtracting RTs for congruent from incongruent target trials provides a measure of
conflict and marks the executive attention network.

The ANT has some useful properties as a measure of attentional efficiency. It does not
use language stimuli, so it can be used with children, speakers of any language, and patients
unable to read or special populations. In about 20 minutes, the test provides a measure of the
efficiency of the alerting, orienting, and conflict networks with reasonable reliability, in addition
to overall RT and error rates. Measuring the three networks scores in the same test also allows
assessment of possible patterns of interactions between them. We do not exclude the possibility
that certain psychopathologies or brain-injured patients would have a dysfunction in the way the
attentional networks interact.
Temp., Atten., & Dev. Psychopathology 37

In a sample of 40 normal persons, network scores were reliable over two successive
presentations. In addition, no correlation was found among the three network scores. An
analysis of RTs in this task showed large main effects for cueing and for the type of target, but
only two small but significant interactions (Fan et al., 2002). They both involved very small
reductions in conflict when either no warning cue is provided or when the cue is at the location
of the target. The latter interaction is to be expected because the effective eccentricity of
flankers is increased when attention is on the central arrow. The first finding appears to arise
because with no warning the subject is generally slow and some conflict resolution may occur
during the longer overall RT. This effect also appears to be responsible for a small but
significant negative correlation between alerting and conflict (r = -.18), found in a larger sample
of more than 200 people who had taken the ANT (Fossella et al., 2002).

Recently, ANT scores have been used as a phenotype to assay the heritability of each of the
attentional functions with a sample of 26 pairs of MZ and DZ twins (Fan, Wu, Fossella, &
Posner, 2001). In accordance with their neuroanatomical and behavioral independence, the three
attentional networks showed different heritability indexes. Despite the small scale of the study,
the executive network scores showed a significant correlation (r = .73) between MZ twins. This
correlation was also significant, although weaker, for the alerting scores (r = .46), while the
orienting network showed no evidence of heritability.

A child version of the ANT task has also been developed that is appropriate for children
from about four years of age. In previous work we found that children work best when there is a
story and when there is clear feedback on their performance (Berger, Jones, Rothbart, & Posner,
2000). In the child version of the ANT, five colorful fish replaced the arrows that typically
appear in the flanker task. Children are invited to help the experimenter feed the middle fish by
pressing a button corresponding to the direction in which it is swimming. Visual feedback (the
middle fish smiles and bubbles come out of its mouth) and auditory feedback (a woohoo
sound) is provided when the response has been successful. In each trial, the flanker fish are
swimming in the same (congruent) or opposite (incongruent) direction as the center fish.

Using the child ANT, the developmental course of the attentional networks has been
studied (Fan, Rueda, et al., 2003; Rueda, Posner, & Rothbart, 2004). As in the adult data, the
child study also revealed independence among the three networks scores. Despite a common
decline in RT, each network also showed a different developmental course. Significant
improvement in conflict resolution was found up until age seven as compared to younger
children, but a remarkable stability in both RT and accuracy conflict scores from the age seven to
adulthood. Alerting scores showed some improvement in late childhood and continued
development between ten year olds and adults. Finally, the orienting score was similar to adult
levels at the age of six. The normal development of these networks may be important in
understanding issues related to developmental pathologies such as those in autism, ADHD and
anxiety and depression.

DEVELOPMENT AND ATTENTIONAL PATHOLOGIES

Attention is a very frequent symptom of many forms of psychopathology. However,


without a real understanding of the neural substrates of attention, this has been a somewhat
Temp., Atten., & Dev. Psychopathology 38

empty classification. This situation has been changed with the systematic application of our
understanding of attentional networks to pathological issues. Viewing attention in terms of
underlying neural networks provides a means of classifying disorders that differs from the usual
internalizing versus externalizing classification that we have applied earlier in the paper. A
number of abnormalities involving attention, including Alzheimers dementia, anxiety, attention
deficit hyperactivity disorder (ADHD), autism, borderline personality disorder, depression and
schizophrenia have been studied either with the Attention Network Task (ANT; Fan et al., 2002),
that measures the efficiency of all three networks (See Figure 1), or parts of the task that measure
one or two of the networks.

Alerting

The alerting network is obtained in the ANT by subtracting the RT following a warning
of when a target will occur, from RTs when participants do not receive any warning signal.
Alerting shows the longest period of development of any of the networks, not reaching adult
levels even by age 10 (Rueda, Fan, et al., 2004). Young participants have larger alerting scores
than adults because they have abnormally long RTs without a warning signal. These effects are
reduced with age. One effect of a warning signal is to improve the alert state, but there may be
other influences as well. For example, children may rehearse the task after the warning signal,
and this may allow them to respond more quickly, and they may be caught somewhat unaware
when a target occurs without warning. Since the ANT task is so simple, adults dont seem to
need to rehearse it once it has been practiced.

It is also known that patients with anterior and posterior right hemisphere damage have
problems when tasks are given without warning (Posner, Inhoff, Friedrich, & Cohen 1987;
Robertson, Mattingley, Rorden, & Driver, 1998). This arises because patients, like young
children, have very long RTs in the absence of warning. Robertson has argued that the inability
to maintain the alert state without a warning signal is a major contributor to the reduced attention
to the side of space opposite the lesion found more strongly in patients with right than with left
posterior lesions.

A study using the ANT has shown that normal elderly persons also show dramatically
larger alerting effects than young adults, resembling what is found with children (Fernandez-
Duque & Black, 2004). The study also found that normally developing older adults and
Alzheimers patients showed the same elevated alerting scores, but only the Alzheimers patients
showed increased difficulty in resolving conflict (see also the discussion of executive attention
below).

ADHD. Studies of ADHD have linked the disorder to aspects of both attention and
sensation seeking, a component of extraversion/surgency (Swanson et al., 2001). Many theories
of ADHD have also suggested a deficit in executive functions (Barkley, 1997). However, in
early work using a spatial orienting task, the most compelling deficit appeared to be a difficulty
in maintaining the alert state in the absence of a warning signal (Swanson et al., 1991). This
difficulty might arise from right hemisphere damage, which has also been reported to be a
feature in many studies of ADHD.
Temp., Atten., & Dev. Psychopathology 39

More recent studies using the ANT have replicated problems with alerting in ADHD,
again mostly due to the inability to maintain the alert state when no warning signal was used
(Beane & Marrocco, 2004; Booth, Carlson, & Tucker, 2001). In one of these studies (Booth et
al., 2001), the ANT was used to discriminate the inattentive subtype from normal children and
the combined subtype. The alerting network best distinguished the different forms of ADHD,
with inattentive children showing less ability to maintain the alert state in the absence of warning
signals. Neuroimaging studies of ADHD children have generally shown right frontal, cingulate,
and basal ganglia deficits (Casey et al., 1997). Since right frontal areas have been related both to
problems with the alert state and with inhibitory control, there could be links between the ANT
results and these imaging studies, but this research has not yet been done.

Casey and colleagues (2001) used three different tasks that rely upon different processes
(stimulus selection, response selection, and response execution) to analyze executive functions in
different types of developmental disorders (Schizophrenia, Sydenham's Chorea, Tourettes
Syndrome, and ADHD). They suggest that each of the cognitive operations is implemented in a
particular basal ganglia-thalamo-cortical circuit. Stimulus processing seems to be related to
connections between basal ganglia, thalamus and the dorso-lateral prefrontal cortex, whereas
response processing is associated with connections of subcortical structures to the lateral orbital
frontal cortex. Schizophrenic patients exhibited deficits in stimulus election, Sydenham's chorea
in response selection, Tourette's syndrome children in response execution, and ADHD children
in both stimulus selection and response execution, suggesting that specific cognitive operations
related to executive attention can be independently damaged.

For a number of years it has been thought that ADHD related to problems with dopamine
and other catecholamine function (Wender, 1971). In molecular genetics research, the dopamine
theory found support in replicated findings that one allele (the 7-repeat) of the dopamine 4
receptor gene is over-represented in children with ADHD and also is associated with a
temperament featuring high sensation seeking (see Swanson et al., 2001, for a recent summary).
Sensation seeking is a lower order construct of extraversion. This finding has been well
replicated for the syndrome of ADHD, but the allele has not been associated with a cognitive
deficit in executive function in a study of children with and without the 7-repeat allele (Swanson
et al., 2000). The 7-repeat was associated with ADHD, but showed it was not with the cognitive
deficit that usually accompanies ADHD (Swanson et al., 2000). This finding led to the
suggestion that there may be two routes to ADHD. One would involve a temperamental extreme
of sensation seeking (or extraversion), the frequency of which might be increased by the
presence of the repeat version of the Dopamine 4 receptor gene. This route need not involve a
cognitive deficit. A second route to ADHD would involve cognitive deficits that might be either
genetic, or due to early brain injury or other experiential factors.

Orienting

Anxiety and depression. There is much new information on the role of attention in
anxiety, depression, and mood disorders. In a series of studies, sub-clinical college students with
high scores on trait anxiety were cued to attend to a location by a positive, negative, or neutral
face. Trials on which the cue to the location was negative and invalid produced longer reaction
times for the trait anxiety subjects (Fox, Ricardo, & Dutton, 2002; Fox, Russo, Bowles, &
Temp., Atten., & Dev. Psychopathology 40

Dutton, 2001). These results replicate and extend previous studies by Derryberry and Reed
(1994), and together suggest that trait anxiety influences the ability to disengage, particularly
from threat stimuli. Derryberry & Reed (1994) suggest that this finding may be dependent upon
relatively low levels of effortful control. Subjects high in effortful control but suffering from
depression did not have difficulty disengaging from negative affect.

A symptom of depression is the tendency to dwell on negative ideation (Beck, 1976).


Vasey and Macleod (2001) reviewed recent studies on trait anxious children, finding evidence of
their children making threatening interpretations of ambiguous stimulus material, overestimating
the likelihood of future negative events, and showing a negative bias toward threatening
information. Negative bias has also been noted in childrens speeded detection of dot probes in
the vicinity of threatening words (Bijttebier, 1998; Schippel, Vasey, Cravens-Brown, & Bretveld,
2003). In temperament research, the ability to control attention was found to be negatively
related to negative affect (Derryberry & Rothbart, 1988; Rothbart, Ahadi, et al., 2000), congruent
with the interpretation that good attentional mechanisms may serve to protect against negative
ideation.

Children who have had a history of abuse have also been found to be hypersensitive to
the recognition of angry faces, misclassifying objectively neutral faces as angry (Pollak &
Kistler, 2002). Together these findings suggest that both temperament and experience can
influence aspects of attention toward threat stimuli, perhaps by both enlarging the boundary of
the classification and increasing the time to dwell on these stimuli. Because orienting to cues
and targets can be studied in children from birth onward (Pollak & Kistler, 2002) it would be
possible to use these findings to obtain more information on the effect of anxiety on normal
development, and to determine the degree to which it predicts later problems, given stressful
experience.

Brain systems related to unipolar depression have been explored in a number of imaging
studies (Drevets & Raichle, 1998; Liotti, Maybert, McGinnis, Brannan, & Jerabeck, 2003).
These studies have generally shown the importance of midline frontal areas, including the
anterior cingulate and orbital prefrontal cortex. In normal persons these areas are related to the
experience and control of emotion (Bush et al., 2000), behavior and cognition (Drevets, 2000),
and self image (Gusnard et al., 2003). The areas appear to function abnormally in depressed
persons exposed to material producing a sad mood, whether or not they were currently
depressed. The overlap between brain activities in currently depressed people with those who
have suffered from depression in the past may explain the frequent occurrence of multiple
depressive bouts.

Autism. Autism is a disorder that has been linked to the orienting system (Akshoomoff,
Pierce, & Courchesne, 2002; Landry & Bryson, in press; Rodier, 2002). It is well known that
autistic persons do not normally orient to faces. However, Landry & Bryson report artistic
difficulty in orienting in tasks that involve non-social stimuli similar to those used in the ANT.
Similar deficits in the ability to disengage and move attention have been reported in autism in
relation to abnormal development of the cerebellum (Akshoomoff et al., 2002). We do not know
if this abnormality is due only to cerebellar deficits, since many of the patients also show parietal
Temp., Atten., & Dev. Psychopathology 41

abnormalities as well. Rodier (2002) has some evidence that the orienting abnormalities found
in autism might relate to a gene associated with migration of cells in early development.

Executive Function

Borderline personality. Borderline personality disorder is characterized by very great


lability of affect and problems in interpersonal relations. In some cases, patients are suicidal or
carry out self-mutilation. Because this diagnosis has been studied largely by psychoanalysts and
has a very complex definition, it might at first be thought of as a poor candidate for a specific
pathophysiology involving attentional networks. However, we focused on the temperamentally
based core symptoms of negative emotionality and difficulty in self-regulation (Posner et al.,
2002). We found that patients were very high in negative affect and relatively low in effortful
control (Rothbart, Ahadi, et al., 2000), and defined a temperamentally matched control group of
normal persons without personality disorder who were equivalent in scores on these two
dimensions. Our study with the ANT found a deficit specific to the executive attention network
in borderline patients (Posner et al., 2002). Preliminary imaging results suggested
overgeneralization of responding in the amygdala, and reduced responding in the anterior
cingulate and related midline frontal areas (Posner et al., 2002). Patients with higher effortful
control and lower conflict scores on the ANT were also the most likely to show the effects of
therapy. This methodology shows the utility of focusing on the core deficits of patients, defining
appropriate control groups based on matched temperament, and using specific attentional tests to
help determine how to conduct imaging studies.

Schizophrenia. A number of years ago, never medicated schizophrenic patients were


tested both by imaging and with a cued detection task similar to the orienting part of the ANT.
At rest, these subjects had shown a focal decrease in cerebral blood flow in the left globus
pallidus (Early, Posner, Reiman, & Raichle, 1989), a part of the basal ganglia with close ties to
the anterior cingulate. These subjects showed a deficit in orienting similar to what had been
found for left parietal patients (Early et al., 1989). When their visual attention was engaged, they
had difficulty in shifting attention to the right visual field. However, they also showed deficits in
conflict tasks, particularly when they had to rely on a language cue. It was concluded that the
overall pattern of their behavior was most consistent with a deficit in the anterior cingulate and
basal ganglia, parts of a frontally based executive attention system.

The deficit in orienting rightward has been replicated in first break schizophrenics, but it
does not seem to be true later in the disorder (Maruff, Currie, Hay, McArthur-Jackson, &
Malone, 1995), nor does the pattern appear to be part of the genetic predisposition for
schizophrenia (Pardo et al., 2000). First break schizophrenic subjects often have been shown to
have left hemisphere deficits, and there have been many reports of anterior cingulate and basal
ganglia deficits in patients with schizophrenia (Benes, 1999). The anterior cingulate may be part
of a much larger network of frontal and temporal structures that operate abnormally in
schizophrenia (Benes, 1999).

A recent study using the ANT casts some light on these results (Wang et al., 2004). In
this case, the schizophrenic patients were chronic and they were compared with a similarly aged
control group. The schizophrenic patients had a much greater difficulty resolving conflict than
Temp., Atten., & Dev. Psychopathology 42

did the normal controls. The deficit in patients was also much larger than that found for
borderline personality patients. There was still a great deal of overlap between the patients and
normal subjects, however, indicating that the deficit is not suitable for making a differential
diagnosis. The data showed a much smaller orienting deficit of the type that had been reported
previously. These findings suggest there is a strong executive deficit in chronic schizophrenia,
as would be anticipated by the Benes (1999) theory. It remains to be determined whether this
deficit exists prior to the initial symptoms, or whether it develops with the disorder.

Chromosome 22q11.2 deletion syndrome. This syndrome is a complex one that involves
a number of abnormalities, including facial and heart structure, but also a mental retardation to
deletion of a number of genes. Children with the deletion are at a high risk for developing of
schizophrenia. Among the genes deleted in this syndrome is the COMT gene, which has been
associated with a performance in a conflict task (Diamond, Briand, Fosella & Gehlbach, 2004)
and with schizophrenia (Egan et al., 2003). In light of these findings, it was to be expected that
the disorder would produce a large executive deficit (Simon et al., in press; Sobin, et al., in
press). Sobin has found in a further paper that the deficit in resolving conflict is associated the
ability to inhibit a blink following a cue that a loud noise would be presented shortly (pre-pulse
inhibition). The authors suggest that the association of the high level attention and pre-pulse
inhibition deficit suggests a pathway that includes both the basal ganglia and the anterior
cingulate.

Summary

Many neurological and psychiatric disorders produce a difficulty in attention. Data to


date suggest that disorders have specific influences on attentional networks. However, a number
of disorders influence the same network (e.g., schizophrenia, borderline and Alzheimers
dementia affect executive attention), and the same disorder can influence more than one network
(e.g., schizophrenia may affect both the executive and orienting network although perhaps at
different stages of the disorder). These findings reduce the utility of attention as a means of
classifying disorders, but they may be useful in understanding their etiology and suggesting
methods of treatment, at least of the attentional symptoms.

SUMMARY AND FUTURE DIRECTIONS

In this chapter, we have described recent findings on the structure and development of
temperament, and related temperament and attention to the development of psychopathology.
We have noted that temperament involves a number of interacting systems, each of which
follows its own course of development. We have identified two major control systems of
temperament, reactive fearfulness, and self-regulatory effortful control, with the latter based on
the development of executive of executive attention that serves to moderate or break the
approach and incentive-related systems, and allow activation of lower probability responses.

We have described a hierarchical structure to temperament: general negative affectivity


includes dispositions towards fear, frustration/anger, sadness, and discomfort, as well as low
soothability. General surgency, or extraversion, includes impulsivity, high intensity pleasure,
activity, and sociability. Effortful control includes inhibitory control, low intensity pleasure,
Temp., Atten., & Dev. Psychopathology 43

attentional shirting, and attentional focusing. These three higher order general dimensions of
temperament are related to 3 of the "Big 5" factors, as well as the Big 3 factors in adult
personality.

As suggested by Cicchetti and Cohen (1995), there are multiple routes to


psychopathology, as well as to positive adaptation, and the influence of a given temperamental
characteristic may vary depending on other temperamental or contextual influences. The nature
of the persons experiences and the choice of settings will also be influenced by affect and
regulated by attention. Although experiences may be chosen on the basis of their affective
consequences, they may also be influenced by the planning processes that are part of
temperamental effortful control.

Temperamental factors may predispose children and adults to psychopathological


outcomes, and contribute to the maintenance, reduction, or intensification of an outcome. In
addition, a given temperamental characteristic may predispose to more than one
psychopathology outcome. Temperament is related to both risk and protective factors in the
development of psychopathology; it may influence risk of recurrence of disorder, and may bias
information processing about the self, leading to further risk or protective factors for the
development of psychopathology. In addition, experience, including the experience of disorder,
will affect the range of stimuli activating the temperamental response and the intensity and
regulation of the reaction.

Temperament involves constitutional differences in reactivity and self-regulation. In this


chapter we were able to make connections between disorders and attentional networks. These
connections are possible due the increasing use of brain imaging studies both of common aspects
of brain networks and of individual differences. Temperamental differences in extraversion,
negative affect, and persistence tasks can provide clues for understanding the genes responsible
for the relatively high heritability of temperament scales. We expect the search for brain
mechanisms underlying temperament and for candidate genes involved in producing these
effects to be a major topic of research in future years.

The ability of temperamental differences to predict disorders has been impressive, but it
is not yet possible to determine the direction of these effects. Increased understanding of the
brain networks and genetic mechanisms related to temperamental dimensions could improve this
picture, and allow us to understand which aspects of temperament predictions of disorders are
causative. A large number of disorders have now been shown to produce different specific
effects on attention networks. Although these effects might not be reliable enough to be used as
differential diagnosis, they do provide clues as to the anatomy impaired by the disorder, and in
the future may guide more specific therapeutic strategies.

When we wrote this chapter for the first edition of the Handbook, we speculated on
possible relations between temperament and psychopathology, and on brain systems related to
reactivity and self-regulation. Evidence for these connections has vastly improved in the eight
intervening years. We expect by the time the handbook is modified again, these trends will be
even more developed, and hope that they will lead to both better understanding and more certain
improvements in treatment.
Temp., Atten., & Dev. Psychopathology 44

REFERENCES

Achenbach, T. M., & Edelbrock, C. S. (1978). The classification of child psychology: A review and
analysis of empirical efforts. Psychological Bulletin, 85, 1275-1301.
Ackerson, L. (1942). Children's behavior problems (Vol. 2). Chicago: University of Chicago Press.
Adolphs, R., & Tranel, D., Hamann, S., Young, A. W., Calder, A. J., Phelps, E. A., Anderson, A., Lee, G.
P., & Damasio, A. R. (1999). Recognition of facial emotion in nine individuals with bilateral
amygdala damage. Neuropsychologia, 37(10), 1111-1117.
Ahadi, S. A., & Rothbart, M. K. (1994). Temperament, development and the Big Five. In C. Halverson,
R. Martin, & G. Kohnstamm (Eds.), The developing structure of temperament and personality
from infancy to adulthood (pp. 189-208). Hillsdale, NJ: Erlbaum.
Akshoomoff, N., Pierce, K., & Courchesne, E. (2002). The neurobiological basis of autism from a
developmental perspective. Development and Psychopathology, 14, 613-634.
Anderson, K. E., Lytton, H., & Romney, D. M. (1986). Mothers interactions with normal and conduct-
disordered boys: Who affects whom? Developmental Psychology, 22, 604-609.
Anderson, R.A. (1989). Visual eye movement functions of the posterior parietal cortex. Annual Review
of Neuroscience, 12, 377-403.
Anderson, S. W., Damasio, H., Tranel, D., & Damasio, A. R. (2000). Long-term sequelae of prefrontal
cortex damage acquired in early childhood. Developmental Neuropsychology, 18, 281-296.
Anderson, A. K., & Phelps, E. A. (2002). Is the human amygdala critical for the subjective experience of
emotion?: Evidence of intact dispositional affect in patients with amygdala lesions. Journal of
Cognitive Neuroscience, 14, 709-720.
Asendorpf, J. B. (1989). Shyness as a final common pathway for two different kinds of inhibition.
Journal of Personality and Social Psychology, 53, 481-492.
Asendorpf, J. B. (1990). Development of inhibition during childhood: Evidence for situational specificity
and a two-factor model. Developmental Psychology, 26, 721-730.
Asendorpf, J. B., van Aken, M. A. G. (1994). Traits and relationship status: Stranger versus peer group
inhibition and test intelligence versus peer group competence as early predictors of later self-
esteem. Child Development, 65, 1786-1798.
Barkley, R. A. (1997). Behavioral inhibition, sustained attention, and executive functions: Constructing a
unifying theory of ADHD. Psychological Bulletin, 121, 65-94.
Barron, A. P., & Earls, F. (1984). The relation of temperament and social factors to behavior problems in
three-year-old children. Journal of Child Psychology and Psychiatry, 25, 23-33.
Bates, J. E. (1980). The concept of difficult temperament. Merrill-Palmer Quarterly, 26, 299-319.
Bates, J. E. (1989). Applications of temperament concepts. In G. A. Kohnstamm, J. E. Bates, & M. K.
Rothbart (Eds.), Temperament in childhood (pp. 321-355). New York: Wiley.
Bates, J. E. (1990). Conceptual and empirical linkages between temperament and behavior problems: A
commentary on the Sanson, Prior, and Kyrios study. Merrill-Palmer Quarterly, 36, 193-199.
Bates, J. E., & Bayles, K. (1988). The role of attachment in the development of behavior problems. In J.
Belsky & T. Nezworski (Eds.), Clinical implications of attachment. Child psychology, (pp. 253-
299). Hillsdale, NJ: Erlbaum.
Bates, J. E., Bayles, K., Bennett, D. S., Ridge, B., & Brown, M. M. (1991). Origins of externalizing
behavior problems at eight years of age. In D. Pepler & K. Rubin (Eds.), Development and
treatment of childhood aggression, (pp. 93-120). Hillsdale, NJ: Erlbaum.
Bates, J. E., Freeland, C. A. B., & Lounsbury, M. L. (1979). Measurement of infant difficultness. Child
Development, 50, 794-803.
Temp., Atten., & Dev. Psychopathology 45

Bates, J. E., Maslin, C. A., & Frankel, K. A. (1985). Attachment security, mother-child interaction, and
temperament as predictors of behavior-problem ratings at age three years. In I. Bretherton & E.
Waters (Eds.), Growing points in attachment theory and research. Society for Child Development
Monographs (Vol. Serial No. 209, pp. 167-193).
Bates, J. E., Pettit, G. S., & Dodge, K. A. (1995). Family and child factors in stability and change in
children's aggressiveness in elementary school. In J. McCord (Ed.), Coercion and punishment in
long-term perspectives (pp. 124-138). New York: Cambridge University Press.
Bates, J. E., Pettit, G. S., Dodge, K. A., & Ridge, B. (1998). Interaction of temperamental resistance to
control and restrictive parenting in the development of externalizing behavior. Developmental
Psychology, 34, 982-995.
Beane, M., & Marrocco, R. (2004). Cholinergic and Noradrenergic inputs to the posterior parietal cortex
modulate the components of exogenous attention. In M. I. Posner (Ed.), Cognitive neuroscience
of attention (pp. 313-325). New York: Guilford.
Beauregard, M., Levesque, J., & Bourgouin, P. (2001). Neural correlates of conscious self-regulation of
emotion. Journal of Neuroscience, 21, 6993-7000.
Beck, A. T. (1976). Cognitive therapy and the emotional disorders. Madison, CT: Meridian.
Behar, L., & Stringfield, S. (1974). A behavior rating scale for the preschool child. Developmental
Psychology, 10, 601-610.
Belsky, J., Hsieh, K-H., & Crnic, K. (1998). Mothering, fathering, and infant negativity as antecedents of
boys' externalizing problems and inhibition at age 3 years: Differential susceptibility to rearing
experience? Development & Psychopathology, 10, 301-319.
Benes, F. M. (1999). Model generation and testing to probe neural circuitry in the cingulate cortex of
postmortem schizophrenic brains. Schizophrenia Bulletin, 24, 219-229.
Berger, A., Jones, L., Rothbart, M. K., & Posner, M. I. (2000). Computerized games to study the
development of attention in childhood. Behavioral Research Methods and Instrumentation, 32,
297-303.
Biederman, J., Rosenbaum, J. F., Hirshfeld, D. R., Faraone, S. V., Bolduc, E. A., Gersten, M., Meminger,
S. R., Kagan, J., Snidman, N., & Reznick, S. (1990). Psychiatric correlates of behavioral
inhibition in young children of parents with and without psychiatric disorders. Archives of
General Psychiatry, 47, 21-26.
Bijttebier, P. (1998). Monitoring and blunting coping styles in children. Unpublished doctoral thesis.
Catholic University of Leuven, Belgium.
Blair, R. J. R., Jones, L., Clark, F., & Smith, M. (1997). The psychopath: A lack of responsiveness to
distress cues? Physiology, 34, 192-198.
Blanchard, D. C., & Takahashi, S. N. (1988). No change in intermale aggression after amygdala lesions
which reduce freezing. Physiology and Behavior, 42(6), 613-616.
Block, J. H., & Block, J. (1980). The role of ego-control and ego-resiliency in the organization of
behavior. In W. A. Collins (Ed.), Minnesota symposium on child psychology, (Vol. 13, pp. 39-
101). Hillsdale, NJ: Erlbaum.
Block, J. H., Gjerde, P. F., & Block, J. H. (1991). Personality antecedents of depressive tendencies in 18-
year-olds: A prospective study. Journal of Personality and Social Psychology, 60, 726-738.
Bohart, A. C., & Stipek, D. J. (2001). What have we learned? In A. C. Bohart & D. J. Stipek (Eds.),
Constructive & destructive behavior: Implications for family, school, & society (pp. 367-397).
Washington, DC: American Psychological Association.
Booth, J., Carlson, C. L., & Tucker, D. (2001, June). Cognitive inattention in the ADHD subtypes. Paper
presented at the 10th meeting of the International Society for Research in Child and Adolescent
Psychopathology, Vancouver, Canada.
Botwinick, M. M., Braver, T. S., Barch, D. M., Carter, C. S., & Cohen, J. D. (2001). Conflict monitoring
and cognitive control. Psychological Review, 108, 624-652.
Bronson, W. C. (1972). The role of enduring orientations to the environment in personality development.
Genetic Psychology Monographs, 86, 3-80.
Temp., Atten., & Dev. Psychopathology 46

Burt, C. (1938). The analysis of temperament. British Journal of Medical Psychology, 17, 158-188.
Burton, R. (1921). The anatomy of melancholy. Oxford: Oxford University Press. (Original work
published 1621)
Bush, G., Luu, P., & Posner, M. I. (2000). Cognitive and emotional influences in anterior cingulate
cortex. Trends in Cognitive Sciences, 4, 215-222.
Buss, A. H., & Plomin, R. (1975). A temperament theory of personality development. New York: Wiley.
Buss, A. H., & Plomin, R. (1984). Temperament: Early developing personality traits. Hillsdale, NJ:
Erlbaum.
Buss, K. A., Schumacher, J. R. M., Dolski, I., Kalin, N. H., Goldsmith, H. H., & Davidson, R. J. (2003).
Right frontal brain activity, cortisol, and withdrawal behavior in 6-month-old infants. Behavioral
Neuroscience, 117(1), 11-20.
Calder, A. J., Lawrence, A. D., & Young, A. W. (2001). Neuropsychology of fear and loathing. Nature
Reviews Neuroscience, 2(5), 352-363.
Calkins, S. D., Fox, N. A., & Marshall, T. R. (1996). Behavioral and psychological antecedents of
inhibition in infancy. Child Development, 67, 523-540.
Campbell, S. B. (1991). Longitudinal studies of active and aggressive preschoolers: Individual
differences in early behavior and in outcome. In D. Cicchetti & S. L. Toth (Eds.), Rochester
Symposium on Developmental Psychopathology: Vol. 2. Internalizing and externalizing
expressions of dysfunction, (pp. 57-89). Hillsdale, NJ: Erlbaum.
Canli, T., Zhao, Z., Desmond, J. E., Kang, E. J., Gross, J., & Gabrieli, J. D. E. (2001). An fMRI study of
personality influences on brain reactivity to emotional stimuli. Behavioral Neuroscience, 115,
33-42.
Carey, W. B., & McDevitt, S. C. (1978). Revision of the infant temperament questionnaire. Pediatrics,
61, 735-739.
Carlson, N. R. (1984). Psychology: The science of behavior. Boston: Allyn & Bacon.
Carver, C. S., & White, T. L. (1994). Behavioral inhibition, behavioral activation, and affective
responses to impending reward and punishment: The BIS/BAS scales. Journal of Personality and
Social Psychology, 67, 319-333.
Casey, B. J., Durston, S., & Fossella, J. A. (2001). Evidence for a mechanistic model of cognitive
control. Clinical Neuroscience Research, 1, 267-282.
Casey, B. J., Trainor, R. J., Orendi, J. L., Schubert, A. B., Nystrom, L. E., Giedd, J. N., Castellanos, F. X.,
Haxby, J. V., Noll, D. C., Cohen, J. D., Forman, S. D., Dahl, R. E., & Rapoport, J. L. (1997). A
developmental functional MRI study of prefrontal activation during performance of a go-no-go
task. Journal of Cognitive Neuroscience, 9, 835-846.
Caspi, A., Harrington, H., Milne, B., Amell, J. W., Theodore, R. F., & Moffitt, T. E. (2003). Childrens
behavioral styles at age 3 are linked to their adult personality traits at age 26. Journal of
Personality, 71(4), 495-513.
Caspi, A., Henry, B., McGee, R. O., Moffitt, T. E., & Silva, P. A. (1995). Temperamental origins of child
and adolescent behavior problems: From age three to age fifteen. Child Development, 66, 55-68.
Caspi, A., McClay, J., Moffitt, T., Mill, J., Martin, J., Craig, I. W., Taylor, A., & Poulton, R. (2002).
Role of genotype in the cycle of violence in maltreated children. Science, 297, 851-854.
Caspi, A., & Silva, P. A. (1995). Temperamental qualities at age three predict personality traits in young
adulthood: Longitudinal evidence from a birth cohort. Child Development, 66, 486-498.
Chen, X., Rubin, K. H., & Li, B-S. (1995). Depressed mood in Chinese children: Relations with school
performance and family environment. Journal of Consulting & Clinical Psychology, 63, 938-
947.
Chen, X., Rubin, K. H., & Sun, Y. (1992). Social reputation and peer relationships in Chinese and
Canadian children: A cross-cultural study. Child Development, 63, 1336-1343.
Chess, S., & Thomas, A. (1986). Temperament in clinical practice. New York: Guilford.
Cicchetti, D., & Cohen, D. J. (Eds.). (1995). Developmental psychopathology, Vol. 2: Risk, disorder, and
adaptation. New York: Wiley.
Temp., Atten., & Dev. Psychopathology 47

Cicchetti, D., & Toth, S L. (1991). Rochester Symposium on Developmental Psychopathology: Vol. 2.
Internalizing and externalizing expressions of dysfunction. Hillsdale, NJ: Erlbaum.
Cicchetti, D., & Toth, S. L. (1995). Developmental psychopathology and disorders of affect. In D.
Cicchetti & D. J. Cohen (Eds.), Developmental psychopathology, Vol. 2: Risk, disorder, and
adaptation. Wiley series on personality processes, (pp. 369-420). New York: Wiley.
Cicchetti, D., & Tucker, D. (1994). Development and self-regulatory structures of the mind.
Development and Psychopathology, 6l, 533-549.
Clark, D. A., Beck, A. T., & Stewart, B. (1990). Cognitive specificity and positive-negative affectivity;
Complementary or contradictory view on anxiety and depression? Journal of Abnormal
Psychology, 99, 148-155.
Clark, L. A., Watson, D., & Mineka, S. (1994). Temperament, personality, and the mood and anxiety
disorders. Journal of Abnormal Psychology, 103, 103-116.
Clohessy, A. B., Posner, M. I., & Rothbart, M. K. (2001). Development of the functional visual field.
Acta Psychologica, 106, 51-68.
Cloninger, C. R. (1986). A unified biosocial theory of personality and its role in the development of
anxiety states. Psychiatric Developments, 3, 167-226,
Cloninger, C. R. (1987). A systematic method for clinical description and classification of personality
variants. Archives of General Psychiatry, 44, 573-588.
Corbetta M., & Shulman, G. L. (2002). Control of goal-directed and stimulus-driven attention in the
brain. Nature Reviews Neuroscience, 3, 201-215.
Costa, P. T., Jr., & McCrae, R. R. (1988). From catalog to classification: Murray's needs and the five-
factor model. Journal of Personality and Social Psychology, 55, 258-265.
Culpeper, N. (1657). Galen's art of physik, Translated with a glossary by N. Culpeper. London.
Curran, T., & Keele, S. W. (1993). Attentional and nonattentional forms of sequence learning. Journal
of Experimental Psychology: Learning, Memory, and Cognition, 19, 189-202.
Damasio, A. R. (1994). Descartes error: Emotion, reason and the brain. New York: G. P. Putnam.
Davidson, R. J., Scherer, K. R., Goldsmith, H. H., Pizzagalli, D., Nitschke, J. B., Kalin, N. H., Berridge,
K. C., LaBar, K. S., LeDoux, J. E., Damasio, A. R., Adolphs, R., Damasio, H., McGaugh, J. L.,
Cahill, L., Elliot, R., & Dolan, R. J. (2003). Part I. Neuroscience. In R. J. Davidson, K. R.
Scherer, et al. (Eds.), Handbook of affective sciences. Series in affective science (pp. 3-128).
London: Oxford University Press.
Davidson, R. J., & Fox, N. A. (1989). Frontal brain asymmetry predicts infants' response to maternal
separation. Journal of Abnormal Psychology, 98, 127-131.
Davidson, R. J., Putnam, K. M., & Larson, C. L. (2000). Dysfunction in the neural circuitry of emotion
regulation: A possible prelude to violence. Science, 289, 591-594.
Davidson, R. J., & Tomarken, A. J. (1989). Laterality and emotion: An electrophysiological approach. In
F. Boller & J. Grafman (Eds.), Handbook of neuropsychology (pp. 419-441). Amsterdam, The
Netherlands: Elsevier.
Davis, M., Hitchcock, J. M., & Rosen, J. B. (1987). Anxiety and the amygdala: Pharmacological and
anatomical analysis of the fear-potentiated startle paradigm. In G. Bower (Ed.), The psychology
of learning and motivation: Advances in research and theory (Vol. 21, pp. 263-305). San Diego,
CA: Academic Press.
Depue, R. A., & Collins, P. F. (1999). Neurobiology of the structure of personality: Dopamine,
facilitation of incentive motivation, and extraversion. Behavioral and Brain Sciences, 22, 491-
569.
Depue, R. A., & Iacono, W. G. (1989). Neurobehavioral aspects of affective disorders. In M. R.
Rosenzweig & L. Y. Porter (Eds.), Annual review of psychology, Vol. 40, (pp. 457-492). Palo
Alto, CA: Annual Reviews.
Derryberry, D., & Reed, M. A. (1994). Temperament and the self-organization of personality.
Development and Psychopathology, 6, 653-676.
Temp., Atten., & Dev. Psychopathology 48

Derryberry, D., & Reed, M. A. (1996). Regulatory processes and the development of cognitive
representations. Development and Psychopathology, 8, 215-234.
Derryberry, D., & Rothbart, M. K. (1984). Emotion, attention, and temperament. In C. Izard, J. Kagan,
& R. Zajonc (Eds.), Emotion, cognition, and behavior, (pp. 132-166). Cambridge, UK:
Cambridge University Press.
Derryberry, D., & Rothbart, M. K. (1988). Arousal, affect, and attention as components of temperament.
Journal of Personality and Social Psychology, 55, 958-966.
Derryberry, D., & Tucker, D. M. (1992). Neural mechanisms of emotion. Journal of Consulting &
Clinical Psychology, 60, 329-338.
Diamond, A. (1991). Neuropsychological insights into the meaning of object concept development. In S.
Carey & R. Gelman (Eds.), The epigenesis of mind: Essays on biology and cognition (pp. 67-
110). Hillsdale, NJ: Erlbaum.
Diamond, A., Briand, L., Fossella, J., & Gehlbach, L. (2004). Genetic and neurochemical modulation of
prefrontal cognitive functions in children. American Journal of Psychiatry, 161, 125-132.
Diamond, S. (1974). The roots of psychology. New York: Basic Books.
Digman, J. M., & Inouye, J. (1986). Further specification of the five robust factors of personality.
Journal of Personality and Social Psychology, 50, 116-123.
Drevets, W. C. (2000). Neuroimaging studies of mood disorders. Biological Psychiatry, 18, 813-829.
Drevets, W. C., & Raichle, M. E. (1998). Reciprocal suppression of regional cerebral blood flow during
emotional versus higher cognitive processes: Implications for interactions between emotion and
cognition. Cognition & Emotion, 12, 353-385.
Driver, J., Eimer, M., & Macaluso, E. (in press). Neurobiology of human spatial attention: Modulation,
generation, and integration. In N. Kanwisher & J. Duncan, (Eds.), Attention and Performance
XX: Functional Brain Imaging of Visual Cognition.
Duncan, J., Seitz, R. J., Kolodny, J., Bor, D., Herzog, H., Ahmed, A., Newell, F. N., & Emslie, H. (2000).
A neural basis for general intelligence. Science, 289, 457-460.
Eagly, A. H., & Steffen, V. J. (1986). Gender and aggressive behavior: A meta-analytic review of the
social-psychological literature. Psychological Bulletin, 100, 309-330.
Earls, F., & Jung, K. G. (1987). Temperament and home environment characteristics and causal factors
in the early development of childhood psychopathology. Journal of the American Academy of
Child and Adolescent Psychiatry, 26, 491-498.
Early, T. S., Posner, M.I., Reiman, E. M., & Raichle, M. E. (1989). Left striato-pallidal hyperactivity in
schizophrenia part II. Phenomenology and thought disorder. Psychiatric Developments, 2, 85-
121.
Eaton, W. O. (1994). Temperament, development, and the Five-Factor Model: Lessons from activity
level. In C. F. Halverson, Jr., G. A. Kohnstamm, & R. P. Martin (Eds.), The developing structure
of temperament and personality from infancy to adulthood (pp. 173-187). Hillsdale, NJ: Erlbaum.
Edelbrock, C., & Achenbach, T. M. (1980). A typology of Child Behavior Profile patterns: Distribution
and correlates in disturbed children aged 6 to 16. Journal of Abnormal Child Psychology, 8, 441-
470.
Egan, M. F., Kojima, M., Callicott, J. H., Goldberg, T. E., Kolachana, B. S., Bertolino, A., Zaitsev, E.,
Gold, B., Goldman, D., Dean, M., Lu, B., & Weinberger, D. R. (2003). The BDNF val66met
polymorphism affects activity-dependent secretion of BDNF and H. Publication, vol, pages.
Eisenberg, N., Cumberland, A., Spinrad, T. L., Fabes, R. A., Shepard, S. A., Reiser, M., Murphy, B. C.,
Losoya, S. H., & Guthrie, I. K. (2001). The relations of regulation and emotionality to childrens
externalizing and internalizing problem behavior. Child Development, 72, 1112-1134.
Eisenberg, N., Fabes, R. A., Guthrie, I. K., & Reiser, M. (2000). Dispositional emotionality and
regulation: Their role in predicting quality of social functioning. Journal of Personality & Social
Psychology, 78(1), 136-157.
Temp., Atten., & Dev. Psychopathology 49

Eisenberg, N., Fabes, R. A., Guthrie, I. K., & Reiser, M. (2002). The role of emotionality and regulation
childrens social competence and adjustment. In L. Pulkinnen & A. Caspi (Eds.), Paths to
successful development: Personality in the life course, (pp. 46-70). Cambridge, UK: Cambridge
University Press.
Eisenberg, N., Fabes, R. A., Guthrie, I. K., Murphy, B. C., Maszk, P., Holgren, R., & Suh, K. (1996).
The relations of regulation and emotionality to problem behavior in elementary school children.
Development and Psychopathology, 8, 141-162.
Eisenberg, N., Guthrie, I. K., Fabes, R. A., Shepard, S., Losoya, S., Murphy, B. C., Jones, S., Poulin, R.,
& Reiser, M. (2000). Prediction of elementary school childrens externalizing problem behaviors
from attentional and behavioral regulation and negative emotionality. Child Development, 71,
1367-1382.
Eisenberg, N., Shepard, S. A., Fabes, R. A., Murphy, B. C., & Guthrie, I. K. (1998). Shyness and
childrens emotionality, regulation, and coping: Contemporaneous, longitudinal, and across-
context relations. Child Development, 69, 767-790.
Eley, T. C., & Stevenson, J. (2000). Specific life events and chronic experiences differentially associated
with depression and anxiety in young twins. Journal of Abnormal Child Psychology, 28, 383-
394.
Ellis, L. K. (2002). Individual differences and adolescent psychosocial development. Doctoral
dissertation, University of Oregon, Eugene, OR.
Evans, D., & Rothbart, M. K. (2004). A hierarchal model of temperament and the Big Five. Manuscript
submitted for publication.
Eysenck, H. J. (1944). Types of personality: A factorial study of 700 neurotics. Journal of Mental
Science, 90, 851-861.
Eysenck, H. J. (1947). Dimensions of personality. London: Routledge & Kegan Paul.
Eysenck, H. J. (1970). The structure of human personality (3rd Edition). London: Methuen.
Eysenck, H. J., & Eysenck, M. W. (1985). Personality and individual differences. New York: Plenum.
Fan, J., Flombaum, J. I., McCandliss, B. D., Thomas, K. M., & Posner, M. I. (2003). Cognitive and brain
mechanisms of conflict. NeuroImage, 18, 42-57.
Fan, J., Fossella, J., Sommer, T., Wu, Y., & Posner, M. I. (2003). Mapping the genetic variation of
executive attention onto brain activity. Manuscript submitted for publication.
Fan, J., McCandliss, B. D., Sommer, T., Raz, M., & Posner, M. I. (2002). Testing the efficiency and
independence of attentional networks. Journal of Cognitive Neuroscience, 3, 340-347.
Fan, J., Rueda, M. R., Halparin, J. D., Gruber, D. B., Lercari, L. P., McCandliss, B. D., & Posner, M. I.
(2003). Assaying the development of attentional networks in six- to ten-year-old children.
Manuscript submitted for publication.
Fan, J., Wu, Y., Fossella, J., & Posner, M. I. (2001). Assessing the heritability of attentional networks.
BioMed Central Neuroscience, 2, 14.
Fernandez-Duque, D., & Black, S. (2004). Attentional networks in normal aging and Alzheimers
disease. Manuscript submitted for publication.
Fisher, P. A., & Fagot, B. I. (1992, April). Temperament, parental discipline, and child psychopathology:
A social-interactional model. Paper presented at the Annual Convention of the Western
Psychological Association, Portland, OR.
Fossella, J., Posner, M. I., Fan, J., Swanson, J. M., & Pfaff, D.M. (2002). Attentional phenotypes for the
analysis of higher mental function. The Scientific World Journal, 2, 217-223.
Fox, E., Ricardo, R., & Dutton, K. (2002). Attentional bias for threat: Evidence for delayed
disengagement from emotional faces. Cognition and Emotion, 16, 355-379.
Fox, E., Russo, R., Bowles, R.J., & Dutton, K. (2001). Do threatening stimuli draw or hold attention in
subclinical anxiety? Journal of Experimental Psychology General, 130, 681-700.
Fox, N. A., Calkins, S. D., & Bell, M. A. (1994). Neural plasticity and development in the first two years
of life: Evidence from cognitive and socioemotional domains of research. Development and
Psychopathology, 6, 677-696.
Temp., Atten., & Dev. Psychopathology 50

Fullard, W., McDevitt, S. C., & Carey, W. B. (1984). Assessing temperament in one- to three-year-old
children. Journal of Pediatric Psychology, 9, 205-217.
Gara, M. A., Woolfolk, R. L., Cohen, B. D., Goldston, R. B., Allen, L. A., & Novalany, J. (1993).
Perception of self and other in major depression. Journal of Abnormal Psychology, 102, 93-100.
Gartstein, M. A., & Rothbart, M. K. (2003). Studying infant temperament via the revised Infant Behavior
Questionnaire. Infant Behavior and Development, 26, 64-86.
Gerardi-Caulton, G. (2000). Sensitivity to spatial conflict and the development of self-regulation in
children 24-36 months of age. Developmental Science, 3, 397-404.
Gilliom, M., & Shaw, D. S. (2004). Codevelopment of externalizing and internalizing problems in early
childhood. Development & Psychopathology, 16(2), 313-333.
Gjerde, P. F., Block, J., & Block, J. H. (1988). Depressive symptoms and personality during late
adolescence: Gender differences in the externalization-internalization of symptom expression.
Journal of Abnormal Psychology, 97, 475-486.
Goldberg, L. R. (1990). An alternative "Description of Personality": The Big-Five factor structure.
Journal of Personality and Social Psychology, 59, 1216-1229.
Goldsmith, H. H., & Campos, J. J. (1982). Toward a theory of infant temperament. In R. Emde & R.
Harmon (Eds.), Attachment and affiliative systems (pp. 161-193). New York: Plenum.
Goldsmith, H. H., Rieser-Danner, L. A., & Briggs, S. (1991). Evaluating convergent and discriminant
validity of temperament questionnaires for preschoolers, toddlers, and infants. Developmental
Psychology, 27, 566-579.
Gosling, S. D., & John, O. P. (1999). Personality dimensions in nonhuman animals: A cross-species
review. Current Directions in Psychological Science, 8, 69-75.
Gray, J. A. (1971). The psychology of fear and stress. New York: McGraw-Hill.
Gray, J. A. (1975). Elements of a two-process theory of learning. New York: Academic Press.
Gray, J. A. (1979). A neuropsychological theory of anxiety. In C. E. Izard (Ed.), Emotions in personality
and psychopathology. New York: Oxford University Press.
Gray, J. A. (1982). The neuropsychology of anxiety. New York: Oxford University Press.
Gray, J. A. (1991). The neuropsychology of temperament. In J. Strelau & A. Angleitner (Eds.),
Explorations in temperament: International perspectives on theory and measurement (pp. 105-
128). New York: Plenum.
Gray, J. A. (1994). Framework for a taxonomy of psychiatric disorder. In S. H. M. van Goozen, N. E.
Van de Poll, et al. (Eds.), Emotions: Essays on emotion theory, (pp. 29-59). Hillsdale, NJ:
Erlbaum.
Gray, J. A., & McNaughton, N. (1996). The neuropsychology of anxiety: Reprise. In D. A. Hope (Ed.),
Nebraska Symposium on Motivation: Perspectives on anxiety, panic, and fear. Volume 43, (pp.
61-134). Lincoln, Nebraska: University of Nebraska Press.
Graziano, W. G. (1994). The development of agreeableness as a dimension of personality. In C. F.
Halverson, G. A. Kohnstamm, & R. P. Martin (Eds.), The developing structure of temperament
and personality from infancy to adulthood (pp. 339-354). Hillsdale, NJ: Erlbaum.
Graziano, W. G., & Eisenberg, N. (1997). Agreeableness: A dimension of personality. R. Hogan, J. A.
Johnson, et al. (Eds.), Handbook of personality psychology (pp. 795-824). San Diego, CA:
Academic Press.
Graziano, W. G., & Tobin, R. M. (2002). Agreeableness: Dimension of personality or social desirability
artifact? Journal of Personality, 70(5), 695-727.
Guerin, D. W., Gottfried, A. W., Oliver, P. H., & Thomas, C. W. (1994). Temperament and school
functioning during early adolescence. Journal of Early Adolescence, 14, 200-225.
Gusnard, D. A., Ollinger, J. M., Shulman, G. L., Cloninger, C. R., Price, J. L., Van Essen, D. C., &
Raichle, M. E. (2003). Persistence and brain circuitry. Proceedings of the National Academy of
Sciences of the United States, 100, 3479-3484.
Temp., Atten., & Dev. Psychopathology 51

Hagekull, B. (1989). Longitudinal stability of temperament within a behavioral style framework. In G.


A. Kohnstamm, J. E. Bates, & M. K. Rothbart (Eds.), Temperament in childhood (pp. 283-297).
Chichester, UK: Wiley.
Hagekull, B. (1994). Infant temperament and early childhood functioning: Possible relations to the five-
factor model. In J. C. J. Halverson & G. A. Kohnstamm & R. P. Martin (Eds.), The developing
structure of temperament and personality (pp. 227-240). Hillsdale, NJ: Erlbaum.
Hagekull, B., & Bohlin, G. (1981). Individual stability in dimensions of infant behavior. Infant Behavior
and Development, 4, 97-108.
Haith, M., Hazan, C., & Goodman, G. S. (1988). Expectation and anticipation of dynamic visual events
by 3.5-month-old babies. Child Development, 59, 467-479.
Harman, C., Rothbart, M. K., & Posner, M. I. (1997). Distress and attention interactions in early infancy.
Motivation and Emotion, 21, 27-43.
Harmon-Jones, E., & Allen, J. J. B. (1997). Behavioral activation sensitivity and resting frontal EEG
asymmetry: Covariation of putative indicators related to risk for mood disorders. Journal of
Abnormal Psychology, 106, 159-163.
Hartman, C., Hox, J., Mellenbergh, G. J., Boyle, M. H., Offord, D. R., Racine, Y., McNamee, J., Gadow,
K. D., Sprafkin, J., Kelly, K. L., Nolan, E. E., Tannock, R., Schachar, R., Schut, H., Postma, I.,
Drost, R., & Sergeant, J. A. (2001). DSM-IV internal construct validity: When a taxonomy meets
data. Journal of Child Psychology & Psychiatry & Allied Disciplines, 42, 817-836.
Hegvik, R. L., McDevitt, S. C., & Carey, W. B. (1982). The Middle Childhood Temperament
Questionnaire. Developmental and Behavioral Pediatrics, 3, 197-200.
Hershey, K. L. (1992). Concurrent and longitudinal relationships of temperament to children's
laboratory performance and behavior problems. Doctoral dissertation, University of Oregon,
Eugene.
Holroyd, C. B., & Coles, M. G. H. (2002). The neural basis of human error processing: Reinforcement
learning, dopamine and the error related negativity. Psychological Review, 109, 679-709.
Huey, S. J., Jr., & Weisz, J. R. (1997). Ego control, ego resiliency, and the five-factor model as
predictors of behavioral and emotional problems in clinic-referred children and adolescents.
Journal of Abnormal Psychology, 106,404-415.
Insel, T. R. (2003). Is social attachment an addictive disorder? Physiology & Behavior, 79(3), 351-357.
John, O. P. (1989). Towards a taxonomy of personality descriptors. In D. M. Buss & N. Cantor (Eds.),
Personality psychology: Recent trends and emerging directions (pp. 261-271). New York:
Springer-Verlag.
Jung, C. G. (1923). Psychological types or the psychology of individuation. New York: Harcourt.
Jung, C. G. (1928). Contributions to analytic psychology. New York: Harcourt Brace. Cited in
Fordham, F. (1953). An introduction to Jungs psychology. New York: Plenum.
Kagan, J. (1989). The concept of behavioral inhibition to the unfamiliar. In J. S. Reznick (Ed.),
Perspectives on behavioral inhibition (pp. 1-24). Chicago: University of Chicago Press.
Kagan, J. (1994). Galen's prophecy: Temperament in human nature. New York: Basic Books.
Kagan, J. (1998). Biology and the child. In W. S. E. Damon & N. V. E. Eisenberg (Eds.), Handbook of
child psychology: Vol. 3 Social, emotional and personality development (5th ed., pp. 177-235).
New York: Wiley.
Kagan, J., Snidman, N., Zentner, M., & Peterson, E. (1999). Infant temperament and anxious symptoms
in school age children. Development & Psychopathology, 11, 209-224.
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, 286-292.
Kant, I. (1978). Anthropology from a pragmatic point of view. Carbondale: Southern Illinois University
Press. (Original work published 1798)
Karnath, H-O., Ferber, S., & Himmelbach, M. (2001). Spatial awareness is a function of the temporal not
the posterior parietal lobe. Nature, 411, 951-953.
Temp., Atten., & Dev. Psychopathology 52

Keenan, K., & Shaw, D. S. (1994). The development of aggression in toddlers: A study of low-income
families. Journal of Abnormal Child Psychology, 22, 53-77.
Keiley, M. K., Lofthouse, N., Bates, J. E., Dodge, K. A., & Pettit, G. S. (2003). Differential risks of
covarying and pure components in mother and teacher reports of externalizing and internalizing
behavior across ages 5 to 14. Journal of Abnormal Child Psychology, 31, 264-283.
Kendler, K. S., Heath, A. C., Martin, N., & Eaves, L. J. (1987). Symptoms of anxiety and depression and
generalized anxiety disorder: Same genes, (partly) different environments? Archives of General
Psychiatry, 44, 451-457.
Kendler, K. S., Myers, J., Prescott, C. A., & Neale, M. C. (2001). The genetic epidemiology of irrational
fears and phobias in men. Archives of General Psychiatry, 58, 257-265.
Kendler, K. S., Neale, M. C., Kessler, R. C., Heath, A. C., & Eaves, L. J. (1992). The genetic
epidemiology of phobias in women: The inter-relationship of agoraphobia, social phobia,
situational phobia, and simple phobia. Archives of General Psychiatry, 49, 273-281.
Kovacs, M., & Devlin, B. (1998). Internalizing disorders in childhood. Journal of Child Psychology &
Psychiatry & Allied Disciplines, 39, 47-63.
Kremen, A. M., & Block, J. (1998). The roots of ego-control in young adulthood: Links with parenting in
early childhood. Journal of Personality and Social Psychology, 75, 1062-1075.
Krueger, R. F., Caspi, A., Moffitt, T. E., White, J., & Stouthamer-Loeber, M. (1996). Delay of
gratification, psychopathology, and personality: Is low self-control specific to externalizing
problems? Journal of Personality, 64, 107-129.
Krueger, R. F., Hicks, B. M., & McGue, M. (2001). Altruism and antisocial behavior: Independent
tendencies, unique personality correlates, distinct etiologies. Psychological Science, 12, 397-402.
Landry, R., & Bryson, S. E. (in press). Impaired disengagement of attention in young children with
autism. Journal of Child and Adolescent Psychiatry.
Lawrence, A. D., & Calder, A. J. (in press). Homologizing human emotions. In D. Evans & P. Cruse
(Eds.), Emotion, evolution, and rationality. Oxford, Oxford University Press.
Lawrence, A. D., Calder, A. J., McGowan, S. W., & Grasby, P. M. (2002). Selective disruption of the
recognition of facial expressions of anger. Neuroreport, 13, 881-884.
LeDoux, J. E. (1987). Cognitive-emotional interactions in the brain. Cognition and Emotion, 3(4), 267-
289.
LeDoux, J. E. (1989). Emotion. In F. Plum (Ed.), Handbook of physiology: Vol. V. Higher functions of
the brain. Section 1. The nervous system (pp. 419-460). Bethesda, MD: American Physiological
Society.
Lee, C. L., & Bates, J. E. (1985). Mother-child interaction at age two years and perceived difficult
temperament. Child Development, 56, 1314-1325.
Lemery, K. S., Essex, M. J., & Smider, N. A. (2002). Revealing the relation between temperament and
behavior problem symptoms by eliminating measurement confounding: Expert ratings and factor
analyses. Child Development, 73, 867-882.
Lengua, L. J., West, S. G., & Sandler, I. N. (1998). Temperament as a predictor of symptomatology in
children: Addressing contamination of measures. Child Development, 69, 164-181.
Lengua, L. J., Wolchik, S. A., Sandler, I. N., & West, S. G. (2000). The additive and interactive effects of
parenting and temperament in predicting problems of children of divorce. Journal of Clinical
Child Psychology, 29, 232-244.
Lerner, J. V., Nitz, K., Talwar, R., & Lerner, R. M. (1989). On the functional significance of
temperamental individuality: A developmental contextual view of the concept of goodness of fit.
In G. A. Kohnstamm, J. E. Bates, & M. K. Rothbart (Eds.), Temperament in childhood (pp. 509-
522). Chichester, UK: Wiley.
Lesch, K. P., Bengel, D., Heils, A., Sabol, S. Z., Greenberg, B. D., Petri, S., Benjamin, J., Muller-Reible,
C. R., Hamer, D. H., & Murphy, D. L. (1996). A gene regulatory region polymorphism alters
serotonin transporter expression and is associated with anxiety-related personality traits. Science,
274, 1527-1531.
Temp., Atten., & Dev. Psychopathology 53

Liotti, M., Mayberg, H. S., McGinnis, S., Brannan, S. L., & Jerabeck, P. (2003). Unmasking disease-
specific cerebral flow abnormalities: Mood challenge in patients with remitted unipolar
depression. American Journal of Psychiatry, 159, 183-186.
Lonigan, C. J., & Phillips, B. M. (2001). Temperamental influences on the development of anxiety
disorders. In M. E. Vasey & M. R. Dadds (Eds.), The developmental psychopathology of anxiety,
(pp. 60-91). New York: Oxford University Press.
Luria, A. R. (1961). The role of speech in the regulation of normal and abnormal behavior. New York:
Liveright.
MacDonald, A.W., Cohen, J. D., Stenger, V. A., & Carter, C. S. (2000). Dissociating the role of the
dorsolateral prefrontal and anterior cingulate cortex in cognitive control. Science, 288, 1835-
1838.
Marrocco, R. T., & Davidson, M. C. (1998). Neurochemistry of attention. In R. Parasuraman (Ed.), The
attention brain, (pp. 35-50). Cambridge, MA: MIT Press.
Martin, R., & Presley, R. (1991, April). Dimensions of temperament during the preschool years. Paper
presented at the biennial meeting of the Society for Research in Child Development, Seattle, WA.
Maruff, P., Currie, J., Hay, D., McArthur-Jackson, C., & Malone, V. (1995). Asymmetries in the covert
orienting of visual spatial attention in schizophrenia. Neuropsychologia, 31, 1205-1223.
Maziade, M., Caron, C., Cote, R., Merette, C., Bernier, H., Laplante, B., Boutin, P., & Thivierge, J.
(1990). Psychiatric status of adolescents who had extreme temperaments at age 7. American
Journal of Psychiatry, 147, 1531-1536.
Maziade, M., Cote, R., Bernier, H., Boutin, P., & Thivierge, J. (1989). Significance of extreme
temperament in infancy for clinical status in preschool years. 11: Patterns of temperament and
implications for the appearance of disorders. British Journal of Psychiatry, 154, 544-551.
McClowry, S. G., Giangrande, S. K., Tommasini, N. R., Clinton, W., Foreman, N. S., Lynch, K., &
Ferketich, S. L. (1994). The effects of child temperament, maternal characteristics, and family
circumstances on the maladjustment of school-age children. Research in Nursing & Health, 17,
25-35.
Merenda, P. F. (1987). Toward a four-factor theory of temperament and/or personality. Journal of
Personality Assessment, 51, 367-374.
Merikangas, K. R., Swendsen, J. D., Preisig, M. A., & Chazan, R. Z. (1998). Psychopathology and
temperament in parents and offspring: results of a family study. Journal of Affective Disorders,
51, 63-74.
Mineka, S., Watson, D., & Clark, L. A. (1998). Comorbidity of anxiety and unipolar mood disorders.
Annual Review of Psychology, 49, 377-412.
Moffitt, T. E., Caspi, A., Dickson, N., Silva, P. A., & Stanton, W. (1996). Childhood-onset versus
adolescent-onset antisocial conduct in males: Natural history from age 3 to 18. Development and
Psychopathology, 8, 399-424.
Nichols, S., & Newman, J. P. (1986). Effects of punishment on response latency in extraverts. Journal of
Personality and Social Psychology, 50, 624-630.
Nishijo, H., Ono, T., & Nishino, H. (1988). Single neuron responses in amygdala of alert monkey during
complex sensory stimulation with affective significance. Journal of Neuroscience, 8, 3570-3583.
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, 1215-1229.
Ohman, A. (2000). Fear and anxiety: Evolutionary, Cognitive, and clinical perspectives. In M. Lewis
and J. M. Haviland-Jones (Eds.), Handbook of emotions. Second edition, (pp. 573-593). New
York: Guilford.
Panksepp, J. (1982). Toward a general psychobiological theory of emotions. Behavioral and Brain
Sciences, 5, 407-468.
Panksepp, J. (1986a). The neurochemistry of behavior. Annual Review of Psychology, 37, 77-107.
Temp., Atten., & Dev. Psychopathology 54

Panksepp, J. (1986b). The psychobiology of prosocial behaviors: Separation distress, play, and altruism.
In C. Zahn-Waxler, E. M. Cummings, & R. Iannotti (Eds.), Altruism and aggression: Biological
and social origins (pp. 19-57). Cambridge: Cambridge University Press.
Panksepp, J. (1993). Neurochemical control of moods and emotions: Amino acids to neuropeptides. In
M. Lewis & J. M. Haviland (Eds.), Handbook of emotions (pp. 87-107). New York: Guilford.
Panksepp, J. (1998). Affective neuroscience: The foundations of human and animal emotions. New
York: Oxford University Press.
Pardo, P. J., Knesevich, M. A., Vogler, G. P, Pardo J. V., Towne, B., Cloninger, C. R., & Posner, M. I.
(2000). Genetic and state variables of neurocognitive dysfunction in schizophrenia: A twin study.
Schizophrenia Bulletin, 26, 459-477.
Parker, J. G., & Asher, S. R. (1987). Peer relations and later personal adjustment: Are low-accepted
children at risk? Psychological Bulletin, 102, 357-389.
Patterson, C. M., Kosson, D. S., & Newman, J. P. (1987). Reaction to punishment, reflectivity, and
passive avoidance learning in extraverts. Journal of Personality and Social Psychology, 52, 565-
575.
Patterson, G. R. (1977). Accelerating stimuli for two classes of coercive behaviors. Journal of Abnormal
Child Psychology, 5, 335-350.
Patterson, G. R. (1980). Mothers: The unacknowledged victims. Monographs of the Society for
Research in Child Development, 45(5, Serial No. 186).
Patterson, G. R., & Capaldi, D. M. (1990). A mediational model for boys' depressed mood. In J. Rolf, A.
S. Masten, D. Cicchetti, K. H. Neuchterlin, & S. Weintraub (Eds.), Risk and protective factors in
the development of psychopathology (pp. 141-163). New York: Cambridge University Press.
Pennington, B. F. (2002). The development of psychopathology: Nature and nurture. New York:
Guilford.
Pliszka, S. R. (1989). Effect of anxiety on cognition, behavior, and stimulant response in ADHD.
Journal of the American Academy of Child and Adolescent Psychiatry, 28, 882-887.
Plomin, R. (1982). The concept of temperament: A response to Thomas, Chess, and Korn. Merrill-
Palmer Quarterly, 28, 25-33.
Pollak, S. D., & Kistler, D. J. (2002). Early experience is associated with the development of categorical
representations for facial expressions of emotion. Proceedings of the National Academy of
Sciences of the United States, 99, 9072-9076.
Posner, M. I. (1980). Orienting of attention. The 7th Sir F.C. Bartlett Lecture. Quarterly Journal of
Experimental Psychology, 32, 3-25.
Posner, M. I., & Fan, J. (in press). Attention as an organ system. To J. Pomerantz (Ed.), Neurobiology of
perception and communication: From synapse to society. The Fourth De Lange Conference.
Cambridge, UK: Cambridge University Press.
Posner, M. I., Inhoff, A., Friedrich, F. J., & Cohen, A. (1987). Isolating attentional systems: A cognitive-
anatomical analysis. Psychobiology, 15, 107-121.
Posner, M. I., & Keele, S. W. (1968). On the genesis of abstract ideas. Journal of Experimental
Psychology, 77, 353-363.
Posner, M. I., & Petersen, S. E. (1990). The attention system of the human brain. Annual Review of
Neuroscience, 13, 25-42.
Posner, M. I., & Rothbart, M. K. (1992). Attention and conscious experience. In A. D. Milner & M. D.
Rugg (Eds.), The neuropsychology of consciousness (pp. 91-112). London: Academic Press.
Posner, M. I., & Rothbart, M. K. (1998). Attention, self-regulation, and consciousness. Philosophical
Transactions of the Royal Society of London, B, 353: 1915-1927.
Posner, M. I., & Rothbart, M. K. (2000). Developing mechanisms of self regulation. Development and
Psychopathology, 12, 427-441.
Posner, M. I., Rothbart, M. K., Vizueta, N., Levy, K., Thomas, K. M., & Clarkin, J. (2002). Attentional
mechanisms of borderline personality disorder. Proceedings of the National Academy of
Sciences, 99, 16366-16370.
Temp., Atten., & Dev. Psychopathology 55

Prior, M. R., Sanson, A., Oberklaid, F., & Northam, E. (1987). Measurement of temperament in one- to
three-year-old children. International Journal of Behavioral Development, 10, 131-132.
Puig-Antich, J. (1982). Major depression and conduct disorder in prepuberty. Journal of the American
Academy of Child Psychiatry, 21, 118-128.
Radke-Yarrow, M., & Sherman, T. (1990). Hard growing: Children who survive. In J. E. Rolf, A. S.
Masten, D. Cicchetti, K. H. Neuchterlein, & S. Weintraub (Eds.), Risk and protective factors in
the development of psychopathology (pp. 97-119). New York: Cambridge University Press.
Rainville, P., Duncan, G.H., Price, D.D., Carrier, B., & Bushnell, M.C. (1997). Pain affect encoded in
human anterior cingulate but not somatosensory cortex. Science, 277, 968-970.
Robertson, I. H., Mattingley, J. B., Rorden, C., & Driver, J. (1998). Phasic alerting of neglect patients
overcomes their spatial deficit in visual awareness. Nature, 395, 169-172.
Rodier, P.M. (2002). Converging evidence for brain stem injury during autism. Development and
Psychopathology, 14, 537-559.
Rosch, E. (1973). On the internal structure of perceptual and semantic categories. In T. E. Moore (Ed.),
Cognitive development and the acquisition of language. New York: Academic Press.
Rosenbaum, J. E, Biederman, J., Gersten, M., Hirshfeld, D. R., Meminger, S. R., Herman, J. B., Kagan,
J., Reznick, J. S., & Snidman, N. (1988). Behavioral inhibition in children of parents with panic
disorder and agoraphobia: A controlled study. Archives of General Psychiatry, 45, 463-470.
Rothbart, M. K. (1981). Measurement of temperament in infancy. Child Development, 52, 569-578.
Rothbart, M. K. (1982). The concept of difficult temperament: A critical analysis of Thomas, Chess, &
Korn. Merrill-Palmer Quarterly, 28, 35-40.
Rothbart, M. K. (1986). Longitudinal observation of infant temperament. Developmental Psychology,
22, 356-365.
Rothbart, M. K. (1988). Temperament and the development of inhibited approach. Child Development,
59, 1241-1250.
Rothbart, M. K (1989a). Behavioral approach and inhibition. In S. Reznick (Ed.), Perspectives on
behavioral inhibition (pp. 139-157). Chicago: University of Chicago Press.
Rothbart, M. K. (1989b). Biological processes of temperament. In G. Kohnstamm, J. Bates, & M. K.
Rothbart (Eds.), Temperament in childhood (pp. 77-110). Chichester, UK: Wiley.
Rothbart, M. K (1989c). Temperament in childhood: A framework. In G. Kohnstamm, J. Bates, & M. K.
Rothbart (Eds.), Temperament in childhood (pp. 59-73). Chichester, UK: Wiley.
Rothbart, M. K. (1989d). Temperament and development. In G. Kohnstamm, J. Bates, & M. K. Rothbart
(Eds.), Temperament in childhood (pp. 187-248). Chichester, UK: Wiley.
Rothbart, M. K., & Ahadi, S. A. (1994). Temperament and the development of personality. Journal of
Abnormal Psychology, 103, 55-66.
Rothbart, M. K., Ahadi, S. A., & Evans, D. E. (2000). Temperament and personality: Origins and
outcomes. Journal of Personality and Social Psychology, 78, 122-135.
Rothbart, M. K., Ahadi, S. A., & Hershey, K. L. (1994). Temperament and social behavior in childhood.
Merrill-Palmer Quarterly, 40, 21-39.
Rothbart, M. K., Ahadi, S. A., & Hershey, K. L., & Fisher (2001). Investigations of temperament at three
to seven years: The Children's Behavior Questionnaire. Child Development, 72, 1394-1408.
Rothbart, M. K., & Bates, J. E. (1998). Temperament. In W. Damon (Series Ed.) & N. Eisenberg (Vol.
Ed.), Handbook of child psychology: Vol. 3. Social, emotional and personality development. 5th
ed., (pp. 105-176). New York: Wiley.
Rothbart, M. K., & Bates, J. E. (in press). Temperament in childrens development. In W. Damon, R.
Lerner, & N. Eisenberg (Eds.), Handbook of child psychology: Vol. 3. Social, emotional and
personality development. 6th ed. New York: Wiley.
Rothbart, M. K., & Derryberry, D. (1981). Development of individual differences in temperament. In M.
E. Lamb & A. L. Brown (Eds.), Advances in developmental psychology (Vol. l, pp. 37-86).
Hillsdale, NJ: Erlbaum.
Temp., Atten., & Dev. Psychopathology 56

Rothbart, M. K., & Derryberry, D. (2002). Temperament in children. In C. von Hofsten & L. Bckman
(Eds.), Psychology at the turn of the millennium. Vol. 2: Social, developmental, and clinical
Perspectives (pp. 17-35). East Sussex, UK: Psychology Press.
Rothbart, M. K., Derryberry, D., & Hershey, K. (2000). Stability of temperament in childhood:
Laboratory infant assessment to parent report at seven years. In V. J. Molfese & D. L. Molfese
(Eds.), Temperament and personality development across the life span, (pp. 85-119). Hillsdale,
NJ: Erlbaum.
Rothbart, M. K., Derryberry, D., & Posner, M. I. (1994). A psychobiological approach to the
development of temperament. In J. E. Bates & T. D. Wachs (Eds.), Temperament: Individual
differences at the interface of biology and behavior (pp. 83-116). Washington, DC: American
Psychological Association.
Rothbart, M. K., Ellis, L. K., Rueda, M. R., & Posner, M. I. (in press). Developing mechanisms of
temperamental effortful control. Journal of Personality.
Rothbart, M. K., & Mauro, J. A. (1990). Questionnaire measures of infant temperament. In J. W. Fagen
& J. Colombo (Eds.), Individual differences in infancy: Reliability, stability and prediction (pp.
411-429). Hillsdale, NJ: Erlbaum.
Rothbart, M. K., Posner, M. I., & Hershey, K. (1995). Temperament, attention, and developmental
psychopathology. In D. Cicchetti & J. D. Cohen (Eds.), Manual of developmental
psychopathology (Vol. 1, pp. 315-340). New York: Wiley.
Rothbart, M. K., Posner, M. I., & Kieras, J. (in press). Temperament, attention, and the development of
self-regulation. In K. McCartney, & D. Phillips (Eds.), The Blackwell handbook of early child
development. Blackwell Press.
Rothbart, M. K., & Rueda, M. R. (in press). The development of effortful control. In Developing
individuality in the human brain: A festschrift honoring Michael I. Posner - May, 2003.
Washington, DC: American Psychological Association.
Rothbart, M. K., & Victor, J. B. (2004, October). Temperament and personality development: The CTPQ
(Childrens Temperament and Personality Questionnaire). Paper presented at the annual
Occasional Temperament Conference, Athens, GA.
Rubin, K. H. (1993). The Waterloo Longitudinal Project: Correlates and consequences of social
withdrawal from childhood to adolescence. In H. H. Rubin & J. Adendorph (Eds.), Social
withdrawal, inhibition, and shyness in childhood, (pp. 291-314). Hillsdale, NJ: Lawrence
Erlbaum.
Rubin, K. H., & Asendorpf, J. (Eds.) (1993). Social withdrawal, inhibition, and shyness in childhood.
Hillsdale, NJ: Lawrence Erlbaum.
Rubin, K. H., & Burgess, K. B. (2001). Social withdrawal and anxiety. In M. W. Vasey & M. R. Dadds
(Eds.), The developmental psychopathology of anxiety, (pp. 407-434). New York, Oxford
University Press.
Rueda, M. R., Fan, J., Halparin, J., Gruber, D., Lercari, L. P., McCandliss, B. D., & Posner, M. I. (2004).
Development of attention during childhood. Neuropsychologia, 42, 1029-1040.
Rueda, M. R., Posner, M. I., & Rothbart, M. K. (2004). Attentional control and self regulation. In R. F.
Baumeister & K. D. Vohs (Eds.), Handbook of self-regulation: Research, theory, and
applications (pp. 283-300). New York: Guilford Press.
Rutter, M. (1990). Psychosocial resilience and protective mechanisms. In J. Rolf, A. S. Masten, D.
Cicchetti, K. H. Neuchterlein, & S. Weintraub (Eds.), Risk and protective factors in development
of psychopathology (pp. 181-214). New York: Cambridge University Press.
Rutter, M., & Quinton, D. (1984). Long-term follow-up of women institutionalized in childhood: Factors
promoting good functioning in adult life. British Journal of Developmental Psychology, 18, 225-
234.
Saltz, E., Campbell, S., & Skotko, D. (1983). Verbal control of behavior: The effects of shouting.
Developmental Psychology, 19, 461-464.
Temp., Atten., & Dev. Psychopathology 57

Sanson, A., Prior, M., Garino, E., Oberklaid, F., & Sewell, J. (1987). The structure of infant
temperament: Factor analysis of the revised Infant Temperament Questionnaire. Infant Behavior
and Development, 10, 97-104.
Sanson, A., Prior, M., & Kyrios, M. (1990). Contamination of measures in temperament research.
Merrill-Palmer Quarterly, 36, 179-192.
Schaffer, H. R. (1966). The onset of fear of strangers and the incongruity hypothesis. Journal of Child
Psychology and Psychiatry, 7, 95-106.
Schippel, P., Vasey, M. W., Cravens-Brown, L. M., & Bretveld, R. A. (2003). Suppressed attention to
rejection, ridicule, and failure cues: A unique correlate of reactive but not proactive aggression in
youth. Journal of Clinical Child & Adolescent Psychology, 32, 40-55.
Schwartz, C. E., Snidman N., & Kagan, J. (1999). Adolescent social anxiety as an outcome of inhibited
temperament in childhood. Journal of the American Academy of Child & Adolescent Psychiatry,
38, 1008-1015.
Shaw, D. S., Winslow, E. B., Owens, E. B., Vondra, J. I., Cohn, J. F., Bell, R. Q. (1998). The
development of early externalizing problems among children from low-income families: A
transformational perspective. Journal of Abnormal Child Psychology, 26, 95-107.
Sheeber, L. B. (1995). Empirical dissociations between temperament and behavior problems: A response
to the Sanson, Prior, and Kyrios study. Merrill-Palmer Quarterly, 41, 54-561.
Sheeber, L. B., & Johnson, J. H. (1994). Evaluation of a temperament-focused, parent-training program.
Journal of Clinical Child Psychology, 23, 249-259.
Shiner, R., & Caspi, A. (2003). Personality differences in childhood and adolescence: Measurement,
development, and consequences. Journal of Child Psychology & Psychiatry & Allied Disciplines,
44(1), 2-32.
Shiner, R. L., Masten, A. S., & Roberts, J. M. (2003). Childhood personality foreshadows adult
personality and life outcomes two decades later. Journal of Personality & Social Psychology,
71(6), 1145-1170.
Shiner, R. L., Masten, A. S., & Tellegen, A. (2002). A developmental perspective on personality in
emerging adulthood: Childhood antecedents and concurrent adaptation. Journal of Personality &
Social Psychology, 83(5), 1165-1177.
Simon, T. J., Bish, J. P., Bearden, C. E., Ding, L., Ferrante, S., Nguyen, V., Gee, J., McDonald-McGinn,
D., Zackai, E. H., & Emanuel, B. S. (in press). A multi-level analysis of cognitive dysfunction
and psychopathology associated with chromosome 22q11.2 deletion syndrome in children
development and psychopathology. Developmental Psychopathology.
Smeets, W. J. A. J., & Gonzalez, A. (2000). Catecholamine systems in the brain of vertebrates: New
perspectives through a comparative approach. Brain Research - Brain Research Reviews, 33,
308-379.
Snyder, J. A. (1977). A reinforcement analysis of interaction in problem and nonproblem children.
Journal of Abnormal Psychology, 86, 528-535.
Sobin, C., Kiley-Brabeck, K., Daniels, S., Blundell, M., Anyane-Yeboa, K., & Karayiorgou, M. (in
press). Networks of attention in children with the 22q11 deletion syndrome. Developmental
Neuropsychology.
Spoont, M. R. (1992). Modulatory role of serotonin in neural information processing: Implications for
human psychopathology. Psychological Bulletin, 112, 330-350.
Sprengelmeyer, R., Young, Andrew W., Schroeder, U., Grossenbacher, P. G., Federlein, J., Buettner, T.,
& Przuntek, H. (1999). Knowing no fear. Proceedings of the Royal Society of London - Series B:
Biological Sciences, 266(1437), 2451-2456.
Stevenson-Hinde, J., & Simpson, A. E. (1982). Temperament and relationships. In R. Porter & G. M.
Collins (Eds.), Temperamental differences in infants and young children, (pp. 51-65). London:
Pitman.
Stice, E., & Gonzales, N. (1998). Adolescent temperament moderates the relations of parenting to
antisocial behavior and substance use. Journal of Adolescent Research, 13, 5-31.
Temp., Atten., & Dev. Psychopathology 58

Strelau, J. (1983). Temperament personality activity. New York: Academic Press.


Suls, J., Martin, R., & David, J. P. (1998). Person-environment fit and its limits: Agreeableness,
neuroticism, and emotional reactivity to interpersonal conflict. Personality & Social Psychology
Bulletin, 24(1), 88-98.
Swanson, J., Oosterlaan, J., Murias, M., Moyzis, R., Schuck, S., Mann, M., Feldman, P., Spence, M. A.,
Sergeant, J., Smith, M., Kennedy, J., & Posner, M. I. (2000). ADHD children with 7-repeat allele
of the DRD4 gene have extreme behavior but normal performance on critical neuropsychological
tests of attention. Proceeds of the National Academy of Sciences, 97, 4754-4759.
Swanson, J. M., Posner, M. I., Potkin, S., Bonforte, S., Youpa, D., Cantwell, D., & Crinella, F. (1991).
Activating tasks for the study of visual-spatial attention in ADHD children: A cognitive
anatomical approach. Journal of Child Neurology, 6, S119-S127.
Swanson, J. M., Kraemer, H. C., Hinshaw, S. P., Arnold, L. E., Conners, C. K., Abikoff, H. B.,
Clevenger, W., Davies, M., Elliott, G. R., Greenhill, L. L., Hechtman, L., Hoza, B., Jensen, P. S.,
March, J. S., Newcorn, J. H., Owens, E. B., Pelham, W. E., Schiller, E., Severe, J. B., Simpson,
S., Vitiello, B., Wells, K., Wigal, T., & Wu, M. (2001). Clinical relevance of the primary
findings of the MTA: Success rates based on severity of ADHD and ODD symptoms at the end of
treatment. Journal of the American Academy of Child & Adolescent Psychiatry, 40(2), 168-179.
Teglasi, H., & MacMahon, B. V. (1990). Temperament and common problem behaviors of children.
Journal of Applied Developmental Psychology, 11, 331-349.
Tellegen, A. (1985). Structures of mood and personality and their relevance to assessing anxiety, with an
emphasis on self-report. In A. H. Tuma & J. D. Maser (Eds.), Anxiety and the anxiety disorders
(pp. 681-706). Hillsdale, NJ: Erlbaum.
Thomas, A., & Chess, S. (1977). Temperament and development. New York: Brunner/Mazel.
Thomas, A., Chess, S., & Birch, H. G. (1968). Temperament and behavior disorders in children. New
York: New York University Press.
Thomas, A., Chess, S., Birch, H. G., Hertzig, M. E., & Korn, S. (1963). Behavioral individuality in early
childhood. New York: New York University Press.
Thomas, K. M., Drevets, W. C., Dahl, R. E., Ryan, N. D., Birmaher, B., Eccard, C. H., Axelson, D.,
Whalen, P. J., & Casey, B. J. (2001). Amygdala response to fearful faces in anxious and
depressed children. Archive of General Psychiatry, 58, 1057-1063.
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, 676-687.
Turken, A. U., & Swick, D. (1999). Response selection in the human anterior cingulate cortex. Nature
Neuroscience, 2, 920-924.
Vasey, M. W., & Macleod, C. (2001). Information-processing factors in childhood-anxiety: A review and
developmental perspective. In M. E. Vasey & M. R. Dadds (Eds.), The developmental
psychopathology of anxiety, (pp. 253-277). New York: Oxford University Press.
Wang, K. J., Fan, J., Dong, Y., Wang C., Lee, T. M. C., & Posner, M. I. (2004). Selective impairment of
attentional networks of orienting and executive control in schizophrenia. Manuscript submitted
for publication.
Watson, D., & Clark, L. A. (1984). Negative affectivity: The disposition to experience aversive
emotional states. Psychological Bulletin, 96, 465-490.
Watson, D., & Clark, L. A. (1992). Affects separable and inseparable: A hierarchical model of the
negative affects. Journal of Personality and Social Psychology, 62, 489-505.
Watson, D., & Clark, L. A. (1995). Depression and the melancholic temperament. European Journal of
Personality, 9, 351-366.
Watson, D., & Clark, L. A. (1997). The measurement and mismeasurement of mood: Recurrent and
emergent issues. Journal of Personality Assessment, 68, 267-296.
Watson, D., Clark, L. A., & Harkness, A. R. (1994). Structures of personality and their relevance to
psychopathology. Journal of Abnormal Psychology, 103, 18-31.
Temp., Atten., & Dev. Psychopathology 59

Watson, D., & Pennebaker, J. (1989). Health complaints, stress, and distress: Exploring the central role
of negative affectivity. Psychological Review, 96, 235-254.
Webb, E. (1915). Character and intelligence. British Journal of Psychology Monographs, Nos. 1 & 3,
99.
Webster-Stratton, C. & Eyberg, S. M. (1982). Child temperament: Relationship with child behavior
problems and parent-child interactions. Journal of Clinical Child Psychology, 11, 123-129.
Weissman, M. M. (1989). Anxiety disorders in parents and children: A genetic-epidemiological
perspective. In J. S. Reznick (Ed.), Perspectives on behavioral inhibition (pp. 241-254).
Chicago: University of Chicago Press.
Wender, P. (1971). Minimal brain dysfunction in children. New York: Wiley-Liss.
Werner, E. E. (1986). The concept of risk from a developmental perspective. In B. Keogh (Ed.),
Advances in special education: Developmental problems in infancy and the preschool years.
Greenwich, CT: JAI Press.
Werner, E. E., & Smith, R. S. (1982). Vulnerable, but invincible: A longitudinal study of resilient
children and youth. New York: McGraw-Hill.
Wertlieb, D., Wiegel, C., Springer, T., & Feldstein, M. (1987). Temperament as a moderator of children's
stressful experiences. American Journal of Orthopsychiatry, 57, 234-245.
Wolfson, J., Fields, J. H., & Rose, S. (1987). Symptoms, temperament, resiliency, and control in anxiety-
disordered preschool children. Journal of the American Academy of Child and Adolescent
Psychiatry, 26, 16-22.
Zahn-Waxler, C., Cole, P. M., & Barrett, K. C. (1991). Guilt and empathy: Sex differences and
implications for the development of depression. In J. Garber & K. A. Dodge (Eds.), The
development of emotion regulation and dysregulation (pp. 243-272). New York: Cambridge
University Press.
Zuckerman, M. (1984). Sensation seeking: A comparative approach to a human trait. Behavioral and
Brain Sciences, 7, 413-471.
Zuckerman, M. (1995). Good and bad humors: Biochemical bases of personality and its disorders.
Psychological Science, 6(6), 325-332.

View publication stats

Das könnte Ihnen auch gefallen