You are on page 1of 15

REVIEWS

B R A I N AG E I N G

The cognitive neuroscience of ageing


Cheryl Grady
Abstract | The availability of neuroimaging technology has spurred a marked increase in the
human cognitive neuroscience literature, including the study of cognitive ageing. Although
there is a growing consensus that the ageing brain retains considerable plasticity of
function, currently measured primarily by means of functional MRI, it is less clear how age
differences in brain activity relate to cognitive performance. The field is also hampered by
the complexity of the ageing process itself and the large number of factors that are
influenced by age. In this Review, current trends and unresolved issues in the cognitive
neuroscience of ageing are discussed.

Working memory
Age differences in cognitive function have been stud- brain activity, including a lack of efficiency in the use
The short-term retention and ied for many years, and it is well-established that older of neural resources or a reduction in the selectivity of
utilization of information. A adults have particular difficulty with episodic memory, responses, which is known as dedifferentiation23.
classic example is looking up which is defined as the conscious recollection of events1. Another issue is how brain activity is related to other
a phone number and
In the laboratory, these age differences in episodic aspects of brain ageing, such as changes in structure
remembering it long enough
to dial. memory are manifested by a reduced ability to learn and (volumes or white matter myelination, for example) or
retrieve both non-verbal and verbal material, such as a changes in neurotransmitters. There is also the ques-
list of words2. Substantial age-related differences are also tion of how age differences in brain function might
seen in tasks involving working memory3,4, attention5–7 be affected by undetected neuropathological changes
and task switching 8–10, all of which can be considered as due to dementing illnesses. That is, some otherwise
types of high level ‘executive’ functions. Older adults are healthy older adults might eventually be diagnosed with
also more susceptible to the effects of distracting inter- Alzheimer’s disease, and the ‘silent’ pathological pro-
ference during cognitive tasks11,12 and have a generally cesses in their brains might account for some of the age
slower processing speed13. Nevertheless, some aspects differences reported in the literature. The purpose of this
of cognition are maintained with age, such as seman- Review is to cover some recent developments in the field
tic memory (the accumulation of knowledge about the that address these long-standing issues and to discuss
world14,15) and emotional regulation16,17. In addition, age some interesting new trends in this area of research.
differences in cognition are not immutable: for exam- Before reviewing this work on cognitive ageing and
ple, the experimental conditions under which memory the brain, it is important to note that there is general
is studied in older adults can be modified so that age consensus that the blood-oxygen-level-dependent
differences are reduced or eliminated18. A challenge in (BOLD) signal obtained from fMRI is a reasonable,
this field has been to understand the brain mechanisms although indirect, index of neural activity, especially
that might underlie better or worse performance in the synaptic activity reflected in local field poten-
old adults. tials24,25. In regard to the use of fMRI to study ageing,
It is on this challenge that functional and structural peak stimulus-related BOLD responses are similar
neuroimaging studies of ageing have focused, and in the in young and older adults26–28, although some work
past decade, functional MRI (fMRI) studies have pro- has shown that the magnitude of the BOLD response
vided ample evidence of age differences in task-related can be reduced in older adults, at least in some brain
brain activity 19,20. However, the interpretation of these regions29. In addition, it is important to keep in mind
The Rotman Research differences is difficult, as sometimes brain activity is that there are alterations in the cerebral vasculature
Institute at Baycrest, 3560 reduced in older adults relative to younger adults and with age, and these have the potential to influence the
Bathurst Street, Toronto, sometimes it is increased. Decreased brain activity has BOLD signal in as yet unknown ways 30,31. Although
Ontario M6A 2E1, Canada. typically been interpreted as a reflection of cognitive more research is required to understand the impact of
e-mail:
cgrady@rotman-baycrest.
deficits in older adults21, and increased activity has often ageing on the physiology underlying the BOLD signal,
on.ca been interpreted as compensatory 22. However, other the relatively small age differences noted in the proper-
doi:10.1038/nrn3256 mechanisms may also explain age-related increases of ties of these signals and recent work suggesting a small

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 491

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

vascular contribution to BOLD signals in older adults activation is beneficial for performance in older adults.
during cognitive tasks32 encourage the continued use of Two studies have shown that using TMS to reduce activ-
this technique to study cognitive ageing. ity in either the left PFC during encoding or the right
PFC during retrieval reduces memory performance in
Compensation in the older brain younger adults but has less effect in older adults, pre-
An early idea in the literature was that older adults (that sumably because the unstimulated hemisphere can
is, those above the age of 65) might be able to engage support the function when the other is inactivated50,51.
some brain areas, particularly the frontal lobes, above Therefore, these studies all suggest that older adults can
the level seen in younger adults (in their twenties) to recruit higher levels of brain activity than young adults
compensate for impaired function elsewhere in the (often in the PFC) and that this additional activity can
brain22. In these early studies, it was noted that older aid the performance of the old adults who are best able
adults had more activity in the prefrontal cortex (PFC) to engage it.
during memory tasks relative to younger adults33–35, By contrast, other work has provided evidence that
which was thought to compensate for reduced activ- over-recruitment of brain activity does not necessar-
ity in visual processing regions22,36 (a phenomenon ily lead to better task performance. For example, some
recently termed the ‘posterior–anterior shift with age- researchers42,52,53 have suggested that when performance
ing’ (PASA)36). This PFC activity was often bilateral in is matched between age groups, over-recruitment
the older adults on tasks for which younger adults typi- reflects less efficient use of neural resources in the older
cally showed unilateral PFC activity, leading to the idea group, not compensation. In addition, more activation
that the increased bilaterality of PFC activation in older in old adults can sometimes be associated with poorer,
adults reflects a compensatory mechanism that can aid not better, performance54. Recent studies have reported
cognitive performance37. greater activity in the PFC during memory encoding 55
A compensatory interpretation is often invoked when or retrieval56 in older adults, both of which were cor-
older adults show more activity in a brain region than related with poorer memory. Similarly, higher activity
younger adults when they perform a task at the same in a distributed set of regions, including the PFC and
level as younger adults38, or when increased activity is parietal cortex, was found to be correlated with slower
positively correlated with performance in older adults and more variable reaction times on a set of visual tasks
but not in younger adults35,36,39–41. Several researchers in old adults compared to young adults57,58. To compli-
have suggested that compensatory mechanisms might cate matters further, some of these regions associated
still be involved, even if performance in older adults is with slower responses in older adults are very similar
impaired42. For example, increased activity in an older to the frontoparietal regions reported to support better
adult might not be associated with preserved perfor- inhibitory function in older adults45 (see above), suggest-
mance on a given task to the level seen in a young adult, ing that the association between activity in a given brain
but this performance might be even worse without the region and performance in older adults is task-specific,
over-recruitment. Thus, despite the continued atten- response-specific (for example, accuracy versus response
tion that this idea has received, it is still not clear exactly time) or both (FIG. 1).
which regions might act in a compensatory manner or Together, these results suggest that increased activity
under which conditions this might occur 43,44. in older relative to younger adults can be associated with
Several recent papers have provided further evidence better performance on some tasks, but that this addi-
in favour of the compensatory hypothesis. One paper tional activity is not always compensatory (in the sense
examined age differences in inhibition using a series that it is directly related to better task performance). In
of tasks that assessed the ability to inhibit prepotent some cases, over-recruitment of brain areas may reflect
responses45. Older adults displayed more activity in a set a greater demand on neural resources or less efficient use
of dorsal PFC and parietal regions, sometimes called the of them, and may or may not be related specifically to
dorsal attention network46,47, compared with younger individual differences in behaviour. One explanation for
adults. Importantly, activity in these attention-related this is the ‘partial compensation’ hypothesis55, whereby
regions correlated with better inhibition only in older over-recruitment of the right PFC during memory
adults. This result is consistent with the idea of a com- encoding may aid old adults in carrying out the encod-
pensatory mechanism whereby additional activity in ing task because of less effective use of the left PFC,
task-relevant regions increases the ability of older adults which would normally carry out this task21,59,60. However,
to carry out the task. Similarly, another experiment this additional right PFC activity during encoding can-
examined face perception48 and found a set of regions not compensate for a reduction in encoding effective-
in the right PFC and occipital cortex where increased ness of the left PFC, and so it does not provide a benefit
activity was associated with better face recognition in for subsequent memory of the encoded items. This is
older adults but not in a younger group. Furthermore, similar to the idea that over-recruitment might help cog-
in a task requiring attention to right and left visual nition in a general way but may not be related to perfor-
fields49, only old adults showed increased activity in the mance on a specific task42. Regardless, the papers cited
bilateral PFC that was positively correlated with bet- in this section indicate that one should be careful about
ter performance. Interestingly, studies using transcra- interpreting age-related increases in brain activity as
nial magnetic stimulation (TMS) have also provided being compensatory without sufficient evidence from
some support for the idea that increased bilateral PFC behaviour to support such an interpretation.

492 | JULY 2012 | VOLUME 13 www.nature.com/reviews/neuro

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

Another idea to explain compensatory activity is


the ‘compensation-related utilization of neural circuits
MFG hypothesis’ (CRUNCH)64. The idea of CRUNCH is that
more neural resources are recruited by older adults at
low levels of cognitive load — that is, when tasks are eas-
IFG ier — than by younger adults, who do not need them53.
At higher levels of load, this compensatory mechanism
is no longer effective, leading to equivalent or less acti-
vation in older adults relative to young adults. Data
consistent with this idea have been reported in the PFC
alone65 and in both the PFC and parietal cortex 66,67 dur-
IPL ing working memory tasks that varied in the number
ITC
of items that had to be kept in mind. In these studies,
VC
older adults had more activation at low levels of working
memory load, where performance was equivalent to that
of younger adults, but less activity and lower accuracy
Study 1: increased activity = slower response in older adults at higher loads. This kind of result also has been found
Study 2: increased activity = better accuracy in older adults during episodic memory tasks68: younger adults showed
Increased activity associated with slower response in study 1 recruitment of bilateral PFC during a difficult version
and better accuracy in study 2 of the memory task, whereas older adults showed acti-
Figure 1 | Increased brain activity in older adults may be associated with better or
vation of these areas for both easy and difficult ver-
worse task performance.  This figure summarizes the results of two studies that differ sions of the task. All of these studies are consistent with
in how increased brain activity in older adults was associatedNature
withReviews | Neuroscience
task performance. In CRUNCH (FIG. 2), which suggests that the relationship
one of these studies (study 1), the regions shown in green showed a correlation between between brain activity and cognitive load is ‘S’ shaped
more activity and slower reaction times on perceptual and working memory tasks in and plateaus at higher levels of load regardless of age.
older adults57. In another experiment (study 2), several brain regions (indicated in purple) The older adult curve would be shifted to the left rela-
showed a correlation between more activity and more accurate performance in a go/ tive to younger adults, such that older adults would have
no‑go task requiring inhibition of responses45. Note that some regions (coloured purple greater brain activity at lower levels of load and reach
and green) showed an association with better performance in study 2 and the opposite their plateau at levels where younger adults are still able
effect in study 1. This discrepancy highlights the complexity of trying to relate brain
to increase their brain activity. According to this hypoth-
activity in older adults to their behaviour, and indicates that the specific relationships
between regional brain activity and task performance in older adults depend on the task
esis, old adults engage neural resources, such as the PFC,
demands or on the behavioural measure that is assessed (or both). IFG, inferior frontal at lower task loads to compensate for less effective use
gyrus; IPL, inferior parietal lobe; ITC, inferior temporal cortex; MFG, middle frontal gyrus; of these resources, or perhaps because of degraded input
VC, visual cortex. to the PFC64, thus shifting the curve leftward. Although
this idea has considerable appeal, and may be able to
account for both the age-related increases and decreases
Potential explanations for compensatory activity. One of brain activity described in the literature, it is not clear
idea is that older adults shift from proactive strategies whether one would need to see recruitment of a unique
early in a decision process to reactive strategies that occur region in older adults in order to interpret this activity as
later. Support for this idea was reported in an experi- compensatory. That is, in older adults, the engagement
ment61 that found PFC activity in young adults occurred at lower load of the same region that is active in younger
during the early phase of memory retrieval trials and PFC adults at higher loads might reflect an increase in the
activity in older adults occurred during the later phase. ‘normal’ inter-individual variability in the brain–load
Another experiment examining task switching found that function that must exist even in young adults, rather
younger adults showed sustained PFC activity through- than compensation per se.
out the period in which they had to switch between
tasks, whereas older adults showed transient increases to Dedifferentiation
cues, indicating that a switch was required62. This pat- The concept of dedifferentiation was originally proposed
tern suggests that cognitive control is engaged differently to explain the increased correlations among behavioural
with ageing, and also supports the notion of a shift from measures in older adults69, but was adopted by neuro-
proactive control to a more reactive strategy that occurs imagers because it also seemed to characterize brain
in response to task demands. A similar proactive–reactive activity in older adults. Early examples included bilateral
age difference was reported for the medial temporal lobes prefrontal activity associated with abilities that typically
(MTLs) in older adults during a memory task63. Younger yield lateralized activity in younger adults37,70, more dif-
adults had more activity in these regions during prepara- fuse activation patterns71, and less selective activity in
tion for memory retrieval, whereas older adults showed task-relevant regions across a variety of tasks22,72,73. Like
more activity during retrieval. These studies suggest that the idea of compensation, dedifferentiation continues
a shift in the timing of resource engagement is required to to be a viable explanation for some age differences in
Cognitive control
The effortful use of cognitive
deal with the influence of age on proactive strategies that brain activity.
resources to guide, organize or make these less effective or accessible, and that compen- One way to investigate dedifferentiation is to com-
monitor behaviour. sation in ageing may have a temporal component to it. pare the patterns of activity across tasks to see whether

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 493

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

Implicit memory they are more similar (that is, less selective) in older of three different kinds of memory content: autobio-
Memory without conscious adults. This kind of result was found in an experiment graphical memory (personally relevant), episodic mem-
awareness, specifically a contrasting implicit memory for a sequence of repeated ory (not personally relevant but related to stimuli seen
change in a person’s behaviour visual stimuli to explicit memory for a list of words74. during the experiment) and semantic memory (world
(for example, faster reaction
times) owing to an
Young adults showed more activity in the hippocampus knowledge). These included activity in the MTLs during
experimental manipulation of for explicit learning and more activity in the striatum for autobiographical retrieval, in the dorsolateral PFC and
which they are not aware. implicit learning. Older adults showed equivalent acti- parietal regions during episodic retrieval, and in the left
vation in these regions during the two tasks. Another temporal cortex during retrieval of semantic memories.
Explicit memory
experiment75 also found that implicit memory in younger These patterns of activity were also seen in old adults but
Conscious retrieval of learned
information, such as recalling a
adults was accompanied by increased activity in the were less distinct for the autobiographical and episodic
list of words that has been striatum and decreased activity in the hippocampus, conditions, which is consistent with reported age dif-
previously studied. whereas older adults showed increases in both. Similarly, ferences in autobiographical and episodic memory but
older adults are reported to have less distinctive activ- maintained, or even increased, semantic memory with
ity in the visual cortex during perception and working age. Finally, less selective responses to specific catego-
memory tasks76,77. In both kinds of task, old adults had ries of visual stimuli have also been reported79, and are
less distinctive patterns of activity in the occipital cor- associated with measures of task switching and working
tex than young adults, which is consistent with dedif- memory in old adults80. Thus, all of these studies indi-
ferentiation. Interestingly, distinctiveness in PFC and cate that young adults have activation patterns that are
parietal regions was higher in old adults compared typically quite selective for the particular stimulus fea-
to young adults, which was interpreted as compensa- tures or task demands involved, whereas in older adults
tion. In another study 78, young adults were found to activation can be much less distinct, which is consist-
have unique patterns of activity during the retrieval ent with the idea of dedifferentiation across cognitive
processes. These studies further suggest that the loss of
selective brain responses may be a marker of a more
Younger adults general cognitive disruption.
Low-risk older adults Another way to assess dedifferentiation is to use adap-
High-risk older adults tation, which is a reduction in the response of a given
brain region (or regions) when a stimulus is presented
repeatedly relative to the first presentation81. Several
Change in brain activity

recent studies have used this method to look at selec-


tivity of brain responses in ageing. One assessed activ-
ity in the region of the brain that is most responsive to
faces, the fusiform face area (FFA)82, in response to faces
that were the same, that had been morphed by varying
amounts (that is, faces that were similar) or that were
different 83. Young adults showed more FFA activity dur-
ing the presentation of morphed faces than for the same
face shown repeatedly, indicating that the FFA treated
Cognitive load morphed faces as ‘different’ even though they were rela-
Figure 2 | The ‘compensation-related utilization of tively similar to each other. By contrast, the older adults
neural circuits hypothesis’.  Nature Reviews relating
The function | Neuroscience
the showed equivalent activity for same and morphed faces.
change in brain activity (measured by functional MRI Moreover, discrimination thresholds for distinguish-
during a task of interest) to levels of cognitive load is shown ing same from different faces were correlated with the
for young adults, old adults with a low risk of developing degree of adaptation in the FFA across younger and older
Alzheimer’s disease and old adults with a high risk of adults, indicating that this adaptation was important for
developing dementia. The function in low-risk older adults
behaviour. A similar study 48 assessed adaptation in the
would be shifted to the left relative to that seen in younger
adults. At relatively low levels of cognitive load this shift FFA during the presentation of faces that also varied
would result in higher activity in older relative to younger in viewpoint (right or left orientation). Young adults
adults (green shaded area). However, activity in older adults showed the least activity when the same face was seen
would reach its peak and level off while younger adults’ in the same viewpoint, more activity when the face or
activity is still increasing, so that at higher load levels the viewpoint changed, and the most activity when
there would be no age difference in activity or younger both the face and viewpoint changed. Older adults
adults would have higher activity (red shaded area). A showed no differences in activity in FFA across the
similar effect would be seen when high-risk older adults are conditions and performed worse than young adults
compared to low-risk older adults — higher activity in on a face-matching task involving changes in view-
high-risk groups relative to low-risk groups at low levels of
point. Adaptation in the auditory domain also has been
cognitive load (purple shaded area), with the reverse seen
at higher levels of load. This hypothesized set of examined84 and, like visual adaptation, is seen more
load-dependent functions could explain why studies prominently in younger than in older adults. These
have reported both under- and over-recruitment in experiments show that when adaptation is used to
older adults compared to young adults and in high- examine differentiation of responses in regions of cortex
risk older adults compared to low-risk older adults. that respond to specific features of stimuli, regardless of

494 | JULY 2012 | VOLUME 13 www.nature.com/reviews/neuro

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

modality, older age is associated with dedifferentiation of between the MTL and the PFC were stronger in the
responses that are relatively selective in younger adults. older adults compared to a younger group. The results of
Furthermore, this loss of selectivity may be associated these studies are reminiscent of the PASA effect involv-
with decrements in the ability to discriminate similarities ing more PFC activity in conjunction with less occipital
and/or differences among these stimuli. activity 36, and suggest that successful memory encoding
in older adults might be mediated by similar posterior-
Brain networks and functional connectivity to‑anterior shifts in the functional connectivity of
Cognitive neuroscientists are becoming increasingly memory-related regions in the MTLs.
interested in assessing the integrated activity among Functional connectivity has also been studied using
groups of brain regions as a way of defining brain net- attentional tasks. One recent study 94 showed that atten-
works (BOX 1). One way of doing this is to measure the tion to specific task-related cues was associated with
functional connectivity of a given brain region or set of activation of the dorsal PFC and parietal attention-
regions85,86. Several recent studies have looked at specific related regions in both younger and older adults, but
functional connections during a task and how these are functional connectivity of these regions was higher in
affected by age. One study examined changes in brain young adults than for older adults. Interestingly, increas-
activity during a working memory task with varying ing cue-related functional connectivity was associated
degrees of difficulty 87 and found that young adults had with more efficient performance on the task. In another
load-dependent increases of activity in the PFC. Older study 95 attention and expectancy were manipulated by
adults showed relatively high levels of PFC activity across predictively cueing which type of stimulus would be pre-
all load levels (consistent with CRUNCH, see above), and sented in a working memory task. When the cue indi-
weaker functional connectivity between the premotor cated that a picture of a face would be presented, young
cortex and a left dorsolateral PFC region. Another study adults showed greater functional connectivity between
addressed age differences in the resolution of interfer- the FFA and dorsal attention regions compared to older
ence during working memory for scenes using a delayed adults, which was consistent with their greater memory
match‑to‑sample task88. Interference was introduced by for predictively cued faces. Both of these studies sug-
presenting a face during the delay period and asking par- gest that weakened functional connectivity between PFC
ticipants to make a gender and age judgement about it. and parietal regions may explain the reduced ability of
Connectivity was measured between a brain region that older adults to attend to and make use of stimuli in the
responds preferentially to scenes (the parahippocampal environment. In general, the studies in this section sug-
place area (PPA)89) and a region of the PFC thought to gest that task-relevant functional connections between
be important for resolving interference. In young adults, specific brain regions can be disrupted with age and
the correlation between activity in the PPA and PFC was that these disruptions have a negative impact on task
disrupted when the face was presented but returned to performance.
pre-interruption levels after the face was removed, sug- Another recent trend in the neuroimaging literature
gesting that the resumption of PPA–PFC functional con- is to examine functional connectivity within particular
nectivity resolved the interference effect. Older adults large-scale brain networks, such as the default network
showed a similar disruption of PPA–PFC functional (DN), which is active when people are resting and
connectivity, but the effect persisted after the face was engaged in spontaneous thought 96–101. In young adults,
removed, suggesting a deficient ability to dynamically DN regions maintain strong functional interconnec-
Functional connectivity modulate network connectivity, which is consistent with tions during tasks requiring self-reference or theory of
A measure of how activity
within a network of brain
the poorer performance of the older adults on the task mind101,102 and also during the resting state103,104. Several
regions is correlated, or how in the presence of the interfering face. Both of these stud- studies have found that the reduction of DN activity dur-
activity in a particular brain ies highlight the importance of functional connectivity ing externally driven cognitive tasks is less pronounced
area is correlated with the rest between task-relevant regions and the influence of age on in old adults, relative to young adults58,105–112. Functional
of the brain.
these connections, which in turn might affect behaviour. connectivity of the DN also is reduced with age during
Delayed match-to-sample Age differences in functional connectivity during working memory tasks113 and during periods of rest105,114–
task episodic memory tasks have been studied using a verbal 116
. These reductions in task-related deactivation and
Presentation of a stimulus (the recognition task90. Old adults had reduced functional functional connectivity are seen in the two regions that
sample) followed by a delay of connectivity within a hippocampal–parietotemporal are thought to be the major nodes of the DN, the poste-
several seconds, and then
presentation of one or more
network relative to young adults, but increased con- rior cingulate cortex (PCC) and ventromedial PFC117, as
stimuli that have to be judged nectivity within a parahippocampal–frontal network. well as in other DN regions, such as the MTLs and infe-
as the same or different from This result was interpreted as evidence that older adults rior parietal lobes (FIG. 3). Interestingly, intrinsic connec-
the sample. compensate for hippocampal deficits by relying more on tivity during the resting state among nodes of the DN is
the parahippocampal cortex. Similar results have been related to the performance of older individuals on a vari-
Default network
(DN). A set of functionally reported in studies that measured brain activity during ety of cognitive tasks105,106,111,118. Given that DN modula-
connected brain regions that successful encoding of words91, scenes92 or objects93 by tion is associated with the degree of task difficulty and
is involved in spontaneous, comparing encoding activity for subsequently remem- performance119,120, a deficit in the ability to modulate DN
internally driven cognitive bered versus forgotten stimuli. In these studies, func- activity and functional connectivity with advancing age
processes and is more
active during periods of
tional connectivity during successful encoding between may be a mechanism for deficient resource allocation to
rest than during externally MTL regions and posterior regions, such as the occipi- the task at hand, accounting for some age differences in
driven tasks. tal cortex, was weaker in old adults, but connections cognitive performance108.

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 495

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

Factors influencing brain activity such as brain structure. It has been known for some
An important issue in understanding age differences in time that grey matter structures in the brain undergo
brain activity is how these are related to other factors that changes with ageing, such as reduced volume and thin-
are affected by ageing and that influence brain function, ning of the cortex, particularly in the frontal lobes121,122.

Box 1 | Measuring activity in brain networks


There is currently much interest in using neuroimaging to a
assess brain networks, and a number of methods have 1.0
emerged in recent years for identifying such networks and 0.8
measuring individual or group differences in network 0.6

Correlation z(r)
integration and activity. Most of these methods are based on
0.4
correlations or covariance between regional measures of
activity obtained with functional MRI (fMRI) (or some other 0.2
neuroimaging technique), and they range from relatively 0.0
simple assessments of correlations between the time courses –0.2
of two or more brain regions105 to more complex multivariate –0.4
approaches that assess brain-wide patterns of connectivity,
–0.6
such as independent component analysis or partial least
116
10 20 30 40 50 60 70 80 90 100
squares 198,199
. Another approach is that of graph theory ,
200
Age (years)
which uses the number of correlations that characterize
various regions to identify areas with large numbers
of connections (hubs) and to cluster together subgroups of b White matter
regions with strong connectivity inside larger collections
of areas. Some have attempted to design methods that can
assign temporal or functional causality, such as dynamic
causal modelling201 and Granger causality analysis202, but
these have been somewhat controversial203.
An example of the pair-wise correlation approach is shown
in the figure (panel a)105. In this study, the time course of
resting activity in the posterior cingulate cortex, a main node
of the default network (DN), was correlated with the time
course of activity in the other primary network node, the
.05 .001
medial prefrontal cortex (PFC) (these regions are shown as Bi > Mono
yellow circles on the brain image in panel a). The correlation
values are shown (panel a) for a group of older adults (green c Functional connectivity
dots) and younger adults (red dots). Not only are the
correlations lower in older adults relative to younger adults,
they are also reduced with age even in the older group (note
regression line shown in green). Studies assessing
whole-brain functional connectivity have shown age
differences in global DN functional connectivity58 (FIG. 3),
suggesting weaker network integration overall in older
adults. Functional connectivity in the DN (and other
networks) is weakened further in older adults with
dementia180,204. However, recent work has shown that some
imaging artefacts that are more common in older adults, such .0001 .01 .01 .0001
as the influence of motion in the scanner205, can weaken Mono > Bi Bi > Mono
functional connectivity, so issues such as this need to be
Nature Reviews | Neuroscience
examined further.
Structural connections appear to be important for at least some of the functional connectivity seen in brain networks206.
For example, the posterior cingulate cortex is a hub for structural connections207 as well as functional connections in the
DN117, and DN regions are strongly connected structurally208. In terms of ageing, older adults with better-maintained white
matter show stronger functional connectivity49. In addition, life experience can influence this relationship between
structure and function in older adults. Recently, it was shown that bilingualism was associated with better white matter
integrity as well as more distributed patterns of functional connectivity in older adults209. Bilingual older adults had better
white matter integrity in a number of tracts, including the corpus callosum, as measured with diffusion tensor imaging (see
panel b in the figure, which shows white matter tracts in green on two representative structural images, and the areas with
greater integrity in bilinguals (Bi) than in monolinguals (Mono) in red). The bilingual older adults also had stronger resting
functional connectivity between a region of the inferior PFC (circled in panel b) and posterior brain regions (red areas in
panel c, where brain images correspond to the slices seen in panel b) compared to age- and education-matched
monolinguals, whereas monolinguals had stronger functional connectivity within PFC areas (blue and green regions, panel
c). As older bilinguals typically show better cognitive control than monolinguals210, this finding suggests that
better-maintained white matter structure and more distributed functional connectivity support maintained cognitive
function in older age. Panel a is reproduced, with permission, from REF. 105 (2007) © Cell Press/Elsevier. Panels b and c are
adapted and reproduced, with permission, from REF. 209 (2011) © Society for Neuroscience.

496 | JULY 2012 | VOLUME 13 www.nature.com/reviews/neuro

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

Integrity of white matter, which is typically assessed with related to behaviour in older adults. For example, it has
diffusion tensor imaging (DTI)123, is also reduced in old been shown that stronger functional connectivity in a
adults compared to young adults124–126. In addition to network involving the inferior PFC was associated both
structural changes, age reductions in neurotransmitter with better integrity of the corpus callosum and faster
binding potential and receptor density have been found response times in older adults141 (BOX 1).
for both dopamine127–129 and serotonin130,131. Finally, the Reduced functional activation also has been asso-
incidence of dementing illnesses such as Alzheimer’s ciated with age differences in grey matter volumes142.
disease increases with age132, making the impact of risk One recent study 143 assessed the relationship between
factors for Alzheimer’s disease in healthy older adults an age reductions in grey matter volume of a region in the
area of interest. right middle frontal gyrus (MFG) and brain activity. In
young adults, larger right MFG volume was positively
Brain structure. There is a fairly extensive literature correlated with greater activity in the bilateral dorso-
on age differences and decline in grey and white mat- lateral PFC and inferior parietal cortex, both of which
ter structures in the brain, especially in the frontal have been implicated in memory retrieval144,145. In older
lobes133,134. A longitudinal study 135 showed that both the adults, right MFG volume was not positively correlated
hippocampal volume and integrity of white matter in with activity in any regions that showed correlations in
the corpus callosum were reduced in older adults and young adults, but was negatively correlated with activ-
correlated with declining memory performance. Some ity in several regions, including the parahippocampal
studies49,55 have found that age differences in activation cortex. Less activity in these regions predicted better
within the PFC were mediated by white matter integrity, memory in older adults, suggesting that older adults
such that more intact white matter was related to more with larger right MFG volume may be better able to
activation, but others have failed to find this effect 94,136. compensate for the effects of age on this region by mod-
Despite this inconsistency in results, the use of DTI to ifying activity in other brain regions to help memory
assess white matter integrity holds considerable prom- retrieval. Interestingly, in this case, the compensation, if
ise for the study of cognitive ageing, particularly as the that is what it is, appears to take the form of decreased
integrity of specific tracts has been shown to be related to activity in some regions, which may indicate suppression
speed of performance in older adults137 or to accuracy of of processes that would conflict with memory retrieval.
performance138–140. Of particular interest will be studies Another study 146 assessed the relation between
examining the relations among white matter integrity brain activity and grey matter volume in younger and
underlying specific functional networks, functional con- older adults across the whole brain. There was under-
nectivity in those networks and how these measures are recruitment of the occipital cortex during encoding
of face–name pairs in old adults compared to young
adults, which was mostly accounted for by atrophy in
SFG these regions. At retrieval, older adults over-recruited
a number of regions, including the dorsolateral PFC
and parietal cortex. This over-recruitment was elimi-
nated after accounting for volume loss in the PFC, but
IPL age differences remained in the parietal cortex after
accounting for the effect of age differences in volume.
These results suggest that structural age changes may
account for some, but not all, of the differences in brain
activity between older and younger adults. Perhaps more
important is the evidence that age differences in brain
PCC structure can influence the relationship between activity
in task-related brain regions and behaviour, indicating a
complex interplay between structure and function.
vmPFC
Dopamine. One of the most studied aspects of dopa-
mine is its role in reward. Current conceptions of how
reward is processed in the brain propose that a circuit
MTL of regions, including the ventral striatum and dopamin-
ergic cells in the ventral tegmental area, is necessary for
learning about and using rewards to guide behaviour 147.
Several studies have shown that there are age reductions
Figure 3 | The default network in young and older adults.  The regions making up the in striatal responses to learned reward148 and reward
default network (DN) are shown in red. The DN was defined inNature
younger adults and
Reviews older adults
| Neuroscience anticipation149. Only one study has directly examined
using a multivariate, whole-brain approach198 that identified regions where activity at rest was
correlated with activity in the posterior cingulate cortex (PCC, a major node of the DN). The the relation between functional activation during reward
DN included the inferior parietal lobes (IPLs), ventromedial prefrontal cortex (vmPFC), tasks and dopamine binding levels150. It showed that old
superior frontal gyrus (SFG) and the left medial temporal lobe (MTL). The green regions adults not only have less activity in the ventral striatum
represent a subset of DN areas, both medial and lateral, with weaker resting functional during reward anticipation but also show a weaker rela-
connectivity in older compared to young adults (BOX 1). tionship between this activity and dopamine levels in

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 497

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

the midbrain relative to younger adults, suggesting that adults (by blocking dopamine D1 receptors) resulted in
age-related dysfunction in this neurotransmitter system reduced activation in frontal and parietal regions dur-
could affect multiple everyday decisions that rely on ing a high-load working memory task to levels similar
reward processing. to those seen in older adults157. Performance was also
The role of dopamine in non-reward tasks also has lower in young adults after D1 blockade, although still
been examined. One study 151 assessed brain activity dur- better than that seen in older adults. However, when
ing a low-level working memory task and the influence a dopamine agonist was administered to old adults to
of a common polymorphism in the gene for catechol test the idea that boosting dopamine function would
O-methyltransferase (COMT) in young and old adults. induce similar brain activity to that observed in young
COMT is an enzyme that is thought to regulate dopa- adults when carrying out episodic memory tasks, an
mine levels in the PFC152, and the Val(158)Met polymor- enhancement, rather than a reduction, in age differ-
phism results in differing levels of available dopamine ences was seen158. Much more research is required before
in the brain. The Met variant is associated with lower any strong statements about the interactions among
dopamine-degrading activity relative to the Val variant, dopamine alterations and brain activity in ageing can
leading to greater dopamine levels. Individuals who were be made.
Met carriers showed no age difference in brain activ-
ity, whereas those with the Val allele showed a robust Risk factors for Alzheimer’s disease: apolipoprotein E.
age difference in left PFC activity. Older adults with the There is evidence that memory reductions can be seen at
Val allele, presumably those with lower dopamine lev- least 6 years before a diagnosis of Alzheimer’s disease159,
els, had higher activity than their younger counterparts. suggesting that pathology in memory-related regions is
These findings suggest that the Val(158)Met polymor- well advanced before diagnosis160. Therefore, it is impor-
phism influences the activity of brain regions within tant to assess the potential influence of Alzheimer’s dis-
working memory networks and that over-recruitment ease risk on studies of ‘normal’ ageing. The impact of the
of PFC activity in older adults can be linked to specific different alleles of the apolipoprotein E (APOE) gene has
gene effects. been examined in this context, as the presence of one
Another recent study 153 measured the binding or two ε4 alleles is a known risk factor for Alzheimer’s
potential of dopamine as an index of receptor density disease161. Some studies have reported greater activation
and related it to brain activity during working mem- in memory-related areas, notably the hippocampus, in
ory. Young adults had increased activity in frontal and healthy old adults who were ε4 carriers compared with
parietal regions in a high-load memory condition rela- non-carriers of the ε4 allele162–164, and even in young ε4
tive to low-load conditions, and these load-dependent carriers relative to non-carriers165, suggesting an increase
increases were greater in younger adults than in older in demand on these regions before the appearance
adults. Older adults showed reductions of dopamine of any symptoms of memory loss. However, a couple of
binding potential in the caudate nucleus and dorsolateral studies166,167 have found evidence of lower brain activ-
PFC, and when the contribution of these differences in ity in the hippocampus of aged ε4 allele carriers during
dopamine binding was accounted for, the age effects on memory tasks. These inconsistent findings regarding
frontal and parietal activity were eliminated or greatly brain activity in high-risk individuals compared to
reduced. These findings suggest that some of the age- their low-risk counterparts could be due to differences
related differences seen in brain activity during vary- in specific task demands, the influence of any number
ing cognitive loads (FIG. 2) may be due to alterations in of lifestyle or health factors, or where in the trajectory of
dopaminergic neurotransmission. longitudinal change one happens to measure brain
Unlike binding potential, dopamine synthesis capac- activity and cognition. For example, if a participant is
ity can be increased in old adults relative to younger on a trajectory towards eventual dementia, measuring
adults154, which could reflect an attempt to compen- brain activity early in this trajectory might reveal an
sate for the reduced receptor density. Greater synthesis over-recruitment of activity in a given region, whereas a
capacity in the caudate nucleus correlated with better later measurement might show under-recruitment. It is
verbal working memory performance and more PFC also possible that differential responses to cognitive load
activity during the task in old adults155. However, the could account for over- or under-recruitment in older
relationship between dopamine synthesis capacity and individuals with either high or low risk for Alzheimer’s
task-related modulation of activity in the posterior cin- disease (FIG. 2). It nevertheless seems clear that the APOE
gulate cortex (PCC) (a DN region), is disrupted in old genotype influences age-related changes in brain func-
adults156. These studies suggest that age differences in tion, and that the altered task-related brain activity in
dopamine synthesis capacity, as with binding potential, ε4 carriers may reflect the increased vulnerability of
influence functional activity in multiple brain circuits these individuals to Alzheimer’s disease pathology and
that are relevant for working memory performance, cognitive decline.
but whether these differences have a causal role in the Finally, it was recently shown that over-recruitment
reduced working memory performance in older adults of brain activity in older ε4 carriers is enhanced in those
is still unknown. with greater physical activity 168. Older adults with the ε4
Two studies have manipulated dopamine levels allele who engaged in more physical activity had greater
directly to assess the relationship between dopamine, memory-related activation in posterior temporal and
ageing and cognition. Dopamine depletion in young parietal regions than non‑ε4‑carriers or those with lower

498 | JULY 2012 | VOLUME 13 www.nature.com/reviews/neuro

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

physical activity. This result is particularly interesting as retrieval was correlated with memory in MCI. This
these areas are some of the first regions of the cortex shift was interpreted as a compensatory response to
to show metabolic deficits in early Alzheimer’s dis- dysfunction in the MTLs. The DN has also been studied
ease169–171. This work shows interesting influences of both in older adults with MCI, who show weaker functional
the APOE genotype and physical activity on memory- connectivity in this network compared to healthy elderly
related brain activation in cognitively intact but geneti- controls, consistent with studies showing that patients
cally at‑risk older adults, but it is not clear whether this with Alzheimer’s disease have less deactivation of and
increase is compensatory or protective against future weaker functional connectivity in the DN109,180. These
cognitive decline. effects of MCI have been found in the PCC181 and in its
connections to other regions182. In addition, DN func-
Risk factors for Alzheimer’s disease: mild cognitive tional connectivity is more affected in those individuals
impairment. Mild cognitive impairment (MCI) in older with MCI who later progress to dementia than in those
adults is another risk factor for Alzheimer’s disease, as a who remain stable over time183. Indeed, the weaker func-
relatively high proportion of older adults with MCI, par- tional connectivity of the DN in MCI and Alzheimer’s
ticularly those with amnestic symptoms, will progress to disease, in conjunction with the finding of amyloid
clinical dementia172. There is an extensive literature on deposition and other neuropathological changes in DN
functional and structural brain changes in MCI, much of regions, including the MTL160, has led to the suggestion
which has shown that individuals with MCI have greater that the DN is intimately involved in the neuropathol-
activation in the MTLs during memory tasks relative to ogy of Alzheimer’s disease184. Again, there is a similarity
healthy older controls173,174. Recent research has focused between the vulnerability of the DN in older adults with
on understanding what might underlie this over-activity. risk factors for Alzheimer’s disease (such as APOE and
For example, one recent study 175 examined subregions of MCI) and the vulnerability seen in healthy older adults
the hippocampus using high-resolution fMRI to explore relative to young adults, suggesting that DN activity
the CA3 region, which is thought to be involved in pat- and functional connectivity in older samples might be a
tern separation during memory. Participants with MCI useful marker for predicting cognitive decline.
showed over-recruitment of the CA3 region, but not
other regions, relative to controls, as well as impaired Influence of training on the ageing brain
pattern separation ability, which is consistent with the The influence of expertise on the adult brain has been
idea of a dysfunctional encoding mechanism result- demonstrated185,186 (see BOX 1 for an example of how a
ing from early neuropathological changes in this hip- lifelong experience can influence brain structure and
pocampal region. Interestingly, healthy older adults function), but less is known about how short-term
also show memory-related deficits in CA3 function behavioural training can affect brain activity in older
relative to younger adults176. Another study 177 found adults. This question is important because it has impli-
that over-recruitment of the hippocampus in MCI was cations for rehabilitating cognitive decline in older
related to cognitive load, such that it was only seen at people. A few neuroimaging studies have looked at this
lower levels of memory load during a paired-associates issue and their results are intriguing. One study pro-
task. At higher loads, activity in the hippocampus was vided training to older adults on a divided attention task
lower in the MCI group relative to controls, consistent in five 1‑hour sessions over a 2‑week interval and found
with CRUNCH (similar results have also been reported improved performance and reduced age differences in
in MCI for other brain regions178). These studies point brain activity that were apparent before training 187. PFC
towards specific processing deficits as well as an impair- activity that was greater in older adults before training
ment in the ability to respond to increases in cognitive was reduced to the level seen in younger adults after
demand as potential explanations for MTL over-recruit- training, presumably because the practice on the two
ment in MCI. This work also highlights the similarities tasks had reduced the effort required to carry them out
between age differences in healthy older versus younger simultaneously, reducing the need for PFC-mediated
adults and differences between MCI and healthy older cognitive control. Similarly, a reduction in the ampli-
individuals (for example, both can be characterized by tude of an electrophysiological-evoked response during
CRUNCH and involve over-recruitment of brain activ- a working memory task was reported in older adults
ity). This similarity suggests a continuum of effects due after 10 hours of perceptual discrimination training,
to age and neuropathological brain changes, perhaps and this reduction predicted the increase in accuracy
because both ageing and risk of dementia can affect cog- on the working memory task that was achieved after
nition in a general way that impairs the ability to respond training 188.
to increasing cognitive demand. Increased activity in older adults after episodic mem-
Other recent work has emphasized how MCI affects ory training has been reported in a study 189 that scanned
larger scale brain networks. One such study showed that young and old adults during encoding of words to assess
healthy older adults use a network of regions, including baseline age differences. The older adults then under-
the MTLs, for successful encoding 179. Although partici- went two training sessions (for a total of 2.5 hours), in
pants with MCI showed engagement of this network, which they were trained on the use of three different
activity in it was not associated with memory perfor- learning strategies and then allowed to use the strategy
mance; instead activity in a network involving anterior of their choice to learn lists of words. A post-training
temporal regions thought to be involved in semantic fMRI session was then carried out in the older group.

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 499

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

Box 2 | Assessing behavioural and brain trajectories over the lifespan


An example of how lifespan studies could a Age differences in BOLD variability
add to our knowledge of brain–behaviour
interactions can be found in the study of
variability. It is well known that measures
of behavioural performance, such as
response times, are variable both between
and within individuals120,211,212 and that
behavioural variability is higher in children
and older adults relative to younger
adults213–216. In ageing, behavioural
variability also can serve as a marker for
cognitive decline217,218 and increases before
death219. However, the relation between
behavioural variability and variability in
brain activity has not been extensively
examined, although evidence indicates
that variability in ongoing activity is
important for the expression of evoked
patterns of activity220. The use of functional
MRI to study brain function has relied
primarily on assessing average brain
activation patterns. Nevertheless, brain
activity is inherently variable, and several Old > Young Young > Old
lines of research have shown that our
ability to understand important aspects of
brain function is enhanced by considering
b Hypothetical relation between behavioural and BOLD variability
the variability of brain signals221–223. In
particular, networks that are more variable
may be more robust to disruption and may
explore more neural states, thus enhancing
learning and promoting optimal
performance221,222,224,225. Brain
Recent studies have shown that there are
developmental increases in variability and
Level of variability

complexity of brain activity, from childhood


to the young adult ages, along with
increased accuracy and stability of task
performance226. In addition, a recent study
assessed variability of the blood-oxygen-
level-dependent (BOLD) signal with age,
using the standard deviation of activity in
all brain voxels, and found that the majority Behaviour
of regions with age differences had less
variability in the older group227 (see the
figure, blue regions in panel a). In addition,
lower BOLD variability in these regions was
associated with slower and more variable Age
response times on cognitive tasks57. Thus,
this accumulating evidence suggests that behavioural variability has a U‑shaped function over the lifespan216, with larger
variability in children and older adults compared to young adults (see the figure, panel b), Nature
whereasReviews
variability of brain
| Neuroscience
activity shows the opposite trend (inverted U shape). Lifespan studies examining this kind of question using the same
behavioural and imaging paradigms from children up to older adults would shed much light on how developmental
changes in brain function can affect behaviour. Panel a is adapted and reproduced, with permission, from REF. 227 ©
(2010) Society for Neuroscience. 

The use of the strategies by the older adults during the degree of improvement in memory after training,
the encoding condition at the second fMRI session suggesting a direct role for training in influencing both
coincided with an elimination of the pre-training age brain function and behaviour. These training studies
differences in word memory. Training also increased suggest that increased brain activity after even limited
older adults’ brain activity in the left frontal and tem- training could be a result of the adoption of a different
poral regions that have been previously associated strategy, whereas decreased brain activity after training
with verbal processing and successful encoding 70,190,191. is more likely to be due to a practice-related increase of
These increases of brain activity were correlated with efficiency on a task187.

500 | JULY 2012 | VOLUME 13 www.nature.com/reviews/neuro

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

worse task performance, in which case it is defined as


Brain structure ‘unsuccessful compensation’. In this model, determin-
White Grey ing the relationship between the engagement of addi-
matter matter tional neural resources and task performance is critical
Vasculature for determining whether the compensation has been
Neuro- successful or not. Adopting terms such as unsuccess-
transmitters ful or partial compensation, to make a clear distinction
between these phenomena and successful compensa-
tion in which over-recruitment is clearly linked to bet-
ter performance, may help to remove discrepancies in
the literature and clarify the compensatory role of vari-
Other factors Cognition ous regions. In addition, others have suggested23 that
Education Life Semantics Spatial the term ‘compensation’ should be used only for those
experience instances in which old adults recruit brain activity that
Stress Ageing Language is not seen in younger adults, and the engagement of this
area or areas is directly correlated with better perfor-
Genes Exercise Memory Attention mance only in the older adults and not in young adults
(that is, there is a unique pattern of neural activity that
Diet Perception supports task performance in an age-specific manner).
Another initiative that would be welcome in this field
is the use of lifespan studies to identify the changes that
occur, both in cognitive processes and the brain mecha-
nisms underlying them, from childhood to old age. Such
Brain function developmental changes could take a variety of forms,
including both linear and nonlinear changes. There is
Activity Variability recent evidence suggesting that some behavioural and
brain trends in development might take different forms
Functional (BOX 2), indicating that much could be learned about
connectivity the links between brain and behaviour using a com-
prehensive lifespan approach. In addition, longitudinal
Figure 4 | A hypothetical model of the various dimensions that can interact with studies will be important for understanding brain age-
ageing.  The model is intended to show the interplay among a wide
Nature array| Neuroscience
Reviews of physical and ing. Although cross-sectional studies are easier to carry
behavioural aspects (some of which are discussed in this Review) and the ageing process.
out, and have contributed most of what we know to date
The arrows are bidirectional to indicate that the influence can potentially arise from
these factors on the ageing process, or vice versa. For example, genetic factors could about ageing of the brain, they are vulnerable to cohort
influence how an individual ages, and ageing can enhance the effects of genes on effects, and longitudinal studies are necessary for identi-
specific behaviours. There are other factors that could be included here, such as risk fying the effects of ageing within individuals. There have
factors for vascular disease or dementia, but this incomplete list gives a sense of how been a few longitudinal studies of brain function in older
complex the study of ageing is and how difficult it would be to comprehensively assess adults that have shown decreased task-related activity
these variables in a single experiment. over time193, both decreases and increases depending on
the specific brain region and cognitive demands194,195,
and a greater decline of activity in older individuals with
Conclusions risk factors for Alzheimer’s disease196. With so few data
There are a number of issues in the ageing literature points it is difficult to come to any strong conclusions
that have yet to be resolved satisfactorily, including that about change over time, highlighting the need for these
of compensatory brain activity. For example, over- kinds of studies.
recruitment of brain activity in older adults has been Finally, it is clear that ageing is influenced by a large
interpreted as compensation both when there was a number of factors that vary from individual to indi-
positive correlation between activity and behaviour 39, vidual, including genetics and life experiences (FIG. 4).
and when this correlation was negative146. Although it Although it is probably impossible to account for all of
seems unlikely that both positive and negative correla- these factors in a single study, the current trend is to
tions could be compensatory, perhaps a more careful include an assessment of multiple influencing factors and
and consistent definition of what is ‘compensatory’ is multiple measures of brain structure and function, as the
needed. One model192 defines three different types of experiments reviewed here can attest. Publicly accessible
compensation. When there is a mismatch between avail- databases, such as the Alzheimer Disease Neuroimaging
able cognitive resources and current task demands, this Initiative (ADNI)197, that contain information on a large
leads to the recruitment of additional neural resources, number of individuals collected across multiple labora-
reflected in increased brain activity. At this point, with- tories will aid greatly in this effort. Sharing of data and
out a link to behaviour, the over-recruitment is called meta-analyses will allow for larger scale examinations
‘attempted compensation’. The increase in brain activ- of ageing effects than is possible with data from a single
ity may be associated with better task performance, in laboratory and will ultimately add to our knowledge in
which case it is defined as ‘successful compensation’, or a substantial way.

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 501

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

1. Tulving, E. Elements of Episodic Memory (Oxford Univ. 26. Aizenstein, H. J. et al. The BOLD hemodynamic 49. Davis, S. W., Kragel, J. E., Madden, D. J. & Cabeza, R.
Press, 1983). response in healthy aging. J. Cogn. Neurosci. 16, The architecture of cross-hemispheric communication
2. Craik, F. I. M. & Bosman, E. A. in Gerontechnology: 786–793 (2004). in the aging brain: linking behavior to functional and
Proceedings of the First International Conference on 27. Huettel, S. A., Singerman, J. D. & McCarthy, G. The structural connectivity. Cereb. Cortex 22, 232–242
Technology and Aging (eds Bouma, H. & Graafmans, J.) effects of aging upon the hemodynamic response (2012).
79–92 (IOS Press, 1992). measured by functional MRI. Neuroimage 13, 50. Rossi, S. et al. Age-related functional changes of
3. Balota, D. A., Dolan, P. O. & Duchek, J. M. in The 161–175 (2001). prefrontal cortex in long-term memory: a repetitive
Oxford Handbook of Memory (eds Tulving, E. & 28. D’Esposito, M., Zarahn, E., Aguirre, G. K. & Rypma, B. transcranial magnetic stimulation study. J. Neurosci.
Craik, F.) 395–410 (Oxford Univ. Press, 2000). The effect of normal aging on the coupling of neural 24, 7939–7944 (2004).
4. Zacks, R. T., Hasher, L. & Li, K. Z. H. in The Handbook activity to the BOLD hemodynamic response. 51. Manenti, R., Cotelli, M. & Miniussi, C. Successful
of Aging and Cognition (eds Craik, F. I. M. & Salthouse, Neuroimage 10, 6–14 (1999). physiological aging and episodic memory: a brain
T. A.) 200–230 (Erlbaum, 2000). 29. Buckner, R. L., Snyder, A. Z., Sanders, A. L., stimulation study. Behav. Brain Res. 216, 153–158
5. Connelly, S. L., Hasher, L. & Zacks, R. T. Age and Raichle, M. E. & Morris, J. C. Functional brain imaging (2011).
reading: the impact of distraction. Psychol. Aging 6, of young, nondemented, and demented older adults. 52. Morcom, A. M., Li, J. & Rugg, M. D. Age effects on the
533–541 (1991). J. Cogn. Neurosci. 12, 24–34 (2000). neural correlates of episodic retrieval: Increased
6. Allen, P. A., Makken, D. J., Groth, K. E. & Crozier, L. C. 30. Hillary, F. G. & Biswal, B. The influence of cortical recruitment with matched performance. Cereb.
Impact of age, redundancy, and perceptual noise on neuropathology on the FMRI signal: a measurement Cortex 17, 2491–2506 (2007).
visual search. J. Gerontol. 47, P69–P74 (1992). of brain or vein? Clin. Neuropsychol. 21, 58–72 53. Rypma, B., Eldreth, D. A. & Rebbechi, D. Age-related
7. Madden, D. J. Adult age differences in attentional (2007). differences in activation-performance relations in
selectivity and capacity. Eur. J. Cogn. Psychol. 2, 31. D’Esposito, M., Deouell, L. Y. & Gazzaley, A. delayed-response tasks: a multiple component
229–252 (1990). Alterations in the BOLD fMRI signal with ageing and analysis. Cortex 43, 65–76 (2007).
8. Anderson, N. D., Craik, F. I. M. & Naveh-Benjamin, M. disease: a challenge for neuroimaging. Nature Rev. 54. Stevens, W. D., Hasher, L., Chiew, K. & Grady, C. L.
The attentional demands of encoding and retrieval in Neurosci. 4, 863–872 (2003). A neural mechanism underlying memory failure in
younger and older adults: I. Evidence from divided 32. Kannurpatti, S. S., Motes, M. A., Rypma, B. & older adults. J. Neurosci. 28, 12820–12824 (2008).
attention costs. Psychol. Aging 13, 405–423 (1998). Biswal, B. B. Neural and vascular variability and the 55. de Chastelaine, M., Wang, T. H., Minton, B., Muftuler,
9. Kramer, A. F., Hahn, S. & Gopher, D. Task coordination fMRI-BOLD response in normal aging. Magn. Reson. L. T. & Rugg, M. D. The effects of age, memory
and aging: explorations of executive control processes Imag. 28, 466–476 (2010). performance, and callosal integrity on the neural
in the task switching paradigm. Acta Psychol. (Amst.) 33. Cabeza, R. et al. Age-related differences in neural correlates of successful associative encoding. Cereb.
101, 339–378 (1999). activity during memory encoding and retrieval: a Cortex 21, 2166–2176 (2011).
10. Cepeda, N. J., Kramer, A. F. & Gonzalez de Sather, positron emission tomography study. J. Neurosci. 17, 56. Persson, J., Kalpouzos, G., Nilsson, L. G., Ryberg, M.
J. C. Changes in executive control across the life span: 391–400 (1997). & Nyberg, L. Preserved hippocampus activation in
examination of task-switching performance. Dev. 34. Madden, D. J. et al. Adult age differences in the normal aging as revealed by fMRI. Hippocampus 21,
Psychol. 37, 715–730 (2001). functional neuroanatomy of verbal recognition 753–766 (2011).
11. Hasher, L. & Zacks, R. T. in The Psychology of Learning memory. Hum. Brain Map. 7, 115–135 (1999). 57. Garrett, D. D., Kovacevic, N., McIntosh, A. R. & Grady,
and Motivation Vol. 22 (ed. Bower, G. H.) 193–225 35. Reuter-Lorenz, P. A. et al. Age differences in the C. L. The importance of being variable. J. Neurosci.
(Academic Press, 1988). frontal lateralization of verbal and spatial working 31, 4496–4503 (2011).
12. Healey, M. K., Campbell, K. L. & Hasher, L. in Progress memory revealed by PET. J. Cogn. Neurosci. 12, This is one of a series of papers showing that the
in Brain Research Vol. 169: The Essence of Memory 174–187 (2000). variability of the fMRI signal is lower in older adults,
(eds Sossin, W., Lacaille, J. C., Castellucci, V. F. & 36. Davis, S. W., Dennis, N. A., Daselaar, S. M., Fleck, M. S. compared to younger adults, and that less-variable
Belleville, S.) 353–363 (Elsevier, 2008). & Cabeza, R. Que PASA? The posterior-anterior shift brain signals are associated with greater behavioural
This review describes work characterizing age in aging. Cereb. Cortex 18, 1201–1209 (2008). variability on cognitive tasks. This is a novel approach
differences in inhibition. When younger and older 37. Cabeza, R. Hemispheric asymmetry reduction in older to the study of brain function with fMRI.
adults are presented with a task in the presence of adults: the HAROLD model. Psychol. Aging 17, 58. Grady, C. L. et al. A multivariate analysis of age-
distraction (and told to ignore the distracting 85–100 (2002). related differences in default mode and task-positive
information), older adults have better memory for This paper presents a model to explain networks across multiple cognitive domains. Cereb.
the distracting material when tested subsequently. over-recruitment of the bilateral PFC in older Cortex 20, 1432–1447 (2010).
This effect is thought to be due to an age-related adults and argues that this over-recruitment is 59. Otten, L. J. & Rugg, M. D. Task-dependency of the
reduction in inhibitory effectiveness. compensatory. This idea of bilaterality in ageing is neural correlates of episodic encoding as measured by
13. Salthouse, T. A. The processing-speed theory of adult still being assessed in the literature. fMRI. Cereb. Cortex 11, 1150–1160 (2001).
age differences in cognition. Psychol. Rev. 103, 38. Cabeza, R., Anderson, N. D., Locantore, J. K. & 60. Wagner, A. D. et al. Building memories: remembering
403–428 (1996). McIntosh, A. R. Aging gracefully: compensatory brain and forgetting of verbal experiences as predicted by
14. Craik, F. I. M. & Jennings, J. M. in The Handbook of activity in high-performing older adults. Neuroimage brain activity. Science 281, 1188–1191 (1998).
Aging and Cognition (eds Craik, F. I. M. & 17, 1394–1402 (2002). 61. Velanova, K., Lustig, C., Jacoby, L. L. & Buckner, R. L.
Salthouse, T. A.) 51–110 (Lawrence Erlbaum, 1992). 39. McIntosh, A. R. et al. Recruitment of unique neural Evidence for frontally mediated controlled processing
15. Laver, G. D. Adult aging effects on semantic and systems to support visual memory in normal aging. differences in older adults. Cereb. Cortex 17,
episodic priming in word recognition. Psychol. Aging Curr. Biol. 9, 1275–1278 (1999). 1033–1046 (2007).
24, 28–39 (2009). 40. Della-Maggiore, V. et al. Corticolimbic interactions 62. Jimura, K. & Braver, T. S. Age-related shifts in brain
16. Carstensen, L. L., Fung, H. F. & Charles, S. T. associated with performance on a short-term memory activity dynamics during task switching. Cereb. Cortex
Socioemotional selectivity theory and the regulation of task are modified by age. J. Neurosci. 20, 8410–8416 20, 1420–1431 (2010).
emotion in the second half of life. Motiv. Emot. 27, (2000). 63. Dew, I. T., Buchler, N., Dobbins, I. G. & Cabeza, R.
103–123 (2003). 41. Grady, C. L., McIntosh, A. R. & Craik, F. Task-related Where Is ELSA? The early to late shift in aging. Cereb.
17. Carstensen, L. L. et al. Emotional experience improves activity in prefrontal cortex and its relation to Cortex 23 Nov 2011 (doi:10.1093/cercor/bhr334).
with age: evidence based on over 10 years of experience recognition memory performance in young and old 64. Reuter-Lorenz, P. A. & Cappell, K. A. Neurocognitive
sampling. Psychol. Aging 26, 21–33 (2011). adults. Neuropsychologia 43, 1466–1481 (2005). aging and the compensation hypothesis. Curr. Direct.
18. Rahhal, T. A., May, C. P. & Hasher, L. Truth and 42. Zarahn, E., Rakitin, B., Abela, D., Flynn, J. & Stern, Y. Psychol. Sci. 17, 177–182 (2008).
character: sources that older adults can remember. Age-related changes in brain activation during a This paper presents the CRUNCH model of brain
Psychol. Sci. 13, 101–105 (2002). delayed item recognition task. Neurobiol. Aging 28, function and ageing, and suggests a mechanism to
19. Spreng, R. N., Wojtowicz, M. & Grady, C. L. Reliable 784–798 (2007). explain both under and over-recruitment of brain
differences in brain activity between young and old 43. Rajah, M. N. & D’Esposito, M. Region-specific changes activity in older adults.
adults: a quantitative meta-analysis across multiple in prefrontal function with age: a review of PET and 65. Mattay, V. S. et al. Neurophysiological correlates of
cognitive domains. Neurosci. Biobehav. Rev. 34, fMRI studies on working and episodic memory. Brain age-related changes in working memory capacity.
1178–1194 (2010). 128, 1964–1983 (2005). Neurosci. Lett. 392, 32–37 (2006).
20. Eyler, L. T., Sherzai, A., Kaup, A. R. & Jeste, D. V. 44. Greenwood, P. M. Functional plasticity in cognitive 66. Cappell, K. A., Gmeindl, L. & Reuter-Lorenz, P. A. Age
A review of functional brain imaging correlates of aging: review and hypothesis. Neuropsychology 21, differences in prefontal recruitment during verbal
successful cognitive aging. Biol. Psychiatry 70, 657–673 (2007). working memory maintenance depend on memory
115–122 (2011). 45. Vallesi, A., McIntosh, A. R. & Stuss, D. T. load. Cortex 46, 462–473 (2010).
21. Grady, C. L. et al. Age-related reductions in human Overrecruitment in the aging brain as a function of 67. Schneider-Garces, N. J. et al. Span, CRUNCH, and
recognition memory due to impaired encoding. task demands: evidence for a compensatory view. beyond: working memory capacity and the aging
Science 269, 218–221 (1995). J. Cogn. Neurosci. 23, 801–815 (2011). brain. J. Cogn. Neurosci. 22, 655–669 (2010).
22. Grady, C. L. et al. Age-related changes in cortical 46. Vincent, J. L., Kahn, I., Snyder, A. Z., Raichle, M. E. & 68. Spaniol, J. & Grady, C. Aging and the neural correlates
blood flow activation during visual processing of faces Buckner, R. L. Evidence for a frontoparietal control of source memory: over-recruitment and functional
and location. J. Neurosci. 14, 1450–1462 (1994). system revealed by intrinsic functional connectivity. reorganization. Neurobiol. Aging 33, 425.e3–425.e18
23. Grady, C. L. Cognitive neuroscience of aging. Ann. NY J. Neurophysiol. 100, 3328–3342 (2008). (2012).
Acad. Sci. 1124, 127–144 (2008). 47. Corbetta, M., Patel, G. & Shulman, G. L. The 69. Lindenberger, U. & Baltes, P. B. Sensory functioning
24. Logothetis, N. K., Pauls, J., Augath, M., Trinath, T. & reorienting system of the human brain: from and intelligence in old age: a strong connection.
Oeltermann, A. Neurophysiological investigation of environment to theory of mind. Neuron 58, 306–324 Psychol. Aging 9, 339–355 (1994).
the basis of the fMRI signal. Nature 412, 150–157 (2008). 70. Logan, J. M., Sanders, A. L., Snyder, A. Z., Morris,
(2001). 48. Lee, Y., Grady, C. L., Habak, C., Wilson, H. R. & J. C. & Buckner, R. L. Under-recruitment and
25. Mukamel, R. et al. Coupling between neuronal firing, Moscovitch, M. Face processing changes in normal nonselective recruitment: dissociable neural
field potentials, and FMRI in human auditory cortex. aging revealed by fMRI adaptation. J. Cogn. Neurosci. mechanisms associated with aging. Neuron 33,
Science 309, 951–954 (2005). 23, 3433–3447 (2011). 827–840 (2002).

502 | JULY 2012 | VOLUME 13 www.nature.com/reviews/neuro

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

71. Madden, D. J. et al. Aging and recognition memory: 94. Madden, D. J. et al. Adult age differences in functional 118. Wang, L. et al. Intrinsic connectivity between the
changes in regional cerebral blood flow associated connectivity during executive control. Neuroimage 52, hippocampus and posteromedial cortex predicts
with components of reaction time distributions. 643–657 (2010). memory performance in cognitively intact older
J. Cogn. Neurosci. 11, 511–520 (1999). 95. Bollinger, J., Rubens, M. T., Masangkay, E., Kalkstein, J. individuals. Neuroimage 51, 910–917 (2010).
72. Grady, C. L. Age-related differences in face processing: & Gazzaley, A. An expectation-based memory deficit in 119. McKiernan, K. A., Kaufman, J. N., Kucera-Thompson, J.
a meta-analysis of three functional neuroimaging aging. Neuropsychologia 49, 1466–1475 (2011). & Binder, J. R. A parametric manipulation of factors
experiments. Canad. J. Exp. Psychol. 56, 208–220 96. Gusnard, D. A., Akbudak, E., Shulman, G. L. & affecting task-induced deactivation in functional
(2002). Raichle, M. E. Medial prefrontal cortex and self- neuroimaging. J. Cogn. Neurosci. 15, 394–408
73. Townsend, J., Adamo, M. & Haist, F. Changing referential mental activity: Relation to a default mode (2003).
channels: an fMRI study of aging and cross-modal of brain function. Proc. Natl Acad. Sci. USA 98, 120. Kelly, A. M., Uddin, L. Q., Biswal, B. B., Castellanos,
attention shifts. Neuroimage 31, 1682–1692 (2006). 4259–4264 (2001). F. X. & Milham, M. P. Competition between functional
74. Dennis, N. A. & Cabeza, R. Age-related 97. Raichle, M. E. et al. A default mode of brain function. brain networks mediates behavioral variability.
dedifferentiation of learning systems: an fMRI study of Proc. Natl Acad. Sci. USA 98, 676–682 (2001). Neuroimage 39, 527–537 (2008).
implicit and explicit learning. Neurobiol. Aging 33, 98. Shulman, G. L. et al. Common blood flow changes 121. Raz, N. et al. Selective aging of human cerebral cortex
2318.e17–2318.e30 (2011). across visual tasks: decreases in cerebral cortex. observed in vivo: differential vulnerability of the
75. Rieckmann, A., Fischer, H. & Backman, L. Activation in J. Cogn. Neurosci. 9, 648–663 (1997). prefrontal gray matter. Cereb. Cortex 7, 268–282
striatum and medial temporal lobe during sequence 99. Buckner, R. L. & Carroll, D. C. Self-projection and the (1997).
learning in younger and older adults: relations to brain. Trends Cogn. Sci. 11, 49–57 (2007). 122. Fjell, A. M. et al. High consistency of regional cortical
performance. Neuroimage 50, 1303–1312 (2010). 100. Spreng, R. N., Mar, R. A. & Kim, A. S. N. The common thinning in aging across multiple samples. Cereb.
76. Carp, J., Gmeindl, L. & Reuter-Lorenz, P. A. Age neural basis of autobiographical memory, prospection, Cortex 19, 2001–2012 (2009).
differences in the neural representation of working navigation, theory of mind, and the default mode: a 123. Johansen-Berg, H. & Behrens, T. (eds) Diffusion MRI:
memory revealed by multi-voxel pattern analysis. quantitative meta-analysis. J. Cogn. Neurosci. 21, From Quantitative Measurement to In Vivo
Front. Hum. Neurosci. 4, 217 (2010). 489–510 (2009). Neuroanatomy (Academic Press, 2009).
77. Carp, J., Park, J., Polk, T. A. & Park, D. C. Age 101. Spreng, R. N. & Grady, C. L. Patterns of brain activity 124. Moseley, M. Diffusion tensor imaging and aging — a
differences in neural distinctiveness revealed by multi- supporting autobiographical memory, prospection and review. NMR Biomed. 15, 553–560 (2002).
voxel pattern analysis. Neuroimage 56, 736–743 theory-of‑mind, and their relationship to the default 125. Sullivan, E. V. & Pfefferbaum, A. Diffusion tensor
(2011). mode network. J. Cogn. Neurosci. 22, 1112–1123 imaging and aging. Neurosci. Biobehav. Rev. 30,
78. St‑Laurent, M., Abdi, H., Burianov, H. & Grady, C. L. (2010). 749–761 (2006).
Influence of aging on the neural correlates of 102. Grigg, O. & Grady, C. L. The default network and 126. Head, D. et al. Differential vulnerability of anterior
autobiographical, episodic, and semantic memory processing of personally relevant information: white matter in nondemented aging with minimal
retrieval. J. Cogn. Neurosci. 23, 4150–4163 (2011). converging evidence from task-related modulations acceleration in dementia of the Alzheimer type:
79. Park, D. C. et al. Aging reduces neural specialization in and functional connectivity. Neuropsychologia 48, evidence from diffusion tensor imaging. Cereb. Cortex
ventral visual cortex. Proc. Natl Acad. Sci. USA 101, 3815–3823 (2010). 14, 410–423 (2004).
13091–13095 (2004). 103. Buckner, R. L., Andrews-Hanna, J. R. & Schacter, D. L. 127. Kish, S. J., Zhong, X. H., Hornykiewicz, O. & Haycock,
80. Park, J., Carp, J., Hebrank, A., Park, D. C. & Polk, T. A. The brain’s default network: anatomy, function, and J. W. Striatal 3,4‑dihydroxyphenylalanine
Neural specificity predicts fluid processing ability in relevance to disease. Ann. NY Acad. Sci. 1124, 1–38 decarboxylase in aging: disparity between
older adults. J. Neurosci. 30, 9253–9259 (2010). (2008). postmortem and positron emission tomography
81. Grill-Spector, K., Henson, R. & Martin, A. Repetition This review paper presents a comprehensive studies? Ann. Neurol. 38, 260–264 (1995).
and the brain: neural models of stimulus-specific summary of the DN and the potential relevance of 128. Rinne, J. O., Lonnberg, P. & Marjamaki, P. Age-
effects. Trends Cogn. Sci. 10, 14–23 (2006). disrupted network function to dementia. dependent decline in human brain dopamine D1 and
82. Kanwisher, N. & Yovel, G. The fusiform face area: a 104. Fox, M. D. et al. The human brain is intrinsically D2 receptors. Brain Res. 508, 349–352 (1990).
cortical region specialized for the perception of faces. organized into dynamic, anticorrelated functional 129. Walker, L. C. et al. The neural basis of memory decline
Phil. Trans. R. Soc. B 361, 2109–2128 (2006). networks. Proc. Natl Acad. Sci. USA 102, 9673–9678 in aged monkeys. Neurobiol. Aging 9, 657–666
83. Goh, J. O., Suzuki, A. & Park, D. C. Reduced neural (2005). (1988).
selectivity increases fMRI adaptation with age during 105. Andrews-Hanna, J. R. et al. Disruption of large-scale 130. Meltzer, C. C. et al. Reduced binding of [18F]
face discrimination. Neuroimage 51, 336–344 brain systems in advanced aging. Neuron 56, altanserin to serotonin type 2A receptors in aging:
(2010). 924–935 (2007). persistence of effect after partial volume correction.
84. Grady, C. L., Charlton, R., He, Y. & Alain, C. Age 106. Damoiseaux, J. S. et al. Reduced resting-state brain Brain Res. 813, 167–171 (1998).
differences in FMRI adaptation for sound identity and activity in the “default network” in normal aging. 131. Moller, M., Jakobsen, S. & Gjedde, A. Parametric and
location. Front. Hum. Neurosci. 5, 24 (2011). Cereb. Cortex 18, 1856–1864 (2008). regional maps of free serotonin 5HT1A receptor sites
85. McIntosh, A. R. Mapping cognition to the brain 107. Esposito, F. et al. Independent component model of in human brain as function of age in healthy humans.
through neural interactions. Memory 7, 523–548 the default-mode brain function: combining individual- Neuropsychopharmacology 32, 1707–1714 (2007).
(1999). level and population-level analyses in resting-state 132. Brookmeyer, R., Gray, S. & Kawas, C. Projections of
86. Horwitz, B. in Visuomotor Coordination (eds Ewert, fMRI. Magn. Reson. Imag. 26, 905–913 (2008). Alzheimer’s disease in the United States and the
J.-P. & Arbib, M. A.) 873–892 (Plenum Press, 1989). 108. Grady, C. L., Springer, M. V., Hongwanishkul, D., public health impact of delaying disease onset. Am.
87. Nagel, I. E. et al. Load modulation of BOLD response McIntosh, A. R. & Winocur, G. Age-related changes in J. Publ. Health 88, 1337–1342 (1998).
and connectivity predicts working memory brain activity across the adult lifespan. J. Cogn. 133. Raz, N. in Handbook of Aging and Cognition - II (eds
performance in younger and older adults. J. Cogn. Neurosci. 18, 227–241 (2006). Craik, F. I. M. & Salthouse, T. A.) 1–90 (Lawrence
Neurosci. 23, 2030–2045 (2011). 109. Lustig, C. et al. Functional deactivations: change with Erlbaum, 2000).
88. Clapp, W. C., Rubens, M. T., Sabharwal, J. & age and dementia of the Alzheimer type. Proc. Natl 134. Kaup, A. R., Mirzakhanian, H., Jeste, D. V. & Eyler,
Gazzaley, A. Deficit in switching between functional Acad. Sci. USA 100, 14504–14509 (2003). L. T. A review of the brain structure correlates of
brain networks underlies the impact of multitasking on 110. Miller, S. L. et al. Age-related memory impairment successful cognitive aging. J. Neuropsychiatry Clin.
working memory in older adults. Proc. Natl Acad. Sci. associated with loss of parietal deactivation but Neurosci. 23, 6–15 (2011).
USA 108, 7212–7217 (2011). preserved hippocampal activation. Proc. Natl Acad. 135. Persson, J. et al. Structure-function correlates of
This study suggests the intriguing possibility that Sci. USA 105, 2181–2186 (2008). cognitive decline in aging. Cereb. Cortex 16, 907–915
older adults show a reduced ability to resolve 111. Persson, J., Lustig, C., Nelson, J. K. & Reuter-Lorenz, (2006).
interference (a type of inhibition deficit) because P. A. Age differences in deactivation: a link to cognitive 136. Madden, D. J. et al. Adult age differences in the
interfering stimuli disrupt functional connectivity in control? J. Cogn. Neurosci. 19, 1021–1032 (2007). functional neuroanatomy of visual attention: a
task-relevant brain networks, which does not return 112. Duzel, E., Schutze, H., Yonelinas, A. P. & Heinze, H. J. combined fMRI and DTI study. Neurobiol. Aging 28,
to normal as quickly as in young adults. Functional phenotyping of successful aging in long- 459–476 (2007).
89. Epstein, R. & Kanwisher, N. A cortical representation term memory: preserved performance in the absence 137. Salami, A., Eriksson, J., Nilsson, L. G. & Nyberg, L.
of the local visual environment. Nature 392, of neural compensation. Hippocampus 21, 803–814 Age-related white matter microstructural differences
598–601 (1998). (2011). partly mediate age-related decline in processing speed
90. Daselaar, S. M., Fleck, M. S., Dobbins, I. G., Madden, 113. Sambataro, F. et al. Age-related alterations in default but not cognition. Biochim. Biophys. Acta 1822,
D. J. & Cabeza, R. Effects of healthy aging on mode network: impact on working memory 408–415 (2012).
hippocampal and rhinal memoryfunctions: an event- performance. Neurobiol. Aging 31, 839–852 138. Davis, S. W. et al. Assessing the effects of age on long
related fMRI study. Cereb. Cortex 16, 1771–1782 (2010). white matter tracts using diffusion tensor
(2006). 114. Park, D. C., Polk, T. A., Hebrank, A. C. & Jenkins, L. J. tractography. Neuroimage 46, 530–541 (2009).
91. Dennis, N. A. et al. Effects of aging on the neural Age differences in default mode activity on easy and 139. Lockhart, S. et al. Episodic memory function is
correlates of successful item and source memory difficult spatial judgment tasks. Front. Hum. Neurosci. associated with multiple measures of white matter
encoding. J. Exp. Psychol. Learn. Mem. Cogn. 34, 3, 75 (2010). integrity in cognitive aging. Front. Hum. Neurosci. 6,
791–808 (2008). 115. Hedden, T. et al. Disruption of functional connectivity 56 (2012).
92. St Jacques, P., Dolcos, F. & Cabeza, R. Effects of aging in clinically normal older adults harboring amyloid 140. Charlton, R. A., Barrick, T. R., Lawes, I. N., Markus,
on functional connectivity of the amygdala for burden. J. Neurosci. 29, 12686–12694 (2009). H. S. & Morris, R. G. White matter pathways
subsequent memory of negative pictures: a network 116. Allen, E. A. et al. A baseline for the multivariate associated with working memory in normal aging.
analysis of fMRI data. Psychol. Sci. 20, 74–84 (2009). comparison of resting-state networks. Front. Syst. Cortex 46, 474–489 (2010).
93. Addis, D. R., Leclerc, C. M., Muscatell, K. A. & Neurosci. 5, 2 (2011). 141. Chen, N. K., Chou, Y. H., Song, A. W. & Madden, D. J.
Kensinger, E. A. There are age-related changes in 117. Andrews-Hanna, J. R., Reidler, J. S., Sepulcre, J., Measurement of spontaneous signal fluctuations in
neural connectivity during the encoding of positive, Poulin, R. & Buckner, R. L. Functional-anatomic fMRI: adult age differences in intrinsic functional
but not negative, information. Cortex 46, 425–433 fractionation of the brain’s default network. Neuron connectivity. Brain Struct. Funct. 213, 571–585
(2010). 65, 550–562 (2010). (2009).

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 503

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

142. Thomsen, T. et al. Brain localization of attentional 167. Filippini, N. et al. Differential effects of the APOE 189. Kirchhoff, B. A., Anderson, B. A., Barch, D. M. &
control in different age groups by combining functional genotype on brain function across the lifespan. Jacoby, L. L. Cognitive and neural effects of semantic
and structural MRI. Neuroimage 22, 912–919 Neuroimage 54, 602–610 (2011). encoding strategy training in older adults. Cereb.
(2004). 168. Smith, J. C. et al. Interactive effects of physical Cortex 22, 788–799 (2012).
143. Rajah, M., Languay, R. & Grady, C. Age-related activity and APOE‑ε4 on BOLD semantic memory This is one of the few papers to address the effects
changes in right middle frontal gyrus volumes activation in healthy elders. Neuroimage 54, on the brain of cognitive training in older adults
correlate with altered episodic retrieval activity. 635–644 (2011). and shows strong evidence that such changes are
J. Neurosci. 31, 17941–17954 (2011). 169. Grady, C. L. et al. Longitudinal study of the early important for improved behavioural performance.
144. Cabeza, R., Ciaramelli, E., Olson, I. R. & Moscovitch, M. neuropsychological and cerebral metabolic changes in 190. Wagner, A. D., Pare-Blagoev, E. J., Clark, J. &
The parietal cortex and episodic memory: an dementia of the Alzheimer type. J. Clin. Exp. Poldrack, R. A. Recovering meaning: left prefrontal
attentional account. Nature Rev. Neurosci. 9, Neuropsychol. 10, 576–596 (1988). cortex guides controlled semantic retrieval. Neuron
613–625 (2008). 170. Duara, R. et al. Positron emission tomography in 31, 329–338 (2001).
145. Lepage, M., Ghaffar, O., Nyberg, L. & Tulving, E. Alzheimer’s disease. Neurology 36, 879–887 191. Thompson-Schill, S. L. Neuroimaging studies of
Prefrontal cortex and episodic memory retrieval mode. (1986). semantic memory: inferring “how” from “where”.
Proc. Natl Acad. Sci. USA 97, 506–511 (2000). 171. Frackowiak, R. S. J. et al. Regional cerebral oxygen Neuropsychologia 41, 280–292 (2003).
146. Kalpouzos, G., Persson, J. & Nyberg, L. Local brain supply and utilization in dementia. A clinical and 192. Cabeza, R. & Dennis, N. A. in Principles of Frontal
atrophy accounts for functional activity differences in physiological study with oxygen‑15 and positron Lobe Function (eds Stuss, D. T. & Knight, R. T.) (Oxford
normal aging. Neurobiol. Aging 33, 623.e1–623.e13 tomography. Brain 104, 753–778 (1981). Univ. Press, in the press).
(2012). 172. Petersen, R. C. et al. Alzheimer’s Disease 193. Nyberg, L. et al. Longitudinal evidence for diminished
147. Haber, S. N. & Knutson, B. The reward circuit: linking Neuroimaging Initiative (ADNI): clinical frontal cortex function in aging. Proc. Natl Acad. Sci.
primate anatomy and human imaging. characterization. Neurology 74, 201–209 (2010). USA 107, 22682–22686 (2010).
Neuropsychopharmacology 35, 4–26 (2010). 173. Filippi, M. & Agosta, F. Structural and functional 194. Beason-Held, L. L., Kraut, M. A. & Resnick, S. M. I.
148. Mell, T. et al. Altered function of ventral striatum network connectivity breakdown in Alzheimer’s Longitudinal changes in aging brain function.
during reward-based decision making in old age. disease studied with magnetic resonance imaging Neurobiol. Aging 29, 483–496 (2008).
Front. Hum. Neurosci. 3, 34 (2009). techniques. J. Alzheimers Dis. 24, 455–474 (2011). 195. Beason-Held, L. L., Kraut, M. A. & Resnick, S. M. II.
149. Samanez-Larkin, G. R. et al. Anticipation of monetary 174. Dickerson, B. C. & Sperling, R. A. Functional Temporal patterns of longitudinal change in aging
gain but not loss in healthy older adults. Nature abnormalities of the medial temporal lobe memory brain function. Neurobiol. Aging 29, 497–513 (2008).
Neurosci. 10, 787–791 (2007). system in mild cognitive impairment and Alzheimer’s 196. Thambisetty, M., Beason-Held, L., An, Y., Kraut, M. A.
150. Dreher, J. C., Meyer-Lindenberg, A., Kohn, P. & disease: insights from functional MRI studies. & Resnick, S. M. APOE ε4 genotype and longitudinal
Berman, K. F. Age-related changes in midbrain Neuropsychologia 46, 1624–1635 (2008). changes in cerebral blood flow in normal aging. Arch.
dopaminergic regulation of the human reward system. 175. Yassa, M. A. et al. High-resolution structural and Neurol. 67, 93–98 (2010).
Proc. Natl Acad. Sci. USA 105, 15106–15111 functional MRI of hippocampal CA3 and dentate gyrus 197. Weiner, M. W. et al. The Alzheimer’s Disease
(2008). in patients with amnestic mild cognitive impairment. Neuroimaging Initiative: a review of papers published
151. Sambataro, F. et al. Catechol-O‑methyltransferase Neuroimage 51, 1242–1252 (2010). since its inception. Alzheimers Dement. 8, S1–S68
valine158methionine polymorphism modulates brain 176. Yassa, M. A., Muftuler, L. T. & Stark, C. E. Ultrahigh- (2012).
networks underlying working memory across resolution microstructural diffusion tensor imaging 198. Grigg, O. & Grady, C. L. Task-related effects on the
adulthood. Biol. Psychiatry 66, 540–548 (2009). reveals perforant path degradation in aged humans in temporal and spatial dynamics of resting-state
152. Tunbridge, E. M., Bannerman, D. M., Sharp, T. & vivo. Proc. Natl Acad. Sci. USA 107, 12687–12691 functional connectivity in the default network. PLoS
Harrison, P. J. Catechol-O‑methyltransferase inhibition (2010). ONE 5, e13311 (2010).
improves set-shifting performance and elevates 177. de Rover, M. et al. Hippocampal dysfunction in 199. Krishnan, A., Williams, L. J., McIntosh, A. R. & Abdi, H.
stimulated dopamine release in the rat prefrontal patients with mild cognitive impairment: a functional Partial Least Squares (PLS) methods for neuroimaging:
cortex. J. Neurosci. 24, 5331–5335 (2004). neuroimaging study of a visuospatial paired associates a tutorial and review. Neuroimage 56, 455–475
153. Backman, L. et al. Dopamine D1 receptors and age learning task. Neuropsychologia 49, 2060–2070 (2011).
differences in brain activation during working memory. (2011). 200. Power, J. D., Fair, D. A., Schlaggar, B. L. & Petersen,
Neurobiol. Aging 32, 1849–1856 (2011). 178. Kochan, N. A. et al. Functional alterations in brain S. E. The development of human functional brain
154. Braskie, M. N. et al. Relationship of striatal dopamine activation and deactivation in mild cognitive networks. Neuron 67, 735–748 (2010).
synthesis capacity to age and cognition. J. Neurosci. impairment in response to a graded working memory 201. Stephan, K. E. et al. Dynamic causal models of neural
28, 14320–14328 (2008). challenge. Dement. Geriatr. Cogn. Disord. 30, system dynamics: current state and future extensions.
155. Landau, S. M., Lal, R., O.’Neil, J. P., Baker, S. & 553–568 (2010). J. Biosci. 32, 129–144 (2007).
Jagust, W. J. Striatal dopamine and working memory. 179. Protzner, A. B., Mandzia, J. L., Black, S. E. & 202. Roebroeck, A., Formisano, E. & Goebel, R. Mapping
Cereb. Cortex 19, 445–454 (2009). McAndrews, M. P. Network interactions explain directed influence over the brain using Granger
156. Braskie, M. N. et al. Correlations of striatal dopamine effective encoding in the context of medial temporal causality and fMRI. Neuroimage 25, 230–242
synthesis with default network deactivations during damage in MCI. Hum. Brain Mapp. 32, 1277–1289 (2005).
working memory in younger adults. Hum. Brain Mapp. (2011). 203. Smith, S. M. et al. The danger of systematic bias in
32, 947–961 (2011). 180. Greicius, M. D., Srivastava, G., Reiss, A. L. & Menon, V. group-level FMRI-lag-based causality estimation.
157. Fischer, H. et al. Simulating neurocognitive aging: Default-mode network activity distinguishes Neuroimage 59, 1228–1229 (2012).
effects of a dopaminergic antagonist on brain activity Alzheimer’s disease from healthy aging: evidence from 204. Sperling, R. A. et al. Functional alterations in memory
during working memory. Biol. Psychiatry 67, functional MRI. Proc. Natl Acad. Sci. USA 101, networks in early Alzheimer’s disease. Neuromolecular
575–580 (2010). 4637–4642 (2004). Med. 12, 27–43 (2010).
158. Morcom, A. M. et al. Memory encoding and dopamine 181. Sala-Llonch, R. et al. Greater default-mode network 205. Van Dijk, K. R., Sabuncu, M. R. & Buckner, R. L. The
in the aging brain: a psychopharmacological abnormalities compared to high order visual influence of head motion on intrinsic functional
neuroimaging study. Cereb. Cortex 20, 743–757 processing systems in amnestic mild cognitive connectivity MRI. Neuroimage 59, 431–438 (2012).
(2010). impairment: an integrated multi-modal MRI study. 206. Honey, C. J. et al. Predicting human resting-state
159. Backman, L., Small, B. J. & Fratiglioni, L. Stability of J. Alzheimers Dis. 22, 523–539 (2010). functional connectivity from structural connectivity.
the preclinical episodic memory deficit in Alzheimer’s 182. Han, S. D. et al. Functional connectivity variations in Proc. Natl Acad. Sci. USA 106, 2035–2040 (2009).
disease. Brain 124, 96–102 (2001). mild cognitive impairment: associations with cognitive 207. Hagmann, P. et al. Mapping the structural core of
160. Braak, H., Braak, E. & Bohl, J. Staging of Alzheimer- function. J. Int. Neuropsychol. Soc. 18, 39–48 human cerebral cortex. PLoS Biol. 6, e159 (2008).
related cortical destruction. Eur. Neurol. 33, (2012). 208. Greicius, M. D., Supekar, K., Menon, V. & Dougherty,
403–408 (1993). 183. Petrella, J. R., Sheldon, F. C., Prince, S. E., Calhoun, V. D. R. F. Resting-state functional connectivity reflects
161. Strittmatter, W. J. et al. Apolipoprotein E: high-avidity & Doraiswamy, P. M. Default mode network structural connectivity in the default mode network.
binding to β-amyloid and increased frequency of type connectivity in stable versus progressive mild Cereb. Cortex 19, 72–78 (2009).
4 allele in late-onset familial Alzheimer disease. Proc. cognitive impairment. Neurology 76, 511–517 209. Luk, G., Bialystok, E., Craik, F. & Grady, C. Lifelong
Natl Acad. Sci. USA 90, 1977–1981 (1993). (2011). bilingualism maintains white matter integrity in older
162. Bookheimer, S. Y. et al. Patterns of brain activation in 184. Buckner, R. L. et al. Cortical hubs revealed by intrinsic adults. J. Neurosci. 31, 16808–16813 (2011).
people at risk for Alzheimer’s disease. N. Engl. J. Med. functional connectivity: mapping, assessment of 210. Bialystok, E. & Craik, F. Cognitive and linguistic
343, 450–456 (2000). stability, and relation to Alzheimer’s disease. processing in the bilingual mind. Curr. Direct. Psychol.
163. Dickerson, B. C. et al. Increased hippocampal J. Neurosci. 29, 1860–1873 (2009). Sci. 19, 19–23 (2010).
activation in mild cognitive impairment compared to 185. Scholz, J., Klein, M. C., Behrens, T. E. & Johansen- 211. Mantyla, T. & Backman, L. Encoding variability and
normal aging and AD. Neurology 65, 404–411 Berg, H. Training induces changes in white-matter age-related retrieval failures. Psychol. Aging 5,
(2005). architecture. Nature Neurosci. 12, 1370–1371 545–550 (1990).
164. Trivedi, M. A. et al. fMRI activation during episodic (2009). 212. Morse, C. K. Does variability increase with age? An
encoding and metacognitive appraisal across the 186. Gauthier, I., Skudlarski, P., Gore, J. C. & Anderson, A. W. archival study of cognitive measures. Psychol. Aging 8,
lifespan: risk factors for Alzheimer’s disease. Expertise for cars and birds recruits brain areas 156–164 (1993).
Neuropsychologia 46, 1667–1678 (2008). involved in face recognition. Nature Neurosci. 3, 213. MacDonald, S. W., Nyberg, L. & Backman, L. Intra-
165. Dennis, N. A. et al. Temporal lobe functional activity 191–197 (2000). individual variability in behavior: links to brain
and connectivity in young adult APOE ε4 carriers. 187. Erickson, K. I. et al. Training-induced plasticity in older structure, neurotransmission and neuronal activity.
Alzheimers Dement. 6, 303–311 (2010). adults: effects of training on hemispheric asymmetry. Trends Neurosci. 29, 474–480 (2006).
166. Trivedi, M. A. et al. Reduced hippocampal activation Neurobiol. Aging 28, 272–283 (2007). 214. Hultsch, D. F., MacDonald, S. W. & Dixon, R. A.
during episodic encoding in middle-aged individuals at 188. Berry, A. S. et al. The influence of perceptual training Variability in reaction time performance of younger
genetic risk of Alzheimer’s disease: a cross-sectional on working memory in older adults. PLoS ONE 5, and older adults. J. Gerontol. B Psychol. Sci. Social
study. BMC Med. 4, 1 (2006). e11537 (2010). Sci. 57, P101–P115 (2002).

504 | JULY 2012 | VOLUME 13 www.nature.com/reviews/neuro

© 2012 Macmillan Publishers Limited. All rights reserved


REVIEWS

215. West, R., Murphy, K. J., Armilio, M. L., Craik, F. I. & An early landmark study addressing variability variability in development. PLoS Comput. Biol. 4,
Stuss, D. T. Lapses of intention and performance in brain activity by showing that intrinsic, ongoing e1000106 (2008).
variability reveal age-related increases in fluctuations brain activity influences the way in which 227. Garrett, D. D., Kovacevic, N., McIntosh, A. R. & Grady,
of executive control. Brain Cogn. 49, 402–419 stimulus-evoked activity is expressed. The authors C. L. Blood oxygen level-dependent signal variability is
(2002). describe the effect of such stimuli as: “the more than just noise. J. Neurosci. 30, 4914–4921
216. Williams, B. R., Hultsch, D. F., Strauss, E. H., Hunter, additional ripples caused by tossing a stone into a (2010).
M. A. & Tannock, R. Inconsistency in reaction time wavy sea.”
across the life span. Neuropsychology 19, 88–96 221. Faisal, A. A., Selen, L. P. & Wolpert, D. M. Noise in the Acknowledgements
(2005). nervous system. Nature Rev. Neurosci. 9, 292–303 C.G.’s research is supported by the Canadian Institutes of
217. Dixon, R. A. et al. Neurocognitive markers of cognitive (2008). Health Research, the Canada Research Chairs program, the
impairment: exploring the roles of speed and 222. Stein, R. B., Gossen, E. R. & Jones, K. E. Neuronal Ontario Research Fund and the Canadian Foundation for
inconsistency. Neuropsychology 21, 381–399 variability: noise or part of the signal? Nature Rev. Innovation.
(2007). Neurosci. 6, 389–397 (2005).
218. MacDonald, S. W., Hultsch, D. F. & Dixon, R. A. 223. Ghosh, A., Rho, Y., McIntosh, A. R., Kotter, R. & Jirsa, Competing interests statement
Performance variability is related to change in V. K. Noise during rest enables the exploration of the The author declares no competing financial interests.
cognition: evidence from the Victoria brain’s dynamic repertoire. PLoS Comput. Biol. 4,
Longitudinal Study. Psychol. Aging 18, 510–523 e1000196 (2008).
(2003). 224. Basalyga, G. & Salinas, E. When response DATABASES
219. Macdonald, S. W., Hultsch, D. F. & Dixon, R. A. variability increases neural network robustness to Alzheimer’s Disease Neuroimaging Initiative: http://www.
Predicting impending death: inconsistency in speed is synaptic noise. Neural Comput. 18, 1349–1379 adni-info.org/
a selective and early marker. Psychol. Aging 23, (2006).
595–607 (2008). 225. Ma, W. J., Beck, J. M., Latham, P. E. & Pouget, A. FURTHER INFORMATION
220. Arieli, A., Sterkin, A., Grinvald, A. & Aertsen, A. Bayesian inference with probabilistic population Cheryl Grady’s homepage: http://research.baycrest.org/
Dynamics of ongoing activity: explanation of the large codes. Nature Neurosci. 9, 1432–1438 (2006). cgrady
variability in evoked cortical responses. Science 273, 226. McIntosh, A. R., Kovacevic, N. & Itier, R. J. Increased ALL LINKS ARE ACTIVE IN THE ONLINE PDF
1868–1871 (1996). brain signal variability accompanies lower behavioral

NATURE REVIEWS | NEUROSCIENCE VOLUME 13 | JULY 2012 | 505

© 2012 Macmillan Publishers Limited. All rights reserved

You might also like