Anda di halaman 1dari 9

Neuropsychologia 44 (2006) 2149–2157

What have we learned about cognitive development from neuroimaging?


Sarah Durston a,b,∗ , B.J. Casey a
a Sackler Institute for Developmental Psychobiology, Weill Medical College of Cornell University,
1300 York Avenue, Box 140, New York, NY 10021, USA
b Rudolf Magnus Institute of Neuroscience, Department of Child and Adolescent Psychiatry,

University Medical Center Utrecht, Heidelberglaan 100, 3584 CX, Utrecht, The Netherlands
Received 6 June 2005; received in revised form 22 September 2005; accepted 15 October 2005
Available online 21 November 2005

Abstract
Changes in many domains of cognition occur with development. In this paper, we discuss neuroimaging approaches to understanding these
changes at a neural level. We highlight how modern imaging methods such as functional magnetic resonance imaging (fMRI) and diffusion
tensor imaging (DTI) are being used to examine how cognitive development is supported by the maturation of the brain. Some reports suggest
developmental changes in patterns of brain activity appear to involve a shift from diffuse to more focal activation, likely representing a fine-tuning
of relevant neural systems with experience. One of the challenges in investigating the interplay between cognitive development and maturation
of the brain is to separate the contributions of neural changes specific to development and learning. Examples are given from the developmental
neuroimaging literature. The focus is on the development of cognitive control, as the protracted developmental course of this ability into adolescence
raises key issues. Finally, the relevance of normative studies for understanding neural and cognitive changes in developmental disorders is discussed.
© 2005 Elsevier Ltd. All rights reserved.

Keywords: Cognition; Development; Functional MRI; DTI; ADHD

Over the last decade, investigators have become increasingly ADHD-literature, as an example of how atypical development
interested in how cognitive development is supported by the may be relevant to our understanding of typical developmental
maturation of the brain (Casey, Tottenham, Liston, & Durston, processes.
2005). Investigators interested in the interplay between the two Cognitive control may be defined as the flexible regula-
have used a host of paradigms and methods to investigate this tion of thoughts and actions in the presence of competing
association. In the following paper, we review some of the meth- ones, and is involved in many cognitive functions, including
ods that have been used to examine the behavioral and neural motor inhibition, interference inhibition, cognitive flexibility
basis of the development of cognition. This paper is not intended and attentional control. In recent years, the ability to flexibly
as an exhaustive review of the literature, but rather as a discussion adapt behavior has played a critical role in theories of cogni-
of how the study of development may benefit from neuroimaging tive development. These theories have characterized immature
approaches. We will focus largely on cognitive control, as the cognition as being particularly susceptible to interference (e.g.,
protracted developmental course of this ability into adolescence Brainerd & Reyna, 1993; Dempster, 1992; Munakata, 1998),
is relevant to a number of developmental disorders, such as atten- with younger children having difficulties in overriding compet-
tion deficit hyperactivity disorder (ADHD), childhood-onset ing attentional or behavioral responses. This susceptibility to
schizophrenia, and childhood-onset obsessive–compulsive dis- interference during childhood is illustrated by studies showing
order. Illustrative examples from the developmental literature that performance on Stroop, flanker, and go no-go tasks contin-
of other cognitive functions will also be discussed. In the final ues to develop over childhood and does not reach full maturity
section of this manuscript, we will discuss some points from the until roughly 12 years or later (e.g., Carver, Livesey, & Charles,
2001; Diamond, Cruttenden, & Nederman, 1994; Diamond &
Taylor, 1996; Enns & Akhtar, 1989; Enns & Cameron, 1987;
∗ Corresponding author. Tel.: +31 30 250 8161; fax: +31 30 250 5444. Enns, Brodeur, & Trick, 1998; Gerstadt, Hong, & Diamond,
E-mail address: S.Durston@azu.nl (S. Durston). 1994; Luria, 1961; Passler, Isaac, & Hynd, 1985; Ridderinkhof

0028-3932/$ – see front matter © 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.neuropsychologia.2005.10.010
2150 S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157

& Van der Molen, 1997; Ridderinkhof, Van der Molen, & Band, cation of this methodology to examining development of cog-
1997; Tipper, Bourque, Anderson, & Brehaut, 1989; Van der nitive control will be discussed in a subsequent section of
Meere & Stemerdink, 1999). This work shows a developmen- this paper.
tal trend in the ability to suppress information and actions over
the ages of 4–12 years, as indexed by mean reaction times and 2. Structural brain changes and cognitive development
accuracy rates. Importantly, age-related differences in accuracy
are not observed on these tasks in the absence of interfering By definition, developmental changes in cognition coincide
information (e.g., Enns et al., 1998) illustrating the differential with changes in the brain. In order to inform us about the func-
developmental trajectories for cognitive control relative to more tional organization of the brain, investigators have begun to
simple, target detection tasks. link neuroanatomical development to cognitive changes. For
example, the basal ganglia have a number of projections to and
1. Structural brain development from the prefrontal cortex and both brain regions mature rel-
atively late (Sowell et al., 1999). As this circuitry has been
Although pediatric neuroimaging is a relatively young field, implicated in goal-directed actions, development of this cir-
several studies have begun to track the anatomical course of cuitry has been suggested to relate to the development of cog-
normal brain development (see, Durston et al., 2001, for a nitively driven actions throughout childhood and adolescence.
review). These structural MRI studies have shown an increase Reiss et al. (1996) have shown that the largest increase in
in brain volume over the first few years and then relative sta- prefrontal volume, in fact, takes place during this period of
bility in brain volume by later childhood (Caviness, Kennedy, development consistent with this claim. However, these par-
Richelme, Rademacher, & Filipek, 1996; Courchesne et al., allel changes could be coincident. A number of studies have
2000; Giedd et al., 1996, 1999; Reiss, Abrams, Singer, Ross, begun to link developmental changes in brain anatomy to behav-
& Denckla, 1996; Thompson et al., 2000). The relative stabil- ior (e.g., Olesen, Nagy, Westerberg, & Klingberg, 2003; Sowell
ity of total brain volume in later development belies a dynamic et al., 2004). In one such example, Casey, Castellanos et al.
interplay of simultaneously occurring progressive and regressive (1997) showed correlations between performance on measures
events, with different regions following different time courses. of impulse control and volumetric measures of the prefrontal
Significant changes occur in regional cortical and subcortical cortex and basal ganglia in healthy children relative to children
gray matter volumes (Caviness et al., 1996; Giedd et al., 1996, with ADHD. Such correlational analyses are intriguing as they
1999; Jernigan, Trauner, Hesselink, & Tallal, 1991; Sowell, suggest a possible causal relationship between brain and behav-
Thompson, Holmes, Jernigan, & Togan, 1999; Sowell et al., ior. However, they provide only an indirect link between the
2003). Perhaps the most informative studies to date have been two, whereas advances in current imaging methods allow for
those tracking changes over time within individuals. These more direct assessment of this relationship using non-invasive
studies have shown differential development between different measures.
regions of the brain, where increases in cortical gray matter fol-
lowed by volume loss apppear to occur first in sensori-motor 3. Functional brain changes and cognitive development
and last in higher-order association cortices such as the dorso-
lateral prefrontal cortex (see Giedd, 2004 for a review; Gogtay More direct investigations of structure–function associations
et al., 2004). In another example of this approach, Sowell et al. and their development have been performed using functional
(2004) tracked changes in cortical gray matter thickness longitu- magnetic resonance imaging (fMRI). These studies suggest dif-
dinally in preadolescent children, and showed regionally specific ferent developmental trajectories of brain regions involved in
increases in classical language regions, accompanied by more cognitive control. A number of investigators have examined
widespread thinning in other cortical areas. Interestingly, they differential maturation of this function in frontal and parietal
were also able to show a relationship between cortical thinning cortices (e.g., Booth et al., 2003; Luna et al., 2001; Rubia et al.,
and improvements in language ability. 2000). For example, Adleman et al. (2002) used a Stroop task
Simultaneous with ongoing changes in gray matter, white to show that increases in activation associated with improved
matter shows linear increases in volume during childhood cognitive performance occurred by adolescence for the parietal
and into adulthood, with changes appearing to follow a pat- lobe, whereas increases in prefrontal activation continued into
tern from caudal to more rostral (Giedd et al., 1999; Jernigan adulthood. Similar findings have been reported for the develop-
et al., 1991; Paus et al., 1999; Pfefferbaum et al., 1994). ment of working memory, where increased capacity is supported
Diffusion tensor imaging (DTI), a relatively new technique by higher levels of activation in prefrontal and parietal cortices in
offers greater detail than conventional structural MRI tech- adolescents compared to children (Casey et al., 1995; Klingberg,
niques, as it provides information on the directionality and Forssberg, & Westerberg, 2002). A number of investigators have
regularity of myelinated fiber tracts. Initial studies using this suggested that maturation is not only associated with enhanced
method have confirmed changes in cortical white matter path- activation in areas that are critical for task performance, but also
ways during development (e.g., Klingberg, Vaidya, Gabrieli, with attenuation of activation in non-critical areas. For exam-
Moseley, & Hedehus, 1999). These changes presumably reflect ple, Tamm, Menon, & Reiss (2002) showed that maturation was
the ongoing myelination of axons by oligodendrocytes, enhanc- associated with more focal activation in areas critical for task
ing neuronal conduction and communication. The direct appli- performance on a go no-go task.
S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157 2151

Other studies have shown differential recruitment of sub- 4. Distinguishing between differences in performance
cortical as opposed to cortical regions with development (e.g., with age and developmental changes
Booth et al., 2003; Bunge, Dudukovic, Thomason, Vaidya, &
Gabrieli, 2002; Rubia et al., 2000). For example, Casey, Thomas, As previously discussed, development and learning coincide
Davidson, Kunz, & Franzen (2002) examined development of during childhood. As such, it is a challenge to separate the
neural systems involved in developing new stimulus-response contributions of each, and examine neural changes specific to
mappings, while overriding pre-existing ones. Here, the extent of development, learning and developmental learning. This para-
activation in subcortical structures (hippocampus for new learn- dox may be addressed in part by functional imaging studies with
ing, striatum for overriding old responses) was far greater for paradigms designed to assess which areas are critically related
children than adults, while their performance was significantly to task performance. In particular, studies equating performance
worse, suggesting that as these functions mature the pattern between groups have the potential to address such issues. In
of activation associated with them becomes more focal. This developmental studies, younger children will often have poorer
pattern occurred in parallel with greater recruitment of corti- performance on a paradigm than an older comparison-group.
cal regions with maturation. Others (e.g., Booth et al., 2004) This potentially confounds developmental changes, as differ-
have shown that better performance is not necessarily associ- ences in MR results may then be related to differences in task
ated with greater activation in development, as although greater performance or difficulty.
fronto-striatal activation was associated with better performance Investigators have addressed differences in performance
on a go no-go task, they found that better selective attention per- across developmental age groups in a number of ways. One
formance was associated with less activation in parietal areas approach is to consider subgroups at different ages that are
in a visual search task. Studies investigating cognitive develop- matched on performance. For instance, Schlaggar et al. (2002)
ment in other domains have illustrated that changes in patterns of compared adults and children, aged 7–10 years on a single
activation are not unique to studies involving cognitive control word processing task and found differences in the recruitment
(e.g., Hertz-Pannier et al., 1997; Nelson et al., 2003; Schlaggar of prefrontal and extrastriate regions. As these differences could
et al., 2002). For example, Thomas et al. (2004) showed dif- have been attributable exclusively to performance discrepancies,
ferential patterns of subcortical–cortical activation in a study of they selected a subgroup of their subjects, matched for perfor-
motor-sequence learning. mance and showed that some of the brain regions examined
Taken together, developmental neuroimaging studies of cog- showed differences attributable to age, independent of perfor-
nitive control, as well as other functions, suggest that cognitive mance (Schlaggar et al., 2002).
development is supported by changes in patterns of brain acti- An alternative approach is to incorporate different levels of
vation, including enhancement of activation in critical areas, task difficulty in the paradigm design by using a parametric
attenuation in others, and changes in the extent of activation as manipulation. This allows the post hoc comparison of groups
well as shifts in lateralization (e.g., Booth et al., 2003, 2004; at different levels of task difficulty, while matching for per-
Bunge et al., 2002; Gaillard, Balsamo, Ibrahim, Sachs, & Xu, formance. For instance, we previously developed a task incor-
2003; Hertz-Pannier et al., 1997; Luna & Sweeney, 2004; Nelson porating a parametric manipulation of a go no-go paradigm,
et al., 2003; Rubia et al., 2000; Thomas et al., 2004). In some where no-go trials were preceded by 1, 3 or 5 go trials (Durston,
studies, developmental changes in patterns of brain activation Thomas, Worden, Yang, & Casey, 2002; Durston, Thomas, Yang
could be conceptualized as a shift in patterns of activation from et al., 2002). Subjects showed an increase in errors with increas-
diffuse to more focal, where diffuse refers to larger or more ing interference from the preceding go-trials. We used this task
areas of activation, and focal indicates smaller areas of activa- to investigate the development of the neural basis of response
tion, with greater magnitude of signal change (Casey, Trainor inhibition (the stopping on no-go trials) and found that success-
et al., 1997; Casey et al., 2002; Gaillard et al., 2000; Hertz- ful response inhibition was associated with greater activation
Pannier et al., 1997; Holland et al., 2001; Luna et al., 2001; of prefrontal and parietal regions for children than for adults.
Monk et al., 2003; Moses et al., 2002; Nelson et al., 2003; In adults, activation in ventral prefrontal regions increased with
Schlaggar et al., 2002; Tamm et al., 2002; Turkeltaub, Gareau, increasing interference from go-trials. However, unlike adults,
Flowers, Zeffiro, & Eden, 2003) similar to changes seen in adult the circuitry appeared to be maximally activated in children
learning (Karni et al., 1995, 1998). These changes may repre- when suppressing a behavioral response regardless of the num-
sent a fine-tuning of relevant neural systems (Johnson, Halit, ber of preceding responses. These findings confirm that imma-
Grice, & Karmiloff-Smith, 2002), or related developmental ture cognition is more susceptible to interference, as children
changes, such as new brain regions coming online and reduced make more errors than adults. Furthermore, they show that this
involvement of others, and may be related to shifts in cognitive diffference is paralleled by maturational differences in under-
strategy, in some studies. Differences between developmental lying fronto-striatal circuitry (Durston, Thomas, Yang et al.,
studies with respect to which areas are seen to come online 2002).
with development illustrate that neural changes observed using A third possible approach is to examine correlations between
functional imaging are critically dependent on the task used. neural activity and performance and age. Performance and age
Certain brain regions may become more involved in one aspect can both be examined independently by covarying the other. Sev-
of cognition with development, while becoming less involved eral groups have used this approach to tease apart performance-
in others. based versus age-based differences in developmental studies
2152 S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157

(e.g., Casey, Trainor et al., 1997; Casey et al., 2002; Thomas activation. A parallel cross-sectional analysis based on a sec-
et al., 2004). ond group of children, the same ages as the longitudinal group,
showed less specific results, confirming added sensitivity in lon-
5. Developmental progressions versus single snap-shots gitudinal approaches.
in time
6. Brain connectivity and cognitive development
The previously reviewed studies illustrate that if we are to
explain changes in a dynamic system, it is essential to exam- The developmental shift from diffuse to focal cortical
ine developmental progressions in behavior and the brain, rather activation may reflect the functional consequences of synaptic
than examining behavior at single snap-shots in time. The impor- pruning and other regressive processes in combination with
tance of examining developmental trajectories is perhaps most strengthening of relevant connections. However, few studies to
obvious in understanding and determining whether atypical date have related maturational changes in brain connectivity
development is due to a developmental delay or deficit. A num- to their functional correlates. New imaging techniques such as
ber of behaviors may be completely appropriate at one age but DTI are beginning to allow us to do exactly that (see Fig. 1;
inappropriate at another. Clinical disorders, such as attention e.g., Klingberg et al., 1999; Olesen et al., 2003). DTI measures
deficit hyperactivity disorder, may reflect exaggerated and/or water-diffusion in the brain, and can be used to estimate the
residual processes that do not necessarily diminish or change regularity of white matter, as an indirect measure of myelination
with maturity. Understanding the normal progression in a spe- and connectivity (Mori, Crain, Chacko, & van Zijl, 1999; Paus et
cific behavioral or neural system will have a significant impact al., 2001). For example, Liston et al. recently used this technique
in determining the biological substrates of clinical disorders and to relate functional changes in frontostriatal networks with the
targeting effective treatments and interventions. development of performance on the previously described cog-
Although the majority of imaging studies to date have nitive control task (see Section 4; Durston, Thomas, Worden et
examined developmental progressions using a cross-sectional al., 2002; Durston, Thomas, Yang et al., 2002) to changes in the
approach, there are a number of advantages to tracking changes connectivity between these structures (Liston et al., in press). An
within individuals using a longitudinal approach. This is espe- automated fiber-tracking algorithm was used to delineate white
cially important in biological systems, as there is an extreme matter fibers projecting from ventral prefrontal cortex to the
degree of individual variability in brain structure, relative to caudate nucleus. These two structures are known to contribute to
developmental variability, during childhood and adolescence cognitive control based on previous functional imaging studies
(Caviness et al., 1996). There are several examples of longi- (Casey, Trainor et al., 1997; Durston et al., in press). A posterior
tudinal approaches in Section 1 (see, Giedd, 2004, for a review). tract, not expected to contribute to this ability, was also delin-
In a recent functional imaging study, we used a version of the eated with this method as a control. Diffusion in prefrontal and
previously described go no-go paradigm to track the develop- posterior tracts became more restricted between the ages of 7 and
ment of cognitive control longitudinally in a sample of children 30 years. This shift was paralleled by an age-associated increase
as they reached adolescence (Durston et al., in press). We showed in efficiency in cognitive performance. Further, fronto-striatal
a developmental shift in patterns of cortical activation from radial diffusivities that are sensitive to changes in myelination,
diffuse to focal activity. Here, we showed that sensori-motor predicted individual differences in reaction time, when con-
regions uncorrelated with task performance were recruited less, trolling for age and accuracy. This pattern was not observed in
while a region in the ventral prefrontal cortex showed enhanced posterior tracts (Liston et al., in press). Collectively, these results
recruitment. Activity in this region has been shown to be cor- indicate that maturation of prefrontal white matter and enhanced
related with performance on this task in prior studies (Casey, connectivity between fronto-striatal structures contributes to
Trainor et al., 1997; Durston, Thomas, Worden et al., 2002; a developing capacity and individual variability in cognitive
Durston, Thomas, Yang et al., 2002; Konishi, Nakajima, Uchida, control.
Sekihara, & Miyashita, 1998; Konishi et al., 1999) and in the
current study (Durston et al., in press). In contrast, activa- 7. Atypical development of cognitive control: attention
tion in structures such as the primary motor cortex remained deficit hyperactivity disorder
unchanged for the simple comparison of responding (go) versus
not responding (no-go) during performance of the task. Inter- The emergence of clinical disorders of cognitive control
estingly, the significant negative correlation between accuracy throughout childhood and adolescence underscores the impor-
and activation in ventral prefrontal cortex was accompanied by tance of understanding the neurobiological basis of such control
a trend-level positive correlation between age and activation (see Durston et al., 2001 for a review). Attention Deficit Hyper-
in this region. This suggests that at this crucial developmen- activity Disorder (ADHD) is currently the best-characterized
tal stage, subjects may in fact be responding more impulsively, childhood disorder associated with deficits in this ability, and
related to faster responses and therefore need more top-down a discussion of this disorder may aid in our understanding
control in overriding a response. This finding ties in with anec- of typical development of this ability. ADHD first manifests
dotal evidence of increased impulsive and risk-taking behavior in early childhood, before the age of 7 and is characterized
in adolescence, and illustrates the complexity of the relation- by age-inappropriate levels of impulsive, hyperactive of dis-
ships between development, performance and patterns of brain tractible behavior (i.e., cognitive control deficits). Children with
S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157 2153

Fig. 1. How diffusion tensor imaging can be combined with anatomical and functional MRI to investigate functional networks. Panel A: the caudate nucleus is
defined on an anatomical MRI scan, as a seedpoint for fiber tracking. Panel B: functional activation maps from a single subject during a cognitive control task,
overlayed on the anatomical MRI scan. Panels C and D: vectors calculated from the DTI scan, where the colors represent the direction of the vector, overlayed on
a section of the anatomical scan and, finally. Panel E: fibers tracked on the basis of the vectors in panel D, connecting both caudate nuclei with regions in frontal
cortex, including the frontal region activated in panel B.
2154 S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157

ADHD have poor performance on tasks that require inhibitory children with ADHD showing that the striatum was the most
control, such as go-no go and stop tasks. This finding has robust region of difference between children with and with-
lead some investigators to theorize that a deficit in inhibitory out the disorder. Other studies have shown reduced activation
components of cognitive control may be central to this disorder of frontal circuitry in adolescents with ADHD (Rubia et al.,
(Barkley, 1997). This theory in turn implicates fronto-striatal 1999) or abnormal recruitment of brain networks in ADHD, dur-
circuitry in ADHD given its involvement in cognitive control ing performance of cognitive control tasks (Bush et al., 1999;
(see Durston, 2003, for a review). As discussed previously, the Schweitzer et al., 2000).
development of fronto-striatal circuitry is protracted as synaptic In summary, this literature suggests local abnormalities in
pruning and myelination of prefrontal fibers proceeds slowly cerebral activation in ADHD, with a hypo-perfusion of pre-
throughout late childhood and adolescence (Casey, Castellanos frontal and possibly striatal areas. Functional studies using cog-
et al., 1997; Casey, Trainor et al., 1997; Conel, 1939–1967; nitive control paradigms show deficits in frontal and striatal
Durston, Thomas, Worden et al., 2002; Durston, Thomas, function, further implicating this circuitry. These results con-
Yang et al., 2002; Huttenlocher, 1979; Klingberg et al., 1999; firm the evidence from the developmental fMRI literature that
Paus et al., 2001; Sowell et al., 1999). Children’s capacity the development of cognitive control is supported by the mat-
for cognitive control develops across the first decade with uration of fronto-striatal circuitry, as poor cognitive control in
younger children more susceptible to interference on a variety ADHD appears to be related to changes in this circuitry.
of tasks in this domain (Casey, Trainor et al., 1997; Casey et
al., 2002; Diamond, 1990; Munakata & Yerys, 2001). Children 8. Conclusions
recruit distinct and often larger, more diffuse fronto-striatal
regions when performing cognitive control tasks than adults do, It is well established that, with development, significant
suggesting development within and refinement of projections changes occur in cognitive ability, especially in the context of
to and from these regions with maturation (Bunge et al., 2002; overriding competing thoughts and actions. In this paper, we
Casey, Trainor et al., 1997; Casey et al., 2002; Luna et al., 2001). discuss the added value of neuroimaging approaches to under-
The imaging literature on ADHD implicates fronto-striatal standing these changes in cognitive development. We focus
circuitry. The first functional imaging studies used single pho- on cognitive control, as its protracted developmental course
ton emission computed tomography (SPECT) to look at cerebral into adolescence raises key points. Changes in brain structure
blood flow in children with ADHD. The authors published a occur in parallel with changes in cognition. However, to show
series of papers in which they used this method and found that these changes are not merely coincident, it is necessary
evidence for reduced blood flow in striatal regions in this dis- to directly link developmental changes in brain anatomy to
order (Lou, Henriksen, & Bruhn, 1984; Lou et al., 1989; Lou, behavior.
Henriksen, & Bruhn, 1990). Although these findings constituted Structural brain development involves a dynamic interplay
the first evidence of functional cerebral deficits in ADHD, the of simultaneously occurring progressive and regressive events,
spatial resolution of SPECT is low and ethical constraints meant with different regions following different time courses. Inves-
that adequate control subjects could often not be included in tigators using correlational analyses have suggested that these
these studies. A series of papers looking at cerebral metabolism maturational changes in brain anatomy may be related to cog-
in adults and adolescents with ADHD using positron emis- nitive development. Current imaging methods including func-
sion tomography (PET) at first suggested a reduction in global tional MRI and DTI allow a more direct linking of the two. FMRI
metabolism in ADHD (Zametkin et al., 1990), although later investigations have confirmed early evidence from anatomical
reports suggested that these findings might be more specific to studies implicating fronto-striatal circuitry in the development of
females with ADHD than to males (Ernst et al., 1994, 1998; cognitive control. Generally, developmental changes in patterns
Zametkin et al., 1993). In line with suggestions of reduced base- of brain activity appear to involve a shift from diffuse to more
line levels of cerebral metabolism and perfusion in ADHD, sev- focal activation, likely representing a fine-tuning of relevant neu-
eral studies have shown that administration of methylphenidate ral systems. Here, the shift to focal activation may represent
increases cerebral metabolism and perfusion in frontal regions increases in the magnitude or extent of activation in regions
(Kim et al., 2001; Matochik et al., 1993, 1994). critical to task performance, whereas the shift from diffuse acti-
More recently, studies have been published using functional vation represents attenuation of activation in other, less critical,
MRI to investigate brain activity in ADHD during the perfor- cortical areas, either in the number of regions activated, or in
mance of cognitive control tasks (Vaidya et al., 1998; Rubia et the extent of activated tissue. Other developmental changes may
al., 1999; Bush et al., 1999; Durston et al., 2003). Vaidya et include new brain regions coming online, while the involvement
al. showed that children with ADHD had higher frontal activa- of others is reduced, and may be related to shifts in cognitive
tion and lower striatal activation than control children during strategies.
response inhibition, whereas administration of methylphenidate One of the challenges in investigating the interplay between
lead to improved performance associated with increased frontal cognitive development and maturation of the brain is to sepa-
activation for both groups and an increase in striatal activation rate the contributions of neural changes specific to development
for the children with ADHD, perhaps representing a normaliza- and learning. Training studies in children will be necessary to
tion of striatal function with medication (Vaidya et al., 1998). tease these apart ultimately (e.g., Klingberg et al., 2005), but
Durston et al. (2003), in part, replicated these findings in younger imaging studies equating performance are critical in separating
S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157 2155

which changes are related to maturation versus differences in Casey, B. J., Trainor, R. J., Orendi, J. L., Schubert, A. B., Nystrom, L. E.,
performance between groups. Other, new and promising imag- Giedd, J. N., et al. (1997). A developmental functional MRI study of
prefrontal activation during performance of a go-nogo task. Journal of
ing approaches include controlling for large variation between
Cognitive Neuroscience, 9, 835–847.
individuals by tracking both cognitive and brain development Caviness, V. S., Kennedy, D. N., Richelme, C., Rademacher, J., & Filipek, P.
in the same individuals over time in longitudinal studies. Fur- A. (1996). The human brain age 7–11 years: a volumetric analysis based
thermore, DTI is now being used to relate the maturation of on magnetic resonance images. Cerebral Cortex, 6, 726–736.
white matter tracts to cognitive development, and may refine Conel, J. L. (1939–1967). The postnatal development of the human cerebral
cortex. Cambridge, MA: Harvard University Press.
our understanding of the role of local changes in white matter
Courchesne, E., Chisum, H. J., Townsend, J., Cowles, A., Covington, J.,
to cognitive development (see Fig. 1). Finally, investigations of Egaas, B., et al. (2000). Normal brain development and aging: quantitative
developmental disorders such as ADHD have the potential to analysis at in vivo MR imaging in healthy volunteers. Radiology, 216(3),
inform us on the role of circuitry affected in these disorders, by 672–682.
relating neural to cognitive changes in affected individuals. Dempster, F. N. (1992). The rise and fall of the inhibitory mechanism:
Towards a unified theory of cognitive development and aging. Devel-
opmental Review, 12, 45–75.
Acknowledgements Diamond, A. (1990). Developmental time course in human infants and infant
monkeys, and the neural bases of inhibitory control in reaching. Annals
The authors gratefully acknowledge Drs. R.M. Mandl and of the New York Academy of Sciences, 608, 637–676.
Diamond, A., Cruttenden, L., & Nederman, D. (1994). AB with mul-
M.J. Mulder for their assistance in constructing Fig. 1, and
tiple wells: I. Why are multiple wells sometimes easier than two
two anonymous reviewers for their helpful suggestions. The wells? II. Memory or memory + inhibition. Developmental Psychology,
work described in this paper was supported in part by an NWO 30, 192–205.
VENI-grant to SD, and P01 MH62196, R01 MH63255, and R21 Diamond, A., & Taylor, C. (1996). Development of an aspect of executive
DA15882 to BJC. control: development of the abilities to remember what I said and to “do
as I say, not as I do”. Developmental Psychobiology, 29, 315–334.
Durston, S. (2003). A review of the biological bases of ADHD: What have
References we learned from imaging studies? Mental Retardation and Developmental
Disabilities Research Reviews, 9(3), 184–195.
Adleman, N. E., Menon, V., Blasey, C. M., White, C. D., Warsofsky, I. S., Durston, S., Davidson, M. C., Tottenham, N. T., Galvan, A., Spicer, J., Fos-
Glover, G. H., et al. (2002). A developmental fMRI study of the Stroop sella, J. A., & Casey, B. J. (in press). A shift from diffuse to focal cortical
color-word task. Neuroimage, 16(1), 61–75. activity with development. Developmental Science.
Barkley, R. A. (1997). Behavioral inhibition, sustained attention and execu- Durston, S., Hulshoff, H. E., Casey, B. J., Giedd, J. N., Buitelaar, J. K.,
tive functions: constructing a unifying theory of ADHD. Psychological & Van Engeland, H. (2001). Anatomical MRI of the developing human
Bulletin, 21, 65–94. brain: What have we learned? Journal of the American Academy of Child
Booth, J. R., Burman, D. D., Meyer, J. R., Lei, Z., Trommer, B. L., Daven- and Adolescent Psychiatry, 40(9), 1012–1020.
port, N. D., et al. (2003). Neural development of selective attention and Durston, S., Thomas, K. M., Worden, M. S., Yang, Y., & Casey, B. J. (2002).
response inhibition. Neuroimage, 20, 737–751. The effect of preceding context on inhibition: An event-related fMRI
Booth, J. R., Burman, D. D., Meyer, J. R., Trommer, B. L., Davenport, N. study. Neuroimage, 16(2), 449–453.
D., Parrish, T. B., et al. (2004). Brain-behavior correlation in children Durston, S., Thomas, K. M., Yang, Y., Ulug, A. M., Zimmerman, R. D., &
depends on the neurocognitive network. Human Brain Mapping, 23(2), Casey, B. J. (2002). The development of neural systems involved in over-
99–108. riding behavioral responses: An event-related fMRI study. Developmental
Brainerd, C. J., & Reyna, V. F. (1993). Memory independence and memory Science, 5(4), F9–F16.
interference in cognitive development. Psychological Review, 100, 42–67. Durston, S., Tottenham, N. T., Thomas, K. M., Davidson, M. C., Eigsti, I. M.,
Bunge, S. A., Dudukovic, N. M., Thomason, M. E., Vaidya, C. J., & Gabrieli, Yang, Y., et al. (2003). Differential patterns of striatal activation in young
J. D. E. (2002). Immature frontal lobe contributions to cognitive control children with and without ADHD. Biological Psychiatry, 53, 871–878.
in children: Evidence from fMRI. Neuron, 33, 301–311. Enns, J. T., & Akhtar, N. (1989). A developmental study of filtering in visual
Bush, G., Frazier, J. A., Rauch, S. L., et al. (1999). Anterior cingulate cortex attention. Child Development, 60, 1188–1199.
dysfunction in attention-deficit/hyperactivity disorder revealed by fMRI Enns, J. T., Brodeur, D. A., & Trick, L. M. (1998). Selective attention over
and the Counting Stroop. Biological Psychiatry, 45, 1542–1552. the life span: Behavioral measures. In J. E. Richards (Ed.), Cognitive
Carver, A. C., Livesey, D. J., & Charles, M. (2001). Age related changes neuroscience of attention: A developmental perspective (pp. 393–418).
in inhibitory control as measured by stop signal task performance. The Mahwah, USA: Lawrence Erlbaum Associates, Inc.
International Journal of Neuroscience, 107, 43–61. Enns, J. T., & Cameron, S. (1987). Selective attention in young children:
Casey, B. J., Castellanos, F. X., Giedd, J. N., Marsh, W. L., Hamburger, The relations between visual search, filtering, and priming. Journal of
S. D., Schubert, A. B., et al. (1997). Implication of right frontostriatal Experimental Child Psychology, 44, 38–63.
circuitry in response inhibition and attention-deficit/hyperactivity disorder. Ernst, M., Liebenauer, L. L., King, A. C., et al. (1994). Reduced brain
Journal of the American Academy of Child and Adolescent Psychiatry, metabolism in hyperactive girls. Journal of the American Academy of
36(3), 374–383. Child and Adolescent Psychiatry, 33, 858–868.
Casey, B. J., Cohen, J. D., Jezzard, P., Turner, R., Noll, D. C., Trainor, Ernst, M., Zametkin, A. J., Phillips, R. L., et al. (1998). Age-related changes
R. J., et al. (1995). Activation of prefrontal cortex in children during a in brain glucose metabolism in adults with attention-deficit/hyperactivity
nonspatial working memory task with functional MRI. Neuroimage, 2, disorder and control subjects. The Journal of Neuropsychiatry and Clin-
221–229. ical Neurosciences, 10, 168–177.
Casey, B. J., Thomas, K. M., Davidson, M. C., Kunz, K., & Franzen, P. L. Gaillard, W. D., Balsamo, L. M., Ibrahim, Z., Sachs, B. C., & Xu, B. (2003).
(2002). Dissociating striatal and hippocampal function developmentally fMRI identifies regional specialization of neural networks for reading in
with a stimulus-response compatibility task. Journal of Neuroscience, young children. Neurology, 60(1), 94–100.
22(19), 8647–8652. Gaillard, W. D., Hertz-Pannier, L., Mott, S. H., Barnett, A. S., LeBihan, D.,
Casey, B. J., Tottenham, N. T., Liston, C., & Durston, S. (2005). Imaging & Theodore, W. H. (2000). Functional anatomy of cognitive development:
the developing brain: What have we learned? TICS, 9(3), 104–110. fMRI of verbal fluency in children and adults. Neurology, 11, 180–185.
2156 S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157

Gerstadt, C. L., Hong, Y. J., & Diamond, A. (1994). The relationship between Lou, H. C., Henriksen, L., & Bruhn, P. (1984). Focal cerebral hypoperfusion
cognition and action: Performance of 3 1/2–7 years old on a Stroop-like in children with dysphasia and/or attention deficit disorder. Archives of
day-night test. Cognition, 53, 129–153. Neurology, 41, 825–829.
Giedd, J. N. (2004). Structural magnetic resonance imaging of the adolescent Lou, H. C., Henriksen, L., Bruhn, P., et al. (1989). Striatal dysfunction in
brain. Annals of the New York Academy of Science, 102, 77–85. attention deficit and hyperkinetic disorder. Archives of Neurology, 46,
Giedd, J. N., Blumenthal, J., Jeffries, N. O., Castellanos, F. X., Lui, H., 48–52.
Zijdenbos, A., et al. (1999). Brain development during childhood and Luna, B., & Sweeney, J. A. (2004). The emergence of collaborative brain
adolescence: a logitudinal MRI study. Nature Neuroscience, 10, 861– function: FMRI studies of the development of response inhibition. Annals
863. of the New York Academy of Science, 1021, 296–309.
Giedd, J. N., Snell, J. W., Lange, N., Rajapakse, J. C., Casey, B. J., Kaysen, Luna, B., Thulborn, K. R., Munoz, D. P., Merriam, E. P., Garver, K.
D., et al. (1996). Quantitative magnetic resonance imaging of human brain E., Minshew, N. J., et al. (2001). Maturation of widely distributed
develoopment: ages 4–18. Cerebral Cortex, 6, 551–560. brain function subserves cognitive development. Neuroimage, 13(5), 786–
Gogtay, N., Giedd, J. N., Lusk, L., Hayashi, K. M., Greenstein, D., Vaituzis, 793.
A. C., et al. (2004). Dynamic mapping of human cortical develop- Luria, D. M. (1961). The role of speech in the regulation of normal and
ment during childhood through early adulthood. Proceedings of the abnormal behavior. New York: Liveright.
National Academy of Sciences of the United States of America, 101(21), Matochik, J. A., Liebenauer, L. L., King, A. C., et al. (1994). Cerebral
8174–8179. glucose metabolism in adults with attention deficit hyperactivity disorder
Hertz-Pannier, L., Gaillard, W. D., Mott, S. H., Cuenod, C. A., Bookheimer, after chronic stimulant treatment. American Journal of Psychiatry, 151,
S. Y., Weinstein, S., et al. (1997). Noninvasive assessment of language 658–664.
dominance in children adn adolescents with functional MRI: A prelimi- Matochik, J. A., Nordahl, T. E., Gross, M., et al. (1993). Effects of acute
nary study. Neurology, 48(4), 1003–1012. stimulant medication on cerebral metabolism in adults with hyperactivity.
Holland, S. K., Plante, E., Weber Byars, A., Strawsburg, R. H., Schmithorst, Neuropsychopharmacology, 8, 377–386.
V. J., & Ball, W. S. J. (2001). Normal fMRI brain activation patterns in Monk, C. S., McClure, E. B., Nelson, E. E., Zarahn, E., Bilder, R. M.,
children performing a verb generation task. Neuroimage, 14, 837–843. Leibenluft, E., et al. (2003). Adolescent immaturity in attention-related
Huttenlocher, P. R. (1979). Synaptic density in human frontal brain engagement to emotional facial expressions. Neuroimage, 20(1),
cortex—developmental changes and effects of aging. Brain Research, 420–428.
163(2), 195–205. Mori, S., Crain, B. J., Chacko, V. P., & van Zijl, P. C. M. (1999). Three-
Jernigan, T. L., Trauner, D. A., Hesselink, J. R., & Tallal, P. A. (1991). Mat- dimensional tracking of axonal projections in the brain by magnetic
uration of human cerebrum observed in vivo during adolescence. Brain, resonance imaging. Annals of Neurology, 45, 265–269.
114(Pt 5), 2037–2049. Moses, P., Roe, K., Buxton, R. B., Wong, E. C., Frank, L. R., & Stiles,
Johnson, M. H., Halit, H., Grice, S. J., & Karmiloff-Smith, A. (2002). Neu- J. (2002). Functional MRI of global and local processing in children.
roimaging of typical and atypical development: A perspective from multi- Neuroimage, 16(2), 415–424.
ple levels of analysis. Development & Psychopathology, 14(3), 521–536. Munakata, Y. (1998). Infant perseveration and implications for object perma-
Karni, A., Meyer, G., Rey-Hipolito, C., Jezzard, P., Adams, M. M., Turner, R., nence theories: A PDP model of the AB task. Developmental Science, 1,
et al. (1995). Functional MRI evidence for adult motor cortex plasticity 161–184.
during motor skill learning. Nature, 377, 155–158. Munakata, Y., & Yerys, B. E. (2001). All together now: when dissociations
Karni, A., Meyer, G., Jezzard, P., Adams, M. M., Turner, R., & Ungerleider, between knowledge and action disappear. Psychonomic Science, 12(4),
L. G. (1998). The acquisition of skilled motor performance: fast and 335–337.
slow experience-driven changes in primary motor cortex. Proceedings of Nelson, E. E., McClure, E. B., Monk, C. S., Zarahn, E., Leibenluft, E., Pine,
the National Academy of Sciences of the United States of America, 95, D. S., et al. (2003). Developmental differences in neuronal engagement
861–868. during implicit encoding of emotional faces: an event-related fMRI study.
Kim, B. N., Lee, J. S., Cho, S. C., et al. (2001). Methylphenidate increased Journal of Child Psychology and Psychiatry, 44(7), 1015–1024.
regional cerebral blood flow in subjects with attention deficit/hyperactivity Olesen, P. J., Nagy, Z., Westerberg, H., & Klingberg, T. (2003). Combined
disorder. Yonsei Medical Journal, 42, 19–29. analysis of DTI and fMRI data reveals a joint maturation of white and
Klingberg, T., Fernell, E., Olesen, P. J., Johnson, M., Gustafsson, P., grey matter in a fronto-parietal network. Brain Research. Cognitive Brain
Dahlstrom, K., et al. (2005). Computerized training of working mem- Research, 18(1), 48–57.
ory in children with ADHD–a randomized, controlled trial. Journal of Passler, M. A., Isaac, W., & Hynd, G. W. (1985). Neuropsychological devel-
the American Academy of Child and Adolescent Psychiatry, 44(2), 177– opment of behavior attributed to frontal lobe functioning in children.
186. Developmental Neuropsychology, 1, 349–370.
Klingberg, T., Forssberg, H., & Westerberg, H. (2002). Increased brain activ- Paus, T., Collins, D. L., Evans, A. C., Leonard, G., Pike, B., & Zijdenbos,
ity in frontal and parietal cortex underlies the development of visuospatial A. (2001). Maturation of white matter in the human brain: a review of
working memory capacity during childhood. Journal of Cognitive Neu- magnetic resonance studies. Brain Research Bulletin, 54(3), 255–266.
roscience, 14(1), 1–10. Paus, T., Zijdenbos, A., Worsley, K., Collins, D. L., Blumenthal, J., Giedd,
Klingberg, T., Vaidya, C. J., Gabrieli, J. D., Moseley, M. E., & Hedehus, J. N., et al. (1999). Structural maturation of neural pathways in children
M. (1999). Myelination and organization of the frontal white matter in and adolescents: in vivo study. Science, 283(5409), 1908–1911.
children: a diffusion tensor MRI study. Neuroreport, 10, 2817–2821. Pfefferbaum, A., Mathalon, D. H., Sullivan, E. V., Rawles, J. M., Zipursky,
Konishi, S., Nakajima, K., Uchida, I., Kikyo, H., Kameyama, M., & R. B., & Lim, K. O. (1994). A quantitative magnetic resonance imaging
Miyashita, Y. (1999). Common inhibitory mechanism in inferior prefrontal study of changes in brain morphology from infancy to late adulthood.
cortex revealed by event-related functional MRI. Brain, 122, 981–991. Archives of Neurology, 51, 874–887.
Konishi, S., Nakajima, K., Uchida, I., Sekihara, K., & Miyashita, Y. (1998). Reiss, A. L., Abrams, M. T., Singer, H. S., Ross, J. L., & Denckla, M.
No-go dominant brain activity in human inferior prefrontal cortex revealed B. (1996). Brain development, gender and IQ in children. A volumetric
by functional magnetic resonance imaging. European Journal of Neuro- imaging study. Brain, 119, 1763–1774.
science, 10, 1209–1213. Ridderinkhof, K. R., & Van der Molen, M. W. (1997). Mental resources,
Liston, C., Watts, R., Tottenham, N., Davidson, M.C., Niogi, S., Ulug, A., et processing speed, and inhibitory control: A developmental perspective.
al. (in press). Frontostriatal microstructure predicts individual differences Biological Psychology, 45, 241–261.
in cognitive control. Cerebral Cortex. Ridderinkhof, K. R., Van der Molen, M. W., & Band, G. P. H. (1997). Sources
Lou, H. C., Henriksen, L., & Bruhn, P. (1990). Focal cerebral dysfunction of interference from irrelevant information: A developmental study. Jour-
in developmental learning disabilities. Lancet, 335, 8–11. nal of Experimental Child Psychology, 65, 315–341.
S. Durston, B.J. Casey / Neuropsychologia 44 (2006) 2149–2157 2157

Rubia, K., Overmeyer, S., Taylor, E., Brammer, M., Williams, S. C., Sim- Thomas, K. M., Hunt, R., Vizueta, N., Sommer, T., Durston, S., Yang, Y.,
mons, A., et al. (2000). Functional frontalisation with age: mapping et al. (2004). Evidence of developmental differences in implicit sequence
neurodevelopmental trajectories with age. Neuroscience and Biobehav- learning: An fMRI study in children and adults. Journal of Cognitive
ioral Reviews, 24, 13–19. Neuroscience, 16(8), 1339–1351.
Rubia, K., Overmeyer, S., Taylor, E., et al. (1999). Hypofrontality in attention Tipper, S. P., Bourque, T. A., Anderson, S. H., & Brehaut, J. C. (1989).
deficit hyperactivity disorder during higher-order motor control: a study Mechanisms of attention: A developmental study. Journal of Experimental
with functional MRI. American Journal of Psychiatry, 156, 891–896. Child Psychology, 48, 353–378.
Schlaggar, B. L., Brown, T. T., Lugar, H. M., Visscher, K. M., Miezin, F. M., Thompson, P. M., Giedd, J. N., Woods, R. P., MacDonald, D., Evans, A. C., &
& Petersen, S. E. (2002). Functional neuroanatomical differences between Toga, A. W. (2000). Growth patterns in the developing brain detected by
adults and school-age children in the processing of single words. Science, using continuum mechanical tensor maps. Nature, 404(6774), 190–193.
296(5572), 1476–1479. Turkeltaub, P. E., Gareau, L., Flowers, D. L., Zeffiro, T. A., & Eden, G. F.
Schweitzer, J. B., Faber, T. L., Grafton, S. T., et al. (2000). Alterations (2003). Development of neural mechanisms for reading. Nature Neuro-
in the functional anatomy of working memory in adult attention deficit science, 6(7), 767–773.
hyperactivity disorder. American Journal of Psychiatry, 157, 278–280. Vaidya, C. J., Austin, G., Kirkorian, G., Ridlehuber, H. W., Desmond, J.
Sowell, E. R., Peterson, B. S., Thompson, P. M., Welcome, S. E., Henkenius, E., Glover, G. H., et al. (1998). Selective effects of methylphenidate in
A. L., & Toga, A. W. (2003). Mapping cortical change across the human attention deficit hyperactivity disorder: a functional magnetic resonance
life span. Nature Neuroscience, 6(3), 309–315. study. Proceedings of the National Academy of Sciences of the United
Sowell, E. R., Thompson, P. M., Holmes, C. J., Jernigan, T. L., & Togan, States of America, 95, 14494–14499.
A. W. (1999). In vivo evidence for post-adolescent brain maturation in Van der Meere, J., & Stemerdink, N. (1999). The development of state reg-
frontal and striatal regions. Nature Neuroscience, 2(10), 859–861. ulation in normal children: An indirect comparison with children with
Sowell, E. R., Thompson, P. M., Leonard, C. M., Welcome, S. E., Kan, ADHD. Developmental Neuropsychology, 16, 213–220.
E., & Togan, A. W. (2004). Longitudinal mapping of cortical thickness Zametkin, A. J., Liebenauer, L. L., Fitzgerald, G. A., et al. (1993). Brain
and brain growth in normal children. Journal of Neuroscience, 24(38), metabolism in teenagers with attention-deficit hyperactivity disorder.
8223–8231. Archives of General Psychiatry, 50, 333–340.
Tamm, L., Menon, V., & Reiss, A. L. (2002). Maturation of brain function Zametkin, A. J., Nordahl, T. E., Gross, M., et al. (1990). Cerebral glucose
associated with response inhibition. Journal of the American Academy of metabolism in adults with hyperactivity of childhood onset. New English
Child and Adolescent Psychiatry, 41(10), 1231–1238. Journal of Medicine, 323, 1361–1366.

Anda mungkin juga menyukai