Download as pdf or txt
Download as pdf or txt
You are on page 1of 70

PA R T  

I I

G E N ET IC
14.
THE A PPLE A ND THE TR EE
W H AT FA M I LY A N D OF F SPR I NG S T U DI E S OF BI P OL A R DI S OR DE R
SHOW A B OU T M E CH A N I SM S OF R I SK A N D PR E V E N T ION

Guillermo Perez Algorta, Anna Van Meter, and Eric A. Youngstrom

INTRODUCTION general population (Merikangas et al., 2010; Van Meter,


Moreira, & Youngstrom, 2011). The concordance rate for
Being a parent is a challenging and often emotionally BD in monozygotic twins is 50% to 60% and 13% to 23%
charged life role. Suffering from a mental illness, such as in dizygotic twins and non-twin siblings (Goodwin &
bipolar disorder (BD), adds complications to one’s role as Jamison, 2007), further increasing the likelihood of mul-
a parent and can result in unexpected outcomes. Though tiple cases of BD within a family. Greater familial loading
conventional perspectives focus on the ways in which par- is associated with earlier illness onset (Lapalme et al., 1997;
ents influence and mold their children—through genes, Pauls, 1992), which often leads to a more difficult illness
parenting practices, and environment—children have a course (James, 2011). Like other youth with BD, those who
significant impact on their parents as well. Parenting is a have a family history tend to have an index mood episode
dynamic phenomenon determined by the child´s character- of depression, followed by a high rate of recurrence and/or
istics, contextual sources of stress and support, and parent chronic mood lability (Duffy, Alda, Hajek, & Grof, 2009).
personality (Belsky, 1984). For people with BD, becoming The heritability of BD may vary as a function of the
a parent can be protective, by providing new motivation for type of bipolar or the severity of illness in the parent. In one
better self-care with the goal of remaining healthy and pres- study, the odds ratio of meeting criteria for an Axis I disor-
ent as a parent, but having a child can also be a risk factor der for BD offspring was 15.0 when a parent had BD with
due to significant time and energy expenditures associated a lifetime history of psychotic symptoms versus 3.3 when a
with parenting (Ehlers, Frank, & Kupfer, 1988). Our goal parent had bipolar II disorder (Garcia-Amador et al., 2013).
is to review the role that heritability plays in families with In the same study, half of BD offspring met the diagnos-
BD and to extend the discussion beyond genes and biology tic criteria for at least one Axis I disorder, the most com-
to include the reciprocal social and behavioral influences mon being attention deficit hyperactivity disorder (30%),
parents and offspring have on one another. anxiety disorders (14%), and other affective disorders (10%;
Garcia-Amador et al., 2013), indicating that bipolar family
history is a risk factor for a wide range of pathology and not
B I P OL A R FA M I L I E S:  T H E ROL E specifically for BD.
OF   G E N E S A N D B IOL O G Y In addition to BD, other affective disorders and anxiety
are the most common Axis I disorders among BD offspring
Bipolar disorder is highly heritable (Smoller & Finn, 2003). (Lapalme, Hodgins, & LaRoche, 1997; Vandeleur et  al.,
Among youth with one parent with BD, the rate of BD is 5% 2012), with BD offspring being about three times more
to 15% (Birmaher et al., 2009; Lapalme et al., 1997; Singh likely to develop a non-BD Axis I  disorder than healthy
et al., 2007). Among those youth who have two parents controls, on average. However, some studies have found
with BD, the rate is as high as 30% (Goodwin & Jamison, rates of Axis I disorders as high as 72% among BD offspring
2007; Gottesman, Laursen, Bertelsen, & Mortensen, 2010). (Mesman, Nolen, Reichart, Wals, & Hillegers, 2013).
This is a risk of 5 to 10 times higher than expected in the In addition to increased risk of BD, bilineal BD history

171
is associated with greater severity of symptoms such as to the depressogenic effects of stress later in life (Dougherty
depressed and irritable mood, lack of mood reactivity, and et al., 2013).
rejection sensitivity. In families parented by someone with Certain temperamental styles are thought to represent a
BD, the likelihood of the offspring having BD or another genetic risk factor for the development of affective disorders
Axis I disorder is high, which further impacts the family (Evans et al., 2008; Gonda et al., 2006; Greenwood, Akiskal,
dynamic and functioning. Akiskal, & Kelsoe, 2012), and it has been proposed that
Conversely, better functioning in biological co-parents temperament may be the primary route by which genetic
is associated with a significantly lower frequency of Axis risk for BD is passed on (Evans et  al., 2005; Yuan et  al.,
I  disorder in BD offspring (Garcia-Amador et  al., 2013), 2012). Even before meeting criteria for a psychiatric illness,
suggesting that genes are not fate; parent behavior and children of BD parents often show temperamental char-
environment play an important role too. Family mem- acteristics thought to precede clinical symptoms in many
bers’ behavior can affect the illness course, and the illness cases. These include high motor activity, low frustration
reciprocally impacts family functioning (Lytton, 1990a, tolerance, and emotional sensitivity, which may be evident
1990b). very early in a child’s life, reflecting innate characteristics
Children’s reaction to their family environment further (Chang, Steiner, & Ketter, 2000). These traits increase the
influences the likelihood of Axis I  pathology. Consistent propensity to have exaggerated or inappropriate emotional
with the diathesis-stress model (Zuckerman, 1999), a reactions to negative life events and stressors, along with
genetic vulnerability, plus individual differences in bio- poor affective regulation (Duffy et al., 2007). BD offspring
logical reactivity to stress, may increase susceptibility for tend toward greater mood lability and irritability than
developing a mental illness. One mechanism proposed other children (Birmaher et al., 2013), increasing the risk
to explain the intergenerational transmission of risk for for, and intensity of, conflict with both parents and other
psychopathology and other negative health outcomes in individuals. There also is a tendency for bipolar offspring
the offspring of depressed parents is dysregulation of the to experience heightened distress in response to conflict
hypothalamic-pituitary-adrenal (HPA) axis, one of the and to react in suboptimal ways to psychosocial stressors.
body’s major stress-response systems (Goodman & Gotlib, This situation may develop into a positive feedback loop for
1999). Specifically, heightened HPA axis reactivity to envi- both the child and the parent, whereby each exacerbates
ronmental stressors has been posited to play a role in the the other’s mood symptoms through heightened reactivity
process linking stress to illness (Holsboer, 2000). and conflict. These strong, negative responses can create a
Both genetics and experience play a role in the stress chain reaction, making it difficult for children with BD to
response. Cortisol reactivity is moderately heritable socialize appropriately (Bella et al., 2011; Chang, Steiner, &
(Steptoe, van Jaarsveld, Semmler, Plomin, & Wardle, Ketter, 2003). Some will act aggressively or otherwise alien-
2009), and individual traits, such as temperament and ate peers and family members, limiting social support and
behavior, are linked to cortisol response to stress (Gunnar increasing the risk for affective psychopathology above and
& Vazquez, 2006). Furthermore, environmental influences, beyond the genetic risk conferred.
such as stress, adversity, and parental psychopathology, are
linked to poor HPA functioning in youth, but a positive
family environment can moderate the relation between B I DI R E C T ION A L I T Y
early adversity and offspring neuroendocrine functioning
(Essex et  al., 2011; Luecken & Appelhans, 2006). These The default message communicated by family studies of BD
effects can be long-lasting, with exposure to maternal psy- is that parents with BD confer great risk of psychiatric ill-
chopathology in childhood associated with dysregulated ness on their children. However, children also create risk
basal cortisol into the offspring’s adolescence (Essex et al., for their parents. Even if they do not meet criteria for any
2011). The effects can also be additive; offspring exposed psychiatric illness, children require a great amount of time
to both parental depression and hostile parenting behav- and energy. They test their parents’ patience and are often
iors during early childhood displayed high cortisol levels in associated with increased stress, sleep deprivation, and poor
response to a laboratory stressor (Dougherty, Tolep, Smith, self-care—all risk factors for the onset of a mood episode.
& Rose, 2013). Other studies have found high levels of cor- Social zeitgeber theory suggests that genetically vulner-
tisol in depressed adolescents and adults, suggesting that able individuals may be at high risk for a mood episode
increased stress sensitivity, due to early dysregulation of the when they experience a disruption of their routine. Other
HPA axis, may render high-risk offspring more vulnerable people, obligations, and even the seasons act as zeitgebers,

17 2   •   S E C T I O N I I : I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
imposing structure on individuals and helping them to An emotionally charged family environment may also
maintain regular moods. If a zeitgeber is disturbed, or if a serve as a method of transmission, whereby the behavior of
zeitstörer, an external rhythm disrupter, appears, the regu- a parent (or child) experiencing a mood episode triggers a
larity of routine is threatened and mood episodes are more mood episode in the other family member (Lytton, 1990a).
likely to occur (Frank, Swartz, & Kupfer, 2000). Behavioral geneticists describe this as an “extended pheno-
Children exert powerful influences on parents’ sched- type,” where the genes of risk in the child also are present
ules. Once they begin attending school and have regular in the parent, conveying not only biological risk but also
routines themselves, they may become zeitgebers, imposing changing the environment in a way that further amplifies
structure on their parents with positive effects. However, effects (Plomin, 1995). Factors such as expressed emotion
as new babies, children are undoubtedly zeitstörers, mak- and family environment are important to the onset and
ing routine very difficult with irregular and important maintenance of both mood episodes and periods of euthy-
demands. This is likely a factor in the development of post- mic mood (Calam, 2012; Miklowitz, 2004; Miklowitz &
partum pathology. Johnson, 2009), and many of the empirically supported
Though the birth of a child is usually a happy event, the treatments for BD focus on reducing expressed emotion
incidence of psychiatric disturbance among parents—both and building stable family environments in order to reduce
mothers and fathers—increases in the perinatal and post- mood episode frequency and intensity. When multiple
partum periods (Davenport & Adland, 1982). Perinatal members experience mood disruptions, families are less
factors increase risk of BD specifically (Sharma, 2012). likely to achieve the type of structure and stability that pro-
Some mothers with diagnosed BD may choose or be advised motes good mental health.
by their doctor to discontinue some medications to reduce
risk to the fetus, which increases the likelihood of a mood
episode (Davenport & Adland, 1982). Hormonal changes
may affect a woman’s mood (Meinhard, 2013). Extreme T H E A PPL E :   L I F E F OR   T H E
changes to either parent’s sleep–wake cycle or schedule are C H I L DR E N OF   PA R E N T S
risk factors, as is the potential for discord in the marital W I T H   B I P OL A R DI S OR DE R
relationship (Marks, 1992; Murray, 2010). Finally, the tran-
sitional period around the addition of a child is often very Reports on the psychosocial effects of having a parent with
emotional; for people at risk for BD or who already have BD are inconsistent (Jones & Bentall, 2008). Offspring
affective pathology, heightened emotional responses—and of parents with BD often show aggressive behavior that is
an inability to successfully regulate these responses—can strongly influenced by deficient parenting (Chang et al.,
result in a mood episode (Davenport & Adland, 1982). 2003). Parent behavior, and the associated psychosocial
Emotion-regulation deficits are common among both disturbance in the family, likely confers as much if not
youth and adults with BD (Green, Cahill, & Malhi, 2007; more risk than genes (Rutter & Quinton, 1984). But some
Townsend, 2012). According to Hodgins et  al.’s (2002) healthy offspring of BD parents perceive their mothers as
model, children born with a genetic predisposition to less rejecting, more emotionally warm, and less overpro-
depression or BD inherit “a tendency to react emotionally tective than offspring of healthy parents (Reichart et al.,
to stressors and daily hassles” and are likely to have parent(s) 2007). Interestingly, the same study found that offspring
who also tend to react strongly. This combination of high with BD reported more rejection by their fathers and moth-
reactivity from both child and parents leads to higher levels ers compared with offspring with other diagnoses or with-
of stress and conflict within the family. out diagnosis. Perhaps the offspring’s disorder has a greater
Such emotionally labile environments may provoke impact on the child-rearing practices of the parents than
the onset, or serve to maintain, mood episodes. A  child the BD of the parents, or perhaps the child’s experience of
with a short emotional fuse may also have a parent prone the parents is filtered by the child’s psychopathology.
to strong emotional responses, which creates a family Offspring with BD may be more sensitive to cues from
environment where parent and child repeatedly trigger their parents, particularly negative ones. Studies of emo-
strong emotional exchanges. Additionally, a child with tion perception find that youth with BD tend to interpret
a parent who reacts strongly has fewer opportunities to neutral faces as more hostile and fear provoking (Dickstein
learn how to regulate his or her emotions, making it more et  al., 2007; Rich et  al., 2005). These negative perceptual
likely that emotionally-intense interactions will occur biases in social interactions may then evoke more extreme
frequently—both with family members and others. emotional responses (Green et al., 2007).

T he A pple and the T ree   •   17 3


Children of BD parents face multiple challenges. parent returns home, children may have ambivalent feel-
During manic episodes, parents are overactive, aggres- ings about the parent (Sturges, 1977). Fear and uncertainty
sive, unpredictable, or destructive. During depressive can also interfere with the child’s relationship with his or
episodes, they are inactive, absent, or even suicidal. The her parent, leading children to keep a distance in an effort
consequences—painful emotions, anger, anxiety, concern, to protect themselves (Mordoch & Hall, 2008). Children
and ambivalence—take a toll on the offspring (Hinshaw, may also take on “adult” responsibilities, feeling that they
2008; Sturges, 1977). Compared to depressed or healthy cannot count on a parent to provide for them (Handley,
mothers, mothers with BD can be more negative in Farrell, Josephs, Hanke, & Hazelton, 2001; Maybery
their interactions with their children (Inoff-Germain, et al., 2005). These extra responsibilities can contribute to
Nottelmann, & Radke-Yarrow, 1992). The toddler and poor coping. Children report withdrawing, not attending
preschool years may be a neurodevelopmental period par- school, and keeping feelings to themselves when faced with
ticularly sensitive to negative environmental exposures, pro- adversity at home (Maybery et al., 2005). Although there
ducing lasting perturbations to offspring stress physiology is likely to be greater responsibility placed on BD offspring
(Melhem, Walker, Moritz, & Brent, 2008). From as early than other youth, in many cases they may be unable to meet
as infancy and preschool, children of parents with mood these demands. Children and adolescents with BP experi-
disorders exhibit psychosocial, emotional, and behavioral ence significant functional impairment across multiple
problems (Brennan et  al., 2000; Hammen, 2009), likely domains, including academic, interpersonal, and overall
related to the fact that mood disorders are associated with measures of functioning (Freeman et al., 2009; Goldstein
negative, hostile child-rearing behaviors (Lovejoy, Graczyk, et al., 2009).
O’Hare, & Neuman, 2000). Unfortunately, when a parent is ill, the children are
Later events also have detrimental effects, as children often neglected—at least from a mental health perspective.
gain awareness of their situation. Older children start to Parent hospitalization is a stressful event, associated with
see the effects of medication on a parent; in one study a odds ratios of 3.3 for a child first episode of mania (Kessing,
child described his mother as being “doped up to the hilt” Agerbo, & Mortensen, 2004), and there are few resources
(Maybery, Ling, Szakacs, & Reupert, 2005). A  parent’s available to help children cope in these situations (Handley
decision to stop taking medication, or to make some other et al., 2001; Maybery et al., 2005). There is limited coordi-
change in his or her life, can result in declining mental nation between mental health services and child protective
health and an unnerving or frightening situation for the services in most communities, and institutions are often
child (Spiegelhoff & Ahia, 2011). unaware that a hospitalized patient has children at home,
Children often do not know the specifics of their par- adding to the challenge of marshalling good childcare
ent’s illness, though they are attuned to changing parental while the parent is receiving mental health services (Park,
behavior due to mental illness. Not having an explanation Solomon, & Mandell, 2006).
for why a parent is acting erratically is more upsetting than Additionally, assortative mating increases the likeli-
when behavior changes can be attributed to mental illness hood that the remaining parent may also have a psychiatric
(Mordoch & Hall, 2008). Parents often try to comfort chil- illness (Melhem et al., 2008), introducing other complica-
dren, rather than answering their questions or talking with tions (Sands, Koppelman, & Solomon, 2004). A  child in
them openly about the situation, which can further con- this situation may be sent to stay with unfamiliar people
fuse and scare children (Sturges, 1977). Lack of knowledge, during a period of parental illness or may stay and observe
or lack of conversation, about the illness may also leave the negative behavior changes in a remaining parent. Most
child feeling as though something is “wrong” or that there mothers with mental illness are single parents, and children
is a shameful family secret. from single-parent homes will very likely find themselves
Children also experience negative emotions, includ- in unfamiliar, stressful circumstances when a parent has a
ing fear and anxiety, when a parent must be hospitalized mood episode (Sands et al., 2004).
(Mordoch & Hall, 2008; Sturges, 1977). A lack of openness Parental psychiatric hospitalization doubles the risk of
about the situation can compound these fears, as worries the family’s involvement with child protective services and
about the parent dying or otherwise not returning home triples likelihood of children being removed from the home
arise (Riebschleger, 2004). Children also may feel guilt (Park et al., 2006). Children of mothers with mental illness
related to parent hospitalization, perhaps due to questions are more likely to become foster children, and their par-
about what their role was in the hospitalization or to hav- ents are then put in the position of having to defend their
ing to rely on others during a parent’s absence. When the parental fitness, rather than the burden of proof falling on

174   •   S E C T I O N I I : I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
the state (Spiegelhoff & Ahia, 2011). Neglect and partici- for a parent to meet his or her obligations (Spiegelhoff &
pation in foster/residential care are both associated with Ahia, 2011). Extreme mood variability, suicidality, and fre-
increased suicide risk for youth (Christoffersen, Poulsen, & quent mood episodes will threaten the parent’s ability to
Nielsen, 2003). form a healthy attachment with the child, adding to worse
Research suggests that the amount of stress due to outcomes for both the parent and child.
parental mental illness on a child may be greater than the Difficulty connecting with one’s child may trigger nega-
stress attached to a death or divorce (Sturges, 1977). Sadly, tive feelings in the parent. Parents with BD rate themselves
children of people with BD may experience both; these in a more critical light than control parents (Spitzer, 2002).
youth are at heightened risk of losing a parent to suicide Many of the psychosocial difficulties associated with BD
(Gould, Greenberg, Velting, & Shaffer, 2003; Ljung et al., may also create extra challenges for parents. People with
2013). This can become a cascade. Death of one parent by BD experience declines in job status and income, failure
suicide increases the risk of mental illness in the remaining to marry, interpersonal relationship deficits, lower enjoy-
parent (Melhem et al., 2008), leading to even more stress ment of recreational activities, and overall diminished con-
in the household (Kuramoto, Brent, & Wilcox, 2009). The tentment compared with controls (Calabrese et al., 2003;
effects of parental suicide on the child can be long-lasting Coryell, Scheftner, Keller, Endicott, & al, 1993; Maybery
and include depressive symptoms, interpersonal problems, et al., 2005; Spiegelhoff & Ahia, 2011), all of which have
stigma, guilt, and anger. Parental death by suicide is also the potential to increase the sense of burden associated
associated with the development of BD in the offspring with parenting.
(Tsuchiya, Agerbo, & Mortensen, 2005), with an odds Parents with BD report having both considerable per-
ratio of between 4 and 5, depending on the age at death sonal difficulties and difficulties in rearing their children
(Mortensen, Pedersen, Melbye, Mors, & Ewald, 2003). (Livesley, 1995). Parents with mental illness also report
Parental death by suicide before age 11 has also been associ- higher levels of stress related to parenting and lower levels
ated with more frequent suicide attempts by the offspring of satisfaction in the relationships with their children than
(Goodwin, Beautrais, & Fergusson, 2004). other parents (Park et al., 2006). They also report that their
children have greater adjustment difficulties and higher lev-
els of behavior problems than other youth. Parents with BD
judge their families as less cohesive, less organized, and more
T H E T R E E :  HOW PA R E N T HO OD conflicted than other parents (Chang, Blasey, Ketter, &
A F F E C T S PE OPL E W I T H  Steiner, 2001). Though the offspring of a parent with BD
B I P OL A R DI S OR DE R often have behavioral and mood problems, parents may
also perceive the behaviors as being even worse due to
Adults with BD who are able to manage their symptoms their own stress and compromised coping (De Los Reyes,
effectively have better outcomes. Unfortunately, during the Goodman, Kliewer, & Reid-Quinones, 2008; Youngstrom,
age range when most people start families, many people Loeber, & Stouthamer-Loeber, 2000). Child characteris-
with BD are undiagnosed and not receiving optimal treat- tics, parent characteristics, and parenting situations influ-
ment. Parenting responsibilities often interfere with adults ence functioning and stress in the dyadic system (Abidin,
seeking mental health services for themselves (Knight & 1992). Typically, reports about parenting and mental illness
Wallace, 2003). Drancourt et  al. (2013) report that, on focus on risks to children when a parent has a mental ill-
average, the first full syndromal mood episode occurs at ness (Rutter & Quinton, 1984), but this understates the
25.3  years, and the mean age at first recorded treatment impact of child psychopathology on the parent’s behavior.
with a mood stabilizer is 34.9, giving a mean of duration At the extreme, child behavior problems can exacerbate the
of untreated BD of 9.6 years. Longer duration of untreated risk of suicide in parents with BD (Cerel, Fristad, Weller, &
BD correlates with negative outcomes, including suicide Weller, 1999).
attempts, more major mood episodes, greater mood insta- Risk for hospitalization worsens parents’ self-critical,
bility, and more frequent mixed episodes. At the time when negative feelings as it keeps them from their children for
young children might be drawing most heavily on a parent’s extended periods of time. Parents facing hospitalization can
internal resources, the likelihood that bipolar symptoms have difficulty asking for help from others (Handley et al.,
would be unrecognized or untreated is high. Substance 2001), increasing the likelihood of negative circumstances
abuse frequently occurs with BD, and it often contributes to befalling their children. Even short hospital stays are chal-
delays in diagnosis and treatment, making it more difficult lenging, as parents are more likely to remain symptomatic

T he A pple and the T ree   •   175


and may not be able to return to their responsibilities effec- both to reduce the negative effects and to promote positive
tively once discharged (Thomas & Kalucy, 2003). functioning.
Parents report a scarcity of professional support for their
children during their hospitalization, and they also are
A S S E S SM E N T FOR BE T T E R
uncertain about what and how much to tell their children
DE T EC T ION A N D E A R L I E R
about their mental illness (Maybery et  al., 2005), height-
I N T E RV E N T ION
ening the sense of secrecy and burden. Rates of women
with mental health problems having children are increas- The first steps in intervening are identifying the problem,
ing, a change from the past when mentally ill people were gauging how far it has progressed, and then defining inter-
often hospitalized long term, but mental health and fam- vention targets. We argue that the unit of assessment should
ily services have not kept up (Sands et al., 2004). Because be the family and not just the identified patient: Family-level
professional services are lacking, welfare often becomes the factors play key roles in illness transmission and moderate
default support for women with mental illness and their illness progression; they also have similar roles in efficacious
families. Women and their children are often separated at treatments. Assessing the family provides more contex-
the point of hospitalization or due to disruptive behavior tual information about the individual case, but it also can
during a mood episode (Sands et al., 2004). Mothers with enhance chances of treatment success. There is huge unmet
mental illness are also more likely to lose custody of their need:  More than 50% of people suffering BD are receiv-
children than other mothers (Park et  al., 2006). In one ing no treatment in the United States (Wang et al., 2005).
study, 25% of women with mental illness had involvement The global burden of BD appears similarly high around the
with child welfare services versus only 4% of women with- world, and the chances of obtaining adequate assessment
out mental illness (Park et al., 2006). Mothers often report and treatment are yet lower in countries with less developed
poor communication regarding their children’s status, as mental health service systems.
well as confusion as to whether custodial arrangements One strategy for improving detection would be to
made while they were ill are permanent (Sands et al., 2004). routinely assess family history for every case seeking treat-
Due to worries about losing custody, parents with mental ment for BD. This approach would branch out from the
illness are often reluctant to disclose that they have chil- identified patient to identify other affected and high-risk
dren; however, this also limits professionals’ ability to help individuals in the family tree. Primary care is another set-
the children or to offer parenting resources (Spiegelhoff & ting where many people come who would not otherwise
Ahia, 2011). consider mental health services (Mitchell, Vaze, & Rao,
If mental health services are available for offspring, par- 2009). School-based services provide a third opportunity
ents may still be resistant to their children seeing a therapist to provide education about mental health services and
(Sturges, 1977). Parents with mental illness often worry BD in particular, as well as a potential additional setting
about “passing on” their mental illness to their children and for deployment of mental health services (Weist, Nabors,
may feel guilty about any emotional or behavioral problems Myers, & Armbruster, 2000).
their children demonstrate (Goodwin & Jamison, 1990). The challenge is finding balanced ways to detect more
The suggestion that a child see a therapist, whether due to cases of BD without having an offsetting number of false
his or her mental state or to a situation brought on by the positive diagnoses. Early enthusiasm for screening programs
parent’s illness, may activate feelings of defensiveness or for many other disorders—including breast and prostate
denial. cancer—has been tempered by high rates of false alarms
and poor predictive power in many settings (Gigerenzer,
2002). No single risk factor or positive result on a screening
N U RT U R E C H A NG I NG N AT U R E :  test is discriminating enough to confirm a diagnosis of BD
HOW TO   I N T E RV E N E in isolation (Zimmerman, 2012). Instead, evidence-based
E F F E C T I V E LY TO   H E L P assessment recommends using a series of tools and risk fac-
FA M I L I E S W I T H tors in sequence, starting with low-cost tools (e.g., Algorta
B I P OL A R DI S OR DE R et  al., 2013)  and following up with additional, in-depth
evaluation when multiple findings raise the index of sus-
There are multiple ways that biology and environment picion for bipolarity (Youngstrom, Jenkins, Jensen-Doss,
transact with each other. The complexity of the relations & Youngstrom, 2012). Goals for the evaluation process
offers multiple points for intervention in the family system, should include understanding what strengths and unmet

176   •   S E C T I O N I I : I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
service needs each family member might have and then for- McAndrew, 2008; Wynaden & Orb, 2005). Despite
mulating a plan to maximize benefits while avoiding pos- these obstacles, it is valuable to pursue engagement, edu-
sible treatment-sabotaging dynamics. cation, and support of the family system. People with
From a pharmacological perspective, family assessment BD have commented on the valuable role close family or
is crucial to discern (a) who would benefit from medication, friends can play in helping them to manage their illness
(b)  who can help with medication management, (c)  what and maintain a good quality of life, but caregivers need
family processes might interfere with a sense of congru- information about ways to do this without jeopardizing
ence between treatment recommendations and personal their own health (Michalak, Yatham, Kolesar, & Lam,
beliefs, and (d)  what behaviors might directly complicate 2006; Russell & Browne, 2005). Acknowledging each
treatment. For young children, it may be essential to have family member’s needs and setting some limits defin-
the parent manage the medication; for teenagers or spouses, ing tolerable behaviors will also reduce resentment and
having the collateral person involved with the medications anger. Mental health literacy is not simply a matter of
could be a source of conflict. A  patient might decide to having knowledge; it links knowledge to the possibility
stop taking medication, or to overdose, as a result of high of action that benefits family members’ mental health
expressed negative emotion. On the positive side, support- (Jorm, 2012).
ive and empathic family members promote engagement
with treatment, and relatives who have done well on medi-
cation provide successful role models as well as clues about
C ONC LUS ION
choice of intervention, inasmuch as there may be a genetic
basis to individual differences in treatment response (Alda,
It is well established that BD is a highly heritable illness. A large
Grof, Rouleau, Turecki, & Young, 2005).
part of the heritability is due to genetic factors, although these
From a psychosocial perspective, family engagement
have proven far more complex and less tractable than initially
increases the chances of success for any intervention. High
hoped. The field has also learned that it is not a question of
expressed emotion and conflict predict earlier age of onset,
nature or nurture but rather a matter of understanding how
worse treatment response, and higher rates of relapse.
the two are intertwined. The parent has effects on the child
Psychoeducational interventions not only increase family
through biological, environmental, and social pathways that
member’s understanding of the illness but also improve
also interact with each other. Genes of risk in the parent can
engagement, enhance family processes, and mediate bet-
lead to behaviors that disrupt parenting, further altering the
ter outcomes for pharmacological interventions (Fristad
home environment and amplifying the genetic effects.
& Macpherson, 2014). Psychosocial interventions focusing
But children are not passive recipients of these influ-
on the individual are still likely to benefit from a thorough
ences. In addition to any inherited biology, they also learn
understanding of the family. There is growing evidence
from their parents, absorbing meta-rules for emotions and
that family therapy also directly improves communica-
relationships the same way they absorb language—another
tion, reduces conflict, and leads to more durable improve-
construct that runs in families without a biological basis
ments than pharmacological treatment alone (Miklowitz
for transmission. The child sometimes withdraws, some-
et  al., 2008; Miklowitz & Scott, 2009), but much work
times pushes back against the parent, and both the parent
is needed in terms of dissemination of these approaches.
and child provoke powerful responses in each other that
Family assessment also enhances prospects for early inter-
activate biological as well as cognitive systems. Decades of
vention and prevention programs in at-risk youths (Pfennig
research show how family factors convey risk and mediate
et al., 2013).
some illness processes as well as worsening the course over
Family intervention should not focus only on symp-
time. More recent work is illuminating ways that under-
tom reduction. Family members and caregivers report a
standing family processes, and helping families make sense
lack of information about how to deal with the person’s
of BD, can establish virtuous cycles that promote treatment
illness, and about the changes and losses that commonly
engagement, remission, and positive functioning.
ensue, and report feeling isolated, alone, and unsupported
(Pirkis et al., 2010; Rowe, 2012). Health professionals do Disclosure statement: Anna Van Meter and Guillermo Perez
not always have time to provide information for caregiv- Algorta have no conflicts to disclose. Dr. Eric Youngstrom
ers, and confidentiality concerns compound the difficulty has consulted with Lundbeck and Otsuka and received
of implementing collaborative family-friendly models grant support from the National Institute of Mental
(Peters, Pontin, Lobban, & Morriss, 2011; Wilkinson & Health.

T he A pple and the T ree   •   17 7


R E F E R E NC E S encoding of emotional faces in pediatric bipolar disorder. Bipolar
Disorders, 9, 679–692.
Dougherty, L. R., Tolep, M. R., Smith, V. C., & Rose, S. (2013). Early
Abidin, R. R. (1992). The determinants of parenting behavior. Journal
exposure to parental depression and parenting:  Associations with
of Clinical Child Psychology, 21, 407–412.
young offspring’s stress physiology and oppositional behavior. Jour-
Alda, M., Grof, P., Rouleau, G. A., Turecki, G., & Young, L. T. (2005).
nal of Abnormal Child Psychology, 41(8), 1299–1310.
Investigating responders to lithium prophylaxis as a strategy for
Drancourt, N., Etain, B., Lajnef, M., Henry, C., Raust, A., Cochet,
mapping susceptibility genes for bipolar disorder. Progress in Neuro-
B., . . . Bellivier, F. (2013). Duration of untreated bipolar disor-
psychopharmacology & Biological Psychiatry, 29, 1038–1045.
der: Missed opportunities on the long road to optimal treatment.
Algorta, G. P., Youngstrom, E. A., Phelps, J., Jenkins, M. M., Young-
Acta Psychiatrica Scandinavica, 127, 136–144.
strom, J. K., & Findling, R. L. (2013). An inexpensive family index
Duffy, A., Alda, M., Hajek, T., & Grof, P. (2009). Early course of bipo-
of risk for mood issues improves identification of pediatric bipolar
lar disorder in high-risk offspring: Prospective study. British Journal
disorder. Psychological Assessment, 25, 12–22.
of Psychiatry, 195, 457–458.
Bella, T., Goldstein, T., Axelson, D., Obreja, M., Monk, K., Hickey,
Duffy, A., Alda, M., Trinneer, A., Demidenko, N., Grof, P., & Goodyer,
M. B., . . . Birmaher, B. (2011). Psychosocial functioning in offspring
I.  M. (2007). Temperament, life events, and psychopathology
of parents with bipolar disorder. Journal of Affective Disorders, 133,
among the offspring of bipolar parents. European Child & Adoles-
204–211.
cent Psychiatry, 16, 222–228.
Belsky, J. (1984). The determinants of parenting: a process model. Child
Ehlers, C. L., Frank, E., & Kupfer, D. J. (1988). Social zeitgebers and
Development, 55, 83–96.
biological rhythms: A unified approach to understanding the etiol-
Birmaher, B., Axelson, D., Monk, K., Kalas, C., Goldstein, B., Hickey,
ogy of depression. Archives of General Psychiatry, 45, 948–952.
M.  B.,  . 
. . 
Brent, D. (2009). Lifetime psychiatric disorders in
Essex, M. J., Shirtcliff, E. A., Burk, L. R., Ruttle, P. L., Klein, M. H.,
school-aged offspring of parents with bipolar disorder:  the Pitts-
Slattery, M.  J., . . . Armstrong, J.  M. (2011). Influence of early life
burgh Bipolar Offspring study. Arch Gen Psychiatry, 66(3), 287–296.
stress on later hypothalamic-pituitary-adrenal axis functioning and
Birmaher, B., Goldstein, B. I., Axelson, D. A., Monk, K., Hickey, M. B.,
its covariation with mental health symptoms: A study of the allo-
Fan, J., . . . Kupfer, D.  J. (2013). Mood lability among offspring of
static process from childhood into adolescence. Development and
parents with bipolar disorder and community controls. Bipolar Dis-
Psychopathology, 23, 1039–1058.
orders, 15, 253–263.
Evans, L., Akiskal, H., Keck, P., McElroy, S., Sadovnick, A. D., Remick,
Brennan, P. A., Hammen, C., Andersen, M. J., Bor, W., Najman, J. M., &
R., & Kelsoe, J. (2005). Familiality of temperament in bipolar dis-
Williams, G. M. (2000). Chronicity, severity, and timing of mater-
order: support for a genetic spectrum. Journal of Affective Disorders,
nal depressive symptoms: Relationships with child outcomes at age
85, 153–168.
5. Developmental Psychology, 36, 759–766.
Evans, L.  M., Akiskal, H.  S., Greenwood, T.  A., Nievergelt, C.  M.,
Calabrese, J. R., Hirschfeld, R., Reed, M., Davies, M. A., Frye, M. A.,
Keck, P. E., McElroy, S. L., . . . Kelsoe, J. R. (2008). Suggestive link-
Keck, P. E. J., . . . Wagner, K. D. (2003). Impact of bipolar disorder
age of a chromosomal locus on 18p11 to cyclothymic temperament
on a U.S. community sample. The Journal of Clinical Psychiatry, 64,
in bipolar disorder families. American Journal of Medical Genetics,
425–432.
147B, 326–332.
Calam, R. (2012). Parenting and the emotional and behavioural adjust-
Frank, E., Swartz, H.  A., & Kupfer, D.  J. (2000). Interpersonal and
ment of young children in families with a parent with bipolar disor-
social rhythm therapy: Managing the chaos of bipolar disorder. Bio-
der. Behavioural and Cognitive Psychotherapy, 40, 425–437.
logical Psychiatry, 48, 593–604.
Cerel, J., Fristad, M.  A., Weller, E.  B., & Weller, R.  A. (1999).
Freeman, A.  J., Youngstrom, E.  A., Michalak, E., Siegel, R., Meyers,
Suicide-bereaved children and adolescents: A controlled longitudi-
O. I., & Findling, R. L. (2009). Quality of life in pediatric bipolar
nal examination. Journal of the American Academy of Child & Ado-
disorder. Pediatrics, 123, e446–e452.
lescent Psychiatry, 38, 672–679.
Fristad, M. A., & Macpherson, H. A. (2014). Evidence-based psycho-
Chang, K., Steiner, H., & Ketter, T. (2003). Studies of offspring of par-
social treatments for child and adolescent bipolar spectrum dis-
ents with bipolar disorder. American Journal of Medical Genetics,
orders. Journal of Clinical Child & Adolescent Psychology, 43(3),
123C, 26–35.
339–355.
Chang, K. D., Blasey, C., Ketter, T. A., & Steiner, H. (2001). Family
Garcia-Amador, M., de la Serna, E., Vila, M., Romero, S., Valenti, M.,
environment of children and adolescents with bipolar parents. Bipo-
Sanchez-Gistau, V., . . . Castro-Fornieles, J. (2013). Parents with
lar Disorders, 3, 73–78.
bipolar disorder: Are disease characteristics good predictors of psy-
Chang, K. D., Steiner, H., & Ketter, T. A. (2000). Psychiatric phenom-
chopathology in offspring? European Psychiatry, 28, 240–246.
enology of child and adolescent bipolar offspring. Journal of the
Gigerenzer, G. (2002). Calculated risks:  How to know when numbers
American Academy of Child & Adolescent Psychiatry, 39, 453–460.
deceive you. New York: Simon & Schuster.
Christoffersen, M. N., Poulsen, H. D., & Nielsen, A. (2003). Attempted
Goldstein, T. R., Birmaher, B., Axelson, D., Goldstein, B. I., Gill, M. K.,
suicide among young people: Risk factors in a prospective register
Esposito-Smythers, C., .  .  . Keller, M. (2009). Psychosocial func-
based study of Danish children born in 1966. Acta Psychiatrica
tioning among bipolar youth. Journal of Affective Disorders, 114,
Scandinavica, 108, 350–358.
174–183.
Coryell, W., Scheftner, W., Keller, M., Endicott, J., Maser, J., & Kler-
Gonda, X., Rihmer, Z., Zsombok, T., Bagdy, G., Akiskal, K.  K., &
man, G. L. (1993). The enduring psychosocial consequences of mania
Akiskal, H. S. (2006). The 5HTTLPR polymorphism of the sero-
and depression. American Journal of Psychiatry, 150, 720–727.
tonin transporter gene is associated with affective temperaments as
Davenport, Y.  B., & Adland, M.  L. (1982). Postpartum psychoses in
measured by TEMPS-A. Journal of Affective Disorders, 91, 125–131.
female and male bipolar manic-depressive patients. American Jour-
Goodman, S.  H., & Gotlib, I.  H. (1999). Risk for psychopathology
nal of Orthopsychiatry, 52, 288–297.
in the children of depressed mothers:  A  developmental model for
De Los Reyes, A., Goodman, K. L., Kliewer, W., & Reid-Quinones, K.
understanding mechanisms of transmission. Psychological Review,
(2008). Whose depression relates to discrepancies? Testing relations
106, 458–490.
between informant characteristics and informant discrepancies
Goodwin, F., & Jamison, K. (1990). Manic-depressive illness (1st ed.).
from both informants’ perspectives. Psychological Assessment, 20,
New York: Oxford University Press.
139–149.
Goodwin, F. K., & Jamison, K. R. (2007). Manic-depressive illness (2nd
Dickstein, D., Rich, B., Roberson-Nay, R., Berghorst, L., Vinton,
ed.). New York: Oxford University Press.
D., Pine, D., & Leibenluft, E. (2007). Neural activation during

17 8   •   S E C T I O N I I : I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
Goodwin, R. D., Beautrais, A. L., & Fergusson, D. M. (2004). Famil- Lovejoy, M.  C., Graczyk, P.  A., O’Hare, E., & Neuman, G. (2000).
ial transmission of suicidal ideation and suicide attempts:  Evi- Maternal depression and parenting behavior:  A  meta-analytic
dence from a general population sample. Psychiatry Research, 126, review. Clinical Psychology Review, 20, 561–592.
159–165. Luecken, L.  J., & Appelhans, B.  M. (2006). Early parental loss and
Gottesman, I.  I., Laursen, T.  M., Bertelsen, A., & Mortensen, P.  B. salivary cortisol in young adulthood: The moderating role of family
(2010). Severe mental disorders in offspring with 2 psychiatrically environment. Development and Psychopathology, 18, 295–308.
ill parents. Archives of General Psychiatry, 67, 252–257. Lytton, H. (1990a). Child and parent effects in boys’ conduct disor-
Gould, M., Greenberg, T., Velting, D., & Shaffer, D. (2003). Youth sui- der: A reinterpretation. Developmental Psychology, 26, 683–697.
cide risk and preventive interventions: A review of the past 10 years. Lytton, H. (1990b). Child effects—Still unwelcome? Response to
Journal of the American Academy of Child & Adolescent Psychiatry, Dodge and Wahler. Developmental Psychology, 26, 705–709.
42, 386–405. Marks, M. N. (1992). Contribution of psychological and social factors
Green, M. J., Cahill, C. M., & Malhi, G. S. (2007). The cognitive and to psychotic and non-psychotic relapse after childbirth in women
neurophysiological basis of emotion dysregulation in bipolar disor- with previous histories of affective disorder. Journal of Affective Dis-
der. Journal of Affective Disorders, 103, 29–42. orders, 24, 253–263.
Greenwood, T. A., Akiskal, H. S., Akiskal, K. K., & Kelsoe, J. R. (2012). Maybery, D., Ling, L., Szakacs, E., & Reupert, A. (2005). Children of a
Genome-wide association study of temperament in bipolar disorder parent with a mental illness: Perspectives on need. Advances in Men-
reveals significant associations with three novel loci. Biological Psy- tal Health, 4, 78–88.
chiatry, 72, 303–310. Meinhard, N. (2013). The role of estrogen in bipolar disorder: A review.
Gunnar, M. R., & Vazquez, D. (2006). Stress neurobiology and devel- Nordic Journal of Psychiatry, 1–7.
opmental psychopathology. In D. Cicchetti & D. J. Cohen (Eds.), Melhem, N. M., Walker, M., Moritz, G., & Brent, D. A. (2008). Ante-
Developmental psychopathology:  Developmental neuroscience (2nd cedents and sequelae of sudden parental death in offspring and sur-
ed., Vol. 2, pp. 533–577). Hoboken, NJ: John Wiley. viving caregivers. Archives of Pediatrics & Adolescent Medicine, 162,
Hammen, C. (2009). Adolescent depression:  Stressful interpersonal 403–410.
contexts and risk for recurrence. Current Directions in Psychological Merikangas, K., He, J., Burstein, M., Swanson, S., Avenevoli, S., Cui,
Science, 18, 200–204. L., . . . Swendsen, J. (2010). Lifetime prevalence of mental disor-
Handley, C., Farrell, G.  A., Josephs, A., Hanke, A., & Hazelton, M. ders in U.S. adolescents: Results from the National Comorbidity
(2001). The Tasmanian children’s project:  The needs of children Survey Replication–Adolescent Supplement (NCS-A). Journal
with a parent/carer with a mental illness. Australian & New Zea- of the American Academy of Child & Adolescent Psychiatry, 49,
land Journal of Mental Health Nursing, 10, 221–228. 980–989.
Hinshaw, S.  P. (Ed.). (2008). Breaking the silence. New  York:  Oxford Mesman, E., Nolen, W.  A., Reichart, C.  G., Wals, M., & Hillegers,
University Press. M.  H. (2013). The Dutch Bipolar Offspring Study:  12-year
Hodgins, S., Faucher, B., Zarac, A., & Ellenbogen, M. (2002). Children follow-up. American Journal of Psychiatry, 170, 542–549.
of parents with bipolar disorder: A population at high risk for major Michalak, E.  E., Yatham, L.  N., Kolesar, S., & Lam, R.  W. (2006).
affective disorders. Child & Adolescent Psychiatric Clinics of North Bipolar disorder and quality of life: A patient-centered perspective.
America, 11, 533–553. Quality of Life Research, 15, 25–37.
Holsboer, F. (2000). The corticosteroid receptor hypothesis of depres- Miklowitz, D. J. (2004). The role of family systems in severe and recur-
sion. Neuropsychopharmacology, 23, 477–501. rent psychiatric disorders: A developmental psychopathology view.
Inoff-Germain, G., Nottelmann, E.  D., & Radke-Yarrow, M. (1992). Development and Psychopathology, 16, 667–688.
Evaluative communications between affectively ill and well moth- Miklowitz, D., & Johnson, B. (2009). Social and familial factors in the
ers and their children. Journal of Abnormal Child Psychology, 20, course of bipolar disorder:  Basic processes and relevant interven-
189–212. tions. Clinical Psychology: Science and Practice, 16, 281–296.
James, A. (2011). Early-onset bipolar disorder. In Child psychology and Miklowitz, D. J., Axelson, D. A., Birmaher, B., George, E. L., Taylor,
psychiatry (pp. 210–216). Hoboken, NJ: John Wiley. D. O., Schneck, C. D., . . . Brent, D. A. (2008). Family-focused treat-
Janet Kuramoto, S., Brent, D. A., & Wilcox, H. C. (2009). The impact ment for adolescents with bipolar disorder: Results of a 2-year ran-
of parental suicide on child and adolescent offspring. Suicide & Life domized trial. Archives of General Psychiatry, 65, 1053–1061.
Threatening Behaviors, 39, 137–151. Miklowitz, D. J., & Scott, J. (2009). Psychosocial treatments for bipo-
Jones, S. H., & Bentall, R. P. (2008). A review of potential cognitive and lar disorder: Cost-effectiveness, mediating mechanisms, and future
environmental risk markers in children of bipolar parents. Clinical directions. Bipolar Disorders, 11, 110–122.
Psychology Review, 28, 1083–1095. Mitchell, A. J., Vaze, A., & Rao, S. (2009). Clinical diagnosis of depres-
Jorm, A. F. (2012). Mental health literacy: Empowering the community sion in primary care: A meta-analysis. Lancet, 374, 609–619.
to take action for better mental health. American Psychologist, 67, Mordoch, E., & Hall, W.  A. (2008). Children’s perceptions of living
231–243. with a parent with a mental illness: Finding the rhythm and main-
Kessing, L.  V., Agerbo, E., & Mortensen, P.  B. (2004). Major stress- taining the frame. Qualitative Health Research, 18, 1127–1144.
ful life events and other risk factors for first admission with mania. Mortensen, P. B., Pedersen, C. B., Melbye, M., Mors, O., & Ewald, H.
Bipolar Disorders, 6, 122–129. (2003). Individual and familial risk factors for bipolar affective dis-
Knight, D.  K., & Wallace, G. (2003). Where are the children? An orders in Denmark. Archives of General Psychiatry, 60, 1209–1215.
examination of children’s living arrangements when mothers enter Murray, G. (2010). Circadian rhythms and sleep in bipolar disorder.
residential drug treatment. Journal of Drug Issues, 33, 305–324. Bipolar disorders, 12, 459–472.
Lapalme, M., Hodgins, S., & LaRoche, C. (1997). Children of parents Park, J. M., Solomon, P., & Mandell, D. S. (2006). Involvement in the
with bipolar disorder: A metaanalysis of risk for mental disorders. child welfare system among mothers with serious mental illness.
Canadian Journal of Psychiatry, 42, 623–631. Psychiatric Services, 57, 493–497.
Livesley, W.  J. (1995). The DSM-IV personality disorders. Pauls, D. L., Morton, L. A., & Egeland, J. A. (1992). Risks of affective
New York: Guilford Press. illness among first-degree relatives of bipolar I old-order Amish pro-
Ljung, T., Lichtenstein, P., Sandin, S., D’Onofrio, B., Runeson, B., bands. Arch Gen Psychiatry, 49(9), 703–708.
Långström, N., & Larsson, H. (2013). Parental schizophrenia and Peters, S., Pontin, E., Lobban, F., & Morriss, R. (2011). Involving rela-
increased offspring suicide risk:  Exploring the causal hypothesis tives in relapse prevention for bipolar disorder: A multi-perspective
using cousin comparisons. Psychological Medicine, 43, 581–590. qualitative study of value and barriers. BMC Psychiatry, 11, 172.

T he A pple and the T ree   •   17 9


Pfennig, A., Correll, C.  U., Marx, C., Rottmann-Wolf, M., Meyer, Sturges, J. S. (1977). Talking with children about mental illness in the
T. D., Bauer, M., & Leopold, K. (2013). Psychotherapeutic interven- family. Health & Social Work, 2, 87–109.
tions in individuals at risk of developing bipolar disorder: A system- Thomas, L., & Kalucy, R. (2003). Parents with mental illness: Lacking
atic review. Early Intervention Psychiatry, 8(1), 3-11. motivation to parent. International Journal of Mental Health Nurs-
Pirkis, J., Burgess, P., Hardy, J., Harris, M., Slade, T., & Johnston, A. ing, 12, 153–157.
(2010). Who cares? A profile of people who care for relatives with a Townsend, J. (2012). Emotion processing and regulation in bipolar dis-
mental disorder. Australian & New Zealand Journal of Psychiatry, order: A review. Bipolar Disorders, 14, 326–339.
44, 929–937. Tsuchiya, K. J., Agerbo, E., & Mortensen, P. B. (2005). Parental death
Plomin, R. (1995). Genetics and children’s experiences in the family. and bipolar disorder: A robust association was found in early mater-
Journal of Child Psychology and Psychiatry, 36, 33–68. nal suicide. Journal of Affective Disorders, 86, 151–159.
Reichart, C. G., van der Ende, J., Hillegers, M. H., Wals, M., Bongers, Van Meter, A., Moreira, A., & Youngstrom, E. (2011). Meta-analysis of
I. L., Nolen, W. A., . . . Verhulst, F. C. (2007). Perceived parental rear- epidemiologic studies of pediatric bipolar disorder. Journal of Clini-
ing of bipolar offspring. Acta Psychiatrica Scandinavica, 115, 21–28. cal Psychiatry, 72, 1250–1256.
Rich, B. A., Schmajuk, M., Perez-Edgar, K. E., Pine, D. S., Fox, N. A., Vandeleur, C., Rothen, S., Gholam-Rezaee, M., Castelao, E., Vidal, S.,
& Leibenluft, E. (2005). The impact of reward, punishment, and Favre, S., . . . Preisig, M. (2012). Mental disorders in offspring of par-
frustration on attention in pediatric bipolar disorder. Biological Psy- ents with bipolar and major depressive disorders. Bipolar Disorders,
chiatry, 58, 532–539. 14, 641–653.
Riebschleger, J. (2004). Good days and bad days:  The experiences of Wang, P. S., Lane, M., Olfson, M., Pincus, H. A., Wells, K. B., & Kes-
children of a parent with a psychiatric disability. Psychiatric Reha- sler, R.  C. (2005). Twelve-month use of mental health services in
bilitation Journal, 28, 25–31. the United States: Results from the National Comorbidity Survey
Rowe, J. (2012). Great expectations: A systematic review of the litera- Replication. Archives of General Psychiatry, 62, 629–640.
ture on the role of family carers in severe mental illness, and their Weist, M.  D., Nabors, L.  A., Myers, C.  P., & Armbruster, P. (2000).
relationships and engagement with professionals. Journal of Psychi- Evaluation of expanded school mental health programs. Commu-
atric Mental Health Nursing, 19, 70–82. nity Mental Health Journal, 36, 395–411.
Russell, S. J., & Browne, J. L. (2005). Staying well with bipolar disorder. Wilkinson, C., & McAndrew, S. (2008). “I’m not an outsider, I’m his
Australian & New Zealand Journal of Psychiatry, 39, 187–193. mother!” A  phenomenological enquiry into carer experiences of
Rutter, M., & Quinton, D. (1984). Parental psychiatric disorder: Effects exclusion from acute psychiatric settings. International Journal of
on children. Psychological Medicine, 14, 853–880. Mental Health Nursing, 17, 392–401.
Sands, R. G., Koppelman, N., & Solomon, P. (2004). Maternal custody Wynaden, D., & Orb, A. (2005). Impact of patient confidentiality on
status and living arrangements of children of women with severe carers of people who have a mental disorder. International Journal of
mental illness. Health & Social Work, 29, 317–325. Mental Health Nursing, 14, 166–171.
Sharma, V. (2012). Pregnancy and bipolar disorder: A systematic review. Youngstrom, E. A., Jenkins, M. M., Jensen-Doss, A., & Youngstrom,
Journal of Clinical Psychiatry, 73, 1447–1455. J.  K. (2012). Evidence-based assessment strategies for pediatric
Singh, M.  K., DelBello, M.  P., Stanford, K.  E., Soutullo, C., bipolar disorder. Israel Journal of Psychiatry & Related Sciences,
McDonough-Ryan, P., McElroy, S. L., & Strakowski, S. M. (2007). 49, 15–27.
Psychopathology in children of bipolar parents. J Affect Disord, Youngstrom, E.  A., Loeber, R., & Stouthamer-Loeber, M. (2000).
102(1–3), 131–136. Patterns and correlates of agreement between parent, teacher,
Smoller, J.  W., & Finn, C.  T. (2003). Family, twin, and adoption and male adolescent ratings of externalizing and internaliz-
studies of bipolar disorder. American Journal of Medical Genetics, ing problems. Journal of Consulting and Clinical Psychology, 68,
123C, 48–58. 1038–1050.
Spiegelhoff, S. F., & Ahia, C. E. (2011). Impact of parental severe men- Yuan, C., Yu, S.  Y., Li, Z., Huang, J., Qian, Y., & Fang, Y. (2012).
tal illness:  Ethical and clinical issues for counselors. The Family P-222—The 5HTTLPR is associated with bipolar disorder and
Journal, 19, 389–395. affective temperaments as measured by TEMPS-A in Chinese pop-
Spitzer, R. L. (2002). DSM-IV-TR casebook: A learning companion to the ulation. European Psychiatry, 27(Suppl. 1), 1.
Diagnostic and Statistical Manual of Mental Disorders, fourth edi- Zimmerman, M. (2012). Misuse of the Mood Disorders Questionnaire
tion, text revision (1st ed.). Washington, DC: American Psychiatric as a case-finding measure and a critique of the concept of using a
Publishing screening scale for bipolar disorder in psychiatric practice. Bipolar
Steptoe, A., van Jaarsveld, C. H., Semmler, C., Plomin, R., & Wardle, J. Disorders, 14, 127–134.
(2009). Heritability of daytime cortisol levels and cortisol reactivity Zuckerman, M. (1999). Vulnerability to psychopathology:  A  biosocial
in children. Psychoneuroendocrinology, 34, 273–280. model. Washington, DC: American Psychological Association.

18 0   •   S E C T I O N I I : I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
15.
THE EMERGING GENETICS OF BIPOLAR DISOR DER
Roy H. Perlis

INTRODUCTION when a sibling has bipolar disorder, the other sibling’s rela-
tive risk is ~7.9 (95% CI 7.1–8.8).
Bipolar disorder often recurs within families. This obser- Several other characteristics of bipolar disorder famili-
vation, familiar to most clinicians and dating back many ality are apparent in studies with these designs. First and
decades, prompted aggressive efforts to identify the specific foremost, bipolar disorder does not appear to follow a
genetic variations that may confer risk for this disorder. Mendelian pattern of inheritance, like recessive or domi-
These efforts have begun to bear fruit, with at least five dis- nant transmission. At minimum, this suggests incomplete
tinct regions of the genome demonstrated to contribute to penetrance—unlike some rare genetic diseases, knowing
bipolar disorder liability and evidence that hundreds more parental status does not allow reliable prediction of off-
await discovery. While these discoveries have immediate spring status. Analysis of pedigrees, in conjunction with
implications for investigating the neurobiology of bipolar additional phenotypic data, also strongly suggests that
disorder, they have not yet reached the point of being clini- other recognized inheritance patterns are not present in
cally actionable. most bipolar disorder pedigrees. For example, there is not
strong evidence of genetic anticipation, as observed in
Huntington’s disease and other trinucleotide repeat dis-
FA M I LY A N D T W I N S T U DI E S orders, in which age at onset tends to decrease with sub-
OF   B I P OL A R DI S OR DE R sequent generations; likewise, there is not strong evidence
of maternal inheritance as is observed in mitochondrial
The standard approach to estimating how much bipolar diseases. In general, mathematical modeling was useful
disorder runs in families examines constellations of fam- in excluding certain genetic models of disease (Craddock,
ily members—at minimum, a child and a parent but often Khodel, Van Eerdewegh, & Reich, 1995).
extended pedigrees over multiple generations. These data Second, in addition to increasing the risk for bipolar
allow an estimate of recurrence risk based on family dis- disorder, a bipolar disorder family member increases the
tance from an affected individual. For example, lambda’s risk for other psychiatric illness, including major depres-
refers to the recurrence risk for a sibling, with whom an sive disorder and schizophrenia (Lichtenstein et al., 2009).
individual will share (on average) ~50% of DNA. For example, risk for schizophrenia is elevated in offspring
Initial estimates supported a lambda’s of ~8—that is, of individuals with bipolar disorder, with relative risk esti-
if a parent has bipolar disorder, the child’s relative risk of mated to be 2.4 (95% CI 2.1–2.6). Table 15.1 illustrates the
bipolar disorder is eight-fold greater than average for that relative risk to first-degree relatives (offspring or siblings) of
population (Lichtenstein et al., 2009; Smoller & Finn; individuals affected with bipolar disorder or schizophrenia.
2003). In lieu of the laborious task of identifying and inter- Importantly, a first-degree family member of a patient with
viewing each family member, a recent investigation made bipolar disorder is more likely to have a diagnosis of major
use of health registry data collected in Sweden to derive spe- depression than bipolar disorder. While the relative risk
cific risks for each family relationship. Notably, these values for bipolar disorder is greater than that for major depres-
are not identical for each first-degree relationship. When sive disorder, because the prevalence of major depressive
one parent has bipolar disorder, the child’s relative risk disorder is so much greater (~15% vs. 5%), the absolute risk
for bipolar is ~6.4 (95% confidence interval [CI] 5.9–7.1); for major depression exceeds that of bipolar disorder. This

181
Table 15 .1.  R ECU R R ENCE R ISK S FOR SCHIZOPHR ENI A A N D BIPOL A R DISOR DER A N D CO-MOR BIDIT Y
I N PA R ENT-OFFSPR I NG A ND SI BLI NGS

R ISK FOR SCHIZO- R ISK FOR BIPOLAR R ISK FOR SCHIZO- R ISK FOR BIPOLAR
PHR ENI A W HEN DISOR DER W HEN PHR ENI A W HEN DISOR DER W HEN
PROBAND HAS PROBAND HAS PROBAND HAS PROBAND HAS
SCHIZOPHR ENI A BIPOLAR DISOR DER BIPOLAR DISOR DER SCHIZOPHR ENI A
R ELATION TO R ELATIVE 95% CI R ELATIVE 95% CI R ELATIVE 95% CI R ELATIVE 95% CI
PROBAND PROBAND R ISK R ISK R ISK R ISK

Parent Offspring 9.9 8.5–11.6 6.4 5.9–7.1 2.4 2.1–2.6 5.2 4.4–6.2
Sibling Sibling 9 8.1–9.9 7.9 7.1–8.8 3.9 3.4–4.4 3.7 3.2–4.2

Sibling Maternal 3.6 2.3–5.5 4.5 2.7–7.4 1.4 0.7–2.6 1.2 0.6–2.4
half-sibling

Sibling Paternal 2.7 1.9–3.8 2.4 1.4–4.1 1.6 1.0–2.7 2.2 1.3–3.8
half-sibling
Adoptive relationships

Biological Adopted away 13.7 6.1–30.8 4.3 2.0–9.5 4.5 1.8–10.9 6 2.3–15.2
parent offspring
Sibling Adopted away 7.6 0.7–87.8 — 3.9 0.2–63.3 5 0.3–79.9
biological sibling

Adoptive Adoptee — 1.3 0.5–3.6 1.5 0.7–3.5 —


parent

Sibling Non-biological 1.3 0.1–15.1 — — — 2 0.1–37.8


sibling

crucial point is often missed in clinical practice, where a However, familial is not the same as genetic—specific char-
report of family history of bipolar disorder biases the clini- acteristics may be run in families but not be based on inher-
cian to diagnose bipolar disorder (Perlis, 2005). ited genetic variation. A prime example is wealth: parental
Another key point to emerge from population and wealth is predictive of wealth in offspring, but no genetic
cohort studies of bipolar disorder is that despite the famili- predisposition to wealth has yet been identified.
ality, most cases of disease are sporadic. That is, most indi- To understand the extent to which risk is a result of
viduals with bipolar disorder will not have a first-degree shared genetic variation, the twin-pair design has been
family member with bipolar disorder (Perlis, Dennehy, applied. This strategy takes advantage of the difference
et al., 2009). between monozygotic twins, who share nearly all of their
It should also be noted that some family studies also DNA, and dizygotic twins, who share on average 50% like
suggest cotransmission of risk with other medical disorders. any other pair of siblings. By contrasting the concordance
For example, one study indicated that migraine may seg- for disease between monozygotic and dizygotic twins, one
regate with affective illness in some families (Merikangas, can partition out the shared genetic risk from, for example,
Merikangas, & Angst, 1993). Another report suggested a environmental risk.
similar effect for panic disorder. The evidence for such con- In large twin-pair studies, when one monozygotic twin
transmission is not strong and is not consistent across stud- has bipolar disorder, the other has approximately a 40% to
ies. Still, these investigations provide hypotheses about the 70% likelihood of having that illness (Craddock & Jones,
pleiotropic effects of bipolar disorder risk genes that may 1999). (As family studies would predict, the co-twin also
be investigated as specific risk variants are identified. Given has an elevated likelihood of other disorders.) For dizygotic
the large number of brain genes also expressed elsewhere, it twins, the likelihood is approximately 5% to 10%. These
would not be surprising to find subtle non-brain manifesta- data allow estimates that up to ~90% of the risk for bipo-
tions of bipolar disorder in some individuals. lar disorder is heritable. For example, a large registry-based
Taken together, studies of families with bipolar disorder study of more than 19,000 twin pairs in Finland reported a
probands strongly indicate the familiality of this disease. heritability of 93% (Kieseppa, Partonen, Haukka, Kaprio, &

18 2   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
Lonnqvist, 2004). Using an alternative non-twin meth- to focus on a small subset of genes with strong prior prob-
odology, heritability was estimated at 62%, similar to that abilities. This advantage was also a substantial disadvan-
of schizophrenia (67%) and roughly twice that of major tage, however: with anywhere from 300,000 to as many as
depressive disorder (32%; Wray & Gottesman, 2012). 2.5  million single nucleotide polymorphisms on an array,
Another approach to heritability examines circum- the multiple testing problem became acute. A  rigorous
stances where twins are adopted and thus exposed to dif- approach to such multiple testing mandated correction
ferent environmental risk. These studies also support the for 1 million tests, a p value of 5 × 10–8, as a threshold for
heritability of bipolar disorder—illness in the biological so-called genome-wide significance.
parent, but not the adoptive parent, is predictive of bipolar While initial single-cohort studies using genome-
risk (Lichtenstein et al., 2009). wide approaches did not identify bipolar disorder risk
An underappreciated aspect of bipolar studies such genes at a genome-wide level of significance, investiga-
as these is the implication for the bipolar disorder pheno- tors quickly recognized the need to combine cohorts to
type itself. Despite loud critiques of the Diagnostic and improve statistical power to detect association. Under
Statistical Manual of Mental Disorders and structured the aegis of the Psychiatric Genomics Consortium, the
interviews in general, the fact remains that these criteria size of such studies has rapidly grown.
identify a strongly familial disorder with heritability simi- Among the first bipolar disorder liability genes was the
lar to or greater than many nonpsychiatric disorders. Thus L-type calcium channel subunit, CACNA1C (Ferreira et al.,
the predictive validity of these diagnostic criteria actually 2008). This association has persisted as sample sizes have
compares quite favorably to criteria used for other diseases. increased, while another initial association with ANK3
has been less consistently supported (Green, Hamshere,
et al., 2013).
C OM MON VA R I A N T By late 2012, at least five loci associated with bipolar
A S S O C I AT ION S T U DI E S disorder had been reported (Green, Hamshere, et al., 2013;
Psychiatric GWAS Consortium Bipolar Disorder Working
Buoyed by strong evidence of heritability, initial investi- Group, 2011). In addition to CACNA1C, these included
gations of bipolar disorder focused on so-called candidate variants in ODZ4, NCAN14, and SYNE1 (Green, Grozeva,
genes, primarily genes coding for proteins implicated in et al., 2013). As sample sizes continue to grow, additional
hypothesized models of bipolar disorder liability. Beginning loci are likely to emerge; for example, a recent report of
with neurotransmitter genes, which had known functional nearly 25,000 bipolar cases included genome-wide evidence
variants that were inexpensive to genotype, such stud- for ADCY2 and an intergenic region on 6q16.1 (Muhleisen
ies proceeded to more creative candidate genes, including et al., 2014).
those related to circadian rhythms and those related to the With progress in identifying individual loci, there have
putative mechanisms of action of lithium, a gold-standard been efforts to integrate these loci into pathways that might
treatment for this disorder. facilitate biological investigations and therapeutic develop-
Unfortunately, such studies failed to converge on ment (Holmans et  al., 2009). In particular, such analyses
convincing evidence of association for any of these loci strongly implicate other calcium channel subunit genes
(Seifuddin et  al., 2012). In hindsight, this lack of success in bipolar disorder risk. In addition, a recent analysis of
is perhaps not surprising, given the limited understanding genome-wide data suggested involvement of diverse signal-
of the biology of this disease that motivated such studies, ing pathways related to corticotropin-releasing hormone,
the relatively modest ability to capture genetic variation beta-adrenergic function, phospholipase C, and glutamate
in these studies, and the limited statistical power to detect receptors, as well as endothelin 1 and cardiac hypertrophy
smaller effects. In essence, such investigations looked for (Nurnberger et al., 2014). Notably, some genes in these path-
lost keys under the lightpost, focusing on genes that were ways were differentially expressed in dorsolateral prefrontal
well studied and straightforward to characterize. cortex, one brain region suggested to be affected in bipolar
Advances in genotyping technology facilitated a new disorder. Pathway analysis has also been applied directly
generation of investigations, which sought to examine to try to identify critical cell types associated with illness;
common variation across the genome rather than simply in for example, an analysis of genome-wide association study
particular genes of interest. Such studies had the substan- data suggested involvement of oligodendrocytes in schizo-
tial advantage of being unbiased—that is, they opened the phrenia, while multiple glial pathways were associated with
door to truly new discoveries rather than being constrained bipolar disorder (Duncan et al., 2014). Still, such analyses

T he E merging G enetics of B ipolar D isorder   •   183


must be regarded as hypothesis-generating, indicating path- during development, so important effects may require
ways or brain regions to be prioritized for further study. examining specific stages of neurodevelopment. Emerging
In fact, experience with CACNA1C underscores the approaches that allow specific genetic variants to be intro-
key point that even if individual genetic variations are not duced into cells should accelerate efforts to study the impact
clinically actionable because of modest effect sizes, they of a given variant.
may still point the way to important neurobiological inves-
tigation. In general, much is known about the physiology
of calcium channels; Figure  15.1 illustrates the subunit R A R E VA R I A N T
structure of voltage-gated calcium channels. The functional A S S O C I AT ION S T U DI E S
implications of individual CACNA1C common variants
have not been established—that is, they do not obviously or An alternate hypothesis, albeit not one mutually exclusive
consistently impact expression levels, truncate the protein from the common-variant model, suggests that bipolar dis-
prematurely, or change 3-dimensional protein structure, order is the result of rare genetic variation that may be more
for example. In other words, are the resulting L-type cal- likely to have large effects and thus be highly penetrant.
cium channels underexpressed, overexpressed, or expressed Such variants might be particularly apparent from the
at normal levels but with some gain or loss of function? study of trios—offspring with two unaffected parents—if
As the implications of CACNA1C variants are studied the variants are de novo (i.e., not present in the parents).
at a molecular level, a growing number of studies investi- One form of rare variation is referred to as copy-number
gate the functional implications of these variants in vivo. variation (CNV). This includes deletions or duplications of
For example, one study suggested differences in process- specific regions of chromosome, ranging from relatively
ing facial expressions in individuals with bipolar disorder small to quite large. These variations can be detected by
carrying a risk variant (Dima et  al., 2013). Building on a some genotyping arrays, particularly more recent ones that
long history of investigation of calcium channel antagonists are designed with probes to detect the number of copies
primarily in bipolar mania (Casamassima, Hay, Benedetti, of a particular region present in a sample; they can also be
Lattanzi, Cassano, & Perlis, 2010), which might correct identified using whole-genome or whole-exome sequencing
a gain of function, a pilot study suggested efficacy for the approaches, discussed in more detail later.
calcium channel antagonist isradipine in bipolar depression Copy-number variations have advantages in that their
(Ostacher et al., 2014). functional implications are clearer than many common
More generally, as a growing number of genetic varia- variants. Deleting an entire gene is likely to decrease expres-
tions are associated with bipolar disorder liability, more sion of that gene’s product; deleting an exon or group of
efficient means of understanding the impact of these vari- exons may cause expression of an altered, dysfunctional
ants at a molecular level will be required. These strategies protein. Conversely, duplicating a gene may cause increase
may include examining patterns of expression in brain in expression. Thus extrapolating from a variation to a
regions or cell types relevant to bipolar disorder as these are potential deleterious effect is more straightforward than
identified—keeping in mind that expression may change with most common single-nucleotide polymorphisms. On
the other hand, all apparently healthy individuals carry
some CNVs, and in fact many proteins are entirely normal
despite being hemizygous (i.e., individuals carrying only
one copy of the functional gene).
α2 In some neuropsychiatric disorders, including autism
and schizophrenia, a growing body of evidence suggests
δ that the overall burden of CNV—that is, the number of
α1
CNV’s carried by an individual—is increased. In bipolar
disorder, however, the data regarding CNV burden is less
clear. Some studies suggest an increase in overall CNV bur-
β den (Malhotra et  al., 2011), while others do not (Bergen
et al., 2012).
While CNV studies in bipolar disorder remain a work
in progress, there is evidence that certain recognized syn-
Figure 15.1  Structure of voltage-gated calcium channels dromes involving deletions may be associated with bipolar

18 4   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
symptomatology, or even frank bipolar disorder. Perhaps reached a threshold for being considered genome-wide sig-
the best example to date is a deletion of 22q11.2, so-called nificant (International Schizophrenia Consotrium et  al.,
velocardiofacial syndrome (VCFS), which has been associ- 2009). These aggregate risk scores, sometimes referred to
ated with bipolar spectrum features (Papolos et al., 1996). as polygene scores because they integrate data from a large
Individuals with VCFS may have a variety of congenital number of genetic loci, have multiple applications. First,
anomalies, including cardiac and renal abnormalities, they may be used to examine the relationship between
cleft lip/palate, hearing loss, intellectual disability, and individual disorders by estimating the extent to which risk
immune dysfunction. While there is a characteristic facial (loading) for one disease predicts risk for another. One ini-
appearance involving small mouth and ears and smaller, tial application demonstrated that schizophrenia loading
squared-off ears, these features may be quite subtle, and predicted bipolar disorder risk (but not risk for multiple
the systemic manifestations may also be missed when non-psychiatric disorders), helping to refine psychiatric
mild. To date, the prevalence of individuals with VCFS in nosology and highlighting the overlap between disorders
genetic association studies of bipolar disorder appears to be at a genetic level (International Schizophrenia Consotrium
extremely low, suggesting rates of missed diagnosis may be et al., 2009). The most recent estimates of polygenic effect
low, or that these individuals may be more often charac- suggest that loading for schizophrenia can explain a sub-
terized as having psychiatric disorders other than bipolar. stantial proportion of genetic risk for bipolar disorder,
However, the actual prevalence of VCFS among clinical highlighting this overlap. A  second application simply
populations with bipolar disorder has not been determined. summarizes the extent to which genome-wide studies have
In addition to CNVs, the role of rare single nucleotide succeeded in identifying risk loci. That is, polygenic scores
variation has also begun to be examined in bipolar disor- can be used to summarize how much risk for a given disor-
der. (By convention, “rare” variants are generally considered der is explained by a set of common variants. Finally, poly-
to be those with a frequency of less than 1%, although this gene scores may serve as a useful summary measure for how
threshold reflects genotyping platform design rather than much loading for common risk variants a given individual
any particular discontinuity between common and rare carries. The potential application of such scores is addressed
variation.) Investigations of rare variants generally utilize in more detail in the clinical section later in this chapter.
sequencing methodologies—that is, reading every base pair.
While technology exists for sequencing an entire genome,
a more cost-effective strategy is sequencing only expressed S U BPH E NO T Y PE S
(transcribed) regions of the genome, so-called whole-exome
sequencing. More recently, in an effort to make genotyping In an effort to speed the identification of risk variants in
of rare variants more cost-effective, arrays have been devel- bipolar disorder, a number of groups have tried to refine
oped to genotype panels of rare variants. the bipolar disorder phenotype to identify more strongly
The challenge in such studies, as with investigations of heritable forms of illness (Potash et al., 2007). Some twin
CNV, is that all individuals carry rare variants, the majority studies suggest similar heritability across bipolar I and II,
of which have no apparent functional implications whatso- for example (Edvardsen et al., 2008), even though the latter
ever. To date, definitive results from sequencing studies in group is often excluded from bipolar disorder association
bipolar disorder have not been reported. An intriguing case studies. Other phenotypes have been investigated that cross
report describes an individual with Timothy Syndrome, an diagnostic boundaries. One example is suicide attempt, for
autism-like syndrome with multiple congenital anomalies which risk is increased up to 20-fold in individuals with
that has been shown to arise as a result of rare functional mood disorders (Osby, Brandt,Correia, Ekbom, & Sparen,
variants (mutations) in CACNA1C, who developed bipolar 2001). To date, efforts to identify risk genes for suicide have
disorder in adulthood (Gershon et al., 2013). not yielded results at a genome-wide level of significance
(Perlis, Huang, et al., 2010; Willour et al., 2012), although
efforts to understand this subphenotype in larger cohorts
P OLYG E N IC S T U DI E S are ongoing.
Treatment response in bipolar disorder represents a par-
As a complement to studies of individual loci, it is also pos- ticularly appealing subphenotype, as it might provide con-
sible to examine the cumulative burden of potential risk vergent support for particular loci or pathways in bipolar
variants in an individual. This strategy was demonstrated disorder. That is, genes that influence response to bipolar
to be sensitive even to common variants that have not yet pharmacotherapies might overlap with those influencing

T he E merging G enetics of B ipolar D isorder   •   185


disease pathophysiology. There are a relative paucity of Unfortunately, characterizing such environmental risks
cohorts with prospectively-derived treatment response is challenging because of the sheer breadth of such poten-
data—that is, investigations that do not rely on retrospec- tial risks and the paucity of appropriate data sets. One area
tive report of benefit. Such retrospective reports by patients of focus has been early adversity—major traumas occur-
have been shown in electronic health records to be fre- ring during childhood—in part because data on such envi-
quently discordant from actual outcomes (Simon, Rutter, ronmental risk exists and has been associated with major
Stewart, Pabiniak, & Wehnes, 2012). depressive disorder. One investigation suggests that early
The first genome-wide association study of lithium adversity is also associated with bipolar disorder risk, as
response (Perlis, Smoller, et  al., 2009)  failed to identify has been observed with major depressive disorder, consis-
any loci at a level of genome-wide significance. However, tent with prior, smaller studies (Gilman, Dupuy, & Perlis,
looking across two distinct cohorts, it did identify five 2012). In particular, childhood physical abuse and sexual
loci with suggestive evidence in both, most notably a maltreatment have been associated with later risk for bipo-
variant in the gene coding for the glutamate/alpha-am lar disorder and may sensitize individuals to subsequent
ino-3-hydroxy-5-methyl-4-isoxazolpropionate receptor stressors, such that these later stressors are more apt to pre-
GRIA2. A  subsequent study among 294 Han Chinese cipitate mania (Gilman et al., 2014).
individuals associated common variation in the glutamic An obstacle to understanding gene-by-environment
acid decarboxylase-like 1 gene with differential response, interactions is the need for even larger cohorts to achieve
with a minimum p value of 5.50 × 10 –37 (Chen et al., 2014). adequate power to detect such interactions, particularly
Of note, despite the gene name, it does not appear to play cohorts in which a range of potential environmental risk
a role in glutamatergic synapses, and so the two sets of is assessed. Similar challenges apply to efforts to identify
genome-wide association study results do not necessarily epistatic effects (i.e., gene-by-gene interactions, where indi-
implicate similar pathways. Moreover, multiple other rep- vidual effects are not merely additive).
lication efforts, including one by a consortium of groups
studying lithium response genetics, provided no support
for this locus (Schulze et al., 2010). C L I N IC A L A PPL IC AT ION S
To date, no genome-wide studies of response to other
pharmacotherapies have been reported. A modest number At present, there is no genetic diagnostic tool with clinical
of candidate-gene studies have been reported (e.g., with the utility in bipolar disorder. While investigations of common
combination of olanzapine and fluoxetine; Perlis, Adams, and rare variation explain a growing proportion of bipolar
et al., 2010), but none have yet been replicated. Efforts are risk, in absolute terms this quantity remains modest—the
also ongoing to establish larger lithium-response cohorts variant in CACNA1C of largest effect increases bipolar risk
that rely on retrospective determination of response using by ~18% (Craddock & Sklar, 2013). Still, there is substantial
a standardized measure referred to as the Alda scale (Grof reason to expect that, perhaps in combination with clinical
et  al., 2001). A  preliminary report suggested reasonable risk factors, the emerging genetics of bipolar disorder will
interrater reliability for the efficacy (“A domain”) aspect of begin to contribute to longitudinal intervention studies of
this measure (Manchia et al., 2013). high-risk individuals, biomarker-stratified clinical trials,
and ultimately truly novel diagnostic and therapeutic tools.
Disclosure statement: Dr. Roy Perlis has served on scientific
N E W DI R E C T ION S advisory boards or consulted to Genomind, Healthrageous,
Perfect Health, Psybrain, and RID Ventures.
As main effects are reliably identified for individual loci, the
possibility of detecting gene-by-environment interactions
becomes more plausible. Even for monozygotic twin pairs R E F E R E NC E S
in which one twin is affected, the risk is only ~50% that the
co-twin is affected, which argues for such effects. More gener- Bergen, S. E., O’Dushlaine, C. T., Ripke, S., Lee, P. H., Ruderfer, D. M.,
ally, based on the twin studies addressed earlier in this chap- Akterin, S., . . . Sullivan, P. F. (2012). Genome-wide association study
in a Swedish population yields support for greater CNV and MHC
ter, it is clear that environmental risk (either independently or involvement in schizophrenia compared with bipolar disorder.
through interactions with genetic variation) should account Molecular Psychiatry, 17, 880–886.
for one-third to one-half of overall bipolar disorder risk. Casamassima, F., Hay, A. C., Benedetti, A., Lattanzi, L., Cassano, G. B.,
& Perlis, R.  H. (2010). L-type calcium channels and psychiatric

18 6   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
disorders: A brief review. American Journal of Medical Genetics: Part Kieseppa, T., Partonen, T., Haukka, J., Kaprio, J., & Lonnqvist,
B, Neuropsychiatric Genetics, 153, 1373–1390. J. (2004). High concordance of bipolar I  disorder in a nation-
Chen, C.  H., Lee, C.  S., Lee, M.  T., Ouyang, W.  C., Chen, C.  C., wide sample of twins. The American Journal of Psychiatry, 161,
Chong, M. Y., . . . Cheng, A. T. (2014). Variant GADL1 and response 1814–1821.
to lithium therapy in bipolar I disorder. The New England Journal of Lichtenstein, P., Yip, B.  H., Bjork, C., Pawitan Y., Cannon T.  D.,
Medicine, 370, 119–128. Sullivan P. F., & Hultman C. M. (2009). Common genetic deter-
Cichon, S., Muhleisen, T. W., Degenhardt, F. A., Mattheisen, M., Miró, minants of schizophrenia and bipolar disorder in Swedish fami-
X., Strohmaier, J., . . . Nöthen, M. M. (2011). Genome-wide associa- lies: A population-based study. Lancet, 373, 234–239.
tion study identifies genetic variation in neurocan as a susceptibility Malhotra, D., McCarthy, S., Michaelson, J. J., Vacic, V., Burdick, K. E.,
factor for bipolar disorder. American Journal of Human Genetics, Yoon, S., . . . Sebat, J. (2011). High frequencies of de novo CNVs in
88, 372–381. bipolar disorder and schizophrenia. Neuron, 72, 951–963.
Craddock, N., & Jones, I. (1999). Genetics of bipolar disorder. Journal Manchia, M., Adli, M., Akula, N., Ardau, R., Aubry, J. M., Backlund,
of Medical Genetics, 36, 585–594. L., . . . Alda, M. (2013). Assessment of Response to lithium main-
Craddock, N., & Sklar, P. (2013). Genetics of bipolar disorder. Lancet, tenance treatment in bipolar disorder: A Consortium on Lithium
381, 1654–1662. Genetics (ConLiGen) Report. PLoS One, 8, e65636.
Craddock, N., Khodel, V., Van Eerdewegh, P., & Reich, T. (1995). Math- Merikangas, K. R., Merikangas, J. R., & Angst, J. (1993). Headache syn-
ematical limits of multilocus models:  The genetic transmission of dromes and psychiatric disorders:  Association and familial trans-
bipolar disorder. American Journal of Human Genetics,;57, 690–702. mission. Journal of Psychiatric Research, 27, 197–210.
Dima, D., Jogia, J., Collier, D., Vassos, E., Burdick, K. E., & Frangou, Muhleisen, T.  W., Leber, M., Schulze, T.  G., Strohmaier, J., Degen-
S. (2013). Independent modulation of engagement and connectiv- hardt, F., Treutlein, J., . . . Cichon, S. (2014). Genome-wide asso-
ity of the facial network during affect processing by CACNA1C ciation study reveals two new risk loci for bipolar disorder. Nature
and ANK3 risk genes for bipolar disorder. JAMA Psychiatry, 70, Communications, 5, 3339.
1303–1311. Nurnberger, J.  I. Jr., Koller, D.  L., Jung, J., Edenberg, H. J., Foroud,
Duncan, L. E., Holmans, P. A., Lee, P. H., O’Dushlaine, C. T., Kirby, T., Guella I., . . . Kelsoe, J.  R. (2014). Identification of pathways
A.  W., Smoller, J.  W., . . . Cohen, B.  M. (2014). Pathway analyses for bipolar disorder:  A  meta-analysis. JAMA Psychiatry, 71(6),
implicate glial cells in schizophrenia. PLoS One, 9, e89441. 657–664.
Edvardsen, J., Torgersen, S., Roysamb, E., Lygren, S., Skre, I., Onstad, Osby, U., Brandt, L., Correia, N., Ekbom, A., & Sparen, P. (2001).
S., & Oien, P. A. (2008). Heritability of bipolar spectrum disorders. Excess mortality in bipolar and unipolar disorder in Sweden.
Unity or heterogeneity? Journal of Affective Disorders, 106, 229–240. Archives of General Psychiatry, 58, 844–850.
Ferreira, M. A., O’Donovan, M. C., Meng, Y. A., Jones, I. R., Rud- Ostacher, M. J., Iosifescu, D. V., Hay, A., Blumenthal, S. R., Sklar, P.,
erfer D.  M., Jones, L., . . . Craddock, N. (2008). Collaborative & Perlis, R. H. (2014). Pilot investigation of isradipine in the treat-
genome-wide association analysis supports a role for ANK3 ment of bipolar depression motivated by genome-wide association.
and CACNA1C in bipolar disorder. Nature Genetics, 40, Bipolar Disorders, 16, 199–203.
1056–1058. Papolos, D. F., Faedda, G. L., Veit, S., Goldberg, R., Morrow, B., Kucher-
Gershon, E.  S., Grennan, K., Busnello, J., Badner, J.  A., Ovsiew, F., lapati, R., & Shprintzen, R. J. (1996). Bipolar spectrum disorders in
Memon S., . . . Liu, C. (2013). A rare mutation of CACNA1C in a patients diagnosed with velo-cardio-facial syndrome: Does a hemi-
patient with bipolar disorder, and decreased gene expression asso- zygous deletion of chromosome 22q11 result in bipolar affective dis-
ciated with a bipolar-associated common SNP of CACNA1C in order? The American Journal of Psychiatry, 153, 1541–1547.
brain. Molecular Psychiatry 2013. Perlis, R. H. (2005). Misdiagnosis of bipolar disorder. American Jour-
Gilman, S. E., Dupuy, J. M., & Perlis, R. H. (2012). Risks for the tran- nal of Managed Care, 11, S271–S274.
sition from major depressive disorder to bipolar disorder in the Perlis, R. H., Adams, D. H., Fijal, B., Sutton, V. K., Farmen, M., Breier,
National Epidemiologic Survey on Alcohol and Related Condi- A., & Houston, J. P. (2010). Genetic association study of treatment
tions. The Journal of Clinical Psychiatry, 73, 829–836. response with olanzapine/fluoxetine combination or lamotrigine in
Gilman, S. E., Ni, M. Y., Dunn, E. C., Breslau, J., McLaughlin, K. A., bipolar I depression. The Journal of Clinical Psychiatry, 71, 599–605.
Smoller, J. W., & Perlis, R. H. (2014). Contributions of the social Perlis, R. H., Dennehy, E. B., Miklowitz, D. J., Delbello, M. P., Ostacher,
environment to first-onset and recurrent mania. Molecular Psychia- M., Calabrese, J. R., . . . Sachs, G. (2009). Retrospective age at onset
try. Advance online publication. doi:10.1038/mp.2014.36. of bipolar disorder and outcome during two-year follow-up: Results
Green, E.  K., Grozeva, D., Forty, L., Gordon-Smith, K., Russell, E., from the STEP-BD study. Bipolar Disorders, 11, 391–400.
Farmer, A., . . . Craddock, N. (2013). Association at SYNE1 in both Perlis, R. H., Huang, J., & Purcell, S., Fava, M., Rush, A. J., Sullivan,
bipolar disorder and recurrent major depression. Molecular Psychia- P. F., . . . Smoller, J. W. (2010). Genome-wide association study of sui-
try, 18, 614–617. cide attempts in mood disorder patients. The American Journal of
Green, E. K., Hamshere, M., Forty, L., Gordon-Smith, K., Fraser, C., Psychiatry, 167, 1499–1507.
Russell, E., . . . Craddock, N. (2013). Replication of bipolar disorder Perlis, R. H., Smoller, J. W., Ferreira, M. A., McQuillin, A., Bass, N.,
susceptibility alleles and identification of two novel genome-wide Lawrence, J., . . . Purcell, S. (2009). A genomewide association study
significant associations in a new bipolar disorder case-control sam- of response to lithium for prevention of recurrence in bipolar disor-
ple. Molecular Psychiatry, 18, 1302–1307. der. The American Journal of Psychiatry, 166, 718–725.
Grof, P., Duffy, A., Cavazzoni, P., Grof, E., Garnham, J., MacDougall, Potash, J.  B., Toolan, J., Steele, J., Miller, E.  B., Pearl, J., Zandi,
M., . . . Alda, M. (2002). Is response to prophylactic lithium a famil- P. P., . . . McMahon, F. J. (2007). The bipolar disorder phenome data-
ial trait? Journal of Clinical Psychiatry, 63, 942–947. base: a resource for genetic studies. The American Journal of Psychia-
Holmans, P., Green, E. K., Pahwa, J. S., Ferreira, M. A., Purcell, S. M., try, 164, 1229–1237.
Sklar, P., . . . Craddock, N. (2009). Gene ontology analysis of GWA Psychiatric GWAS Consortium Bipolar Disorder Working Group.
study data sets provides insights into the biology of bipolar disorder. (2011). Large-scale genome-wide association analysis of bipolar
American Journal of Human Genetics, 85, 13–24. disorder identifies a new susceptibility locus near ODZ4. Nature
International Schizophrenia Consotrium, Purcell, S. M., Wray, N. R., Genetics, 43, 977–983.
Stone, J. L., Visscher, P. M., O’Donovan, M. C., . . . Sklar, P. (2009). Schulze, T.  G., Alda, M., Adli, M., Akula, N., Ardau, R., Bui,
Common polygenic variation contributes to risk of schizophrenia E. T., . . . McMahon, F. J. (2010). The International Consortium on
and bipolar disorder. Nature, 460, 748–752. Lithium Genetics (ConLiGen):  an initiative by the NIMH and

T he E merging G enetics of B ipolar D isorder   •   18 7


IGSLI to study the genetic basis of response to lithium treatment. Smoller, J. W., & Finn, C. T. (2003). Family, twin, and adoption stud-
Neuropsychobiology, 62, 72–78. ies of bipolar disorder. American Journal of Medical Genetics, Part
Seifuddin, F., Mahon, P. B., Judy, J., Pirooznia, M., Jancic, D., Taylor, C: Seminars in Medical Genetics, 123, 48–58.
J., . . . Zandi, P. P. (2012). Meta-analysis of genetic association studies Willour, V. L., Seifuddin, F., Mahon, P. B., Jancic, D., Pirooznia, M.,
on bipolar disorder. American Journal of Medical Genetics: Part B, Steele, J., . . . (2012). A genome-wide association study of attempted
Neuropsychiatric Genetics;159, 508–518. suicide. Molecular Psychiatry, 17, 433–444.
Simon, G. E., Rutter, C. M., Stewart, C., Pabiniak, C., & Wehnes, L. (2012). Wray, N. R., & Gottesman, I. I. (2012). Using summary data from the
Response to past depression treatments is not accurately recalled: Com- Danish national registers to estimate heritabilities for schizophre-
parison of structured recall and patient health questionnaire scores in nia, bipolar disorder, and major depressive disorder. Frontiers in
medical records. The Journal of Clinical Psychiatry, 73, 1503–1508. Genetics, 3, 118.

18 8   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
PA R T   I I I

C O G N I T ION
16.
COGNITIV E IMPAIR MENTS A ND EV ERY DAY
DISA BILIT Y IN BIPOLAR ILLNESS
Colin A. Depp, Alexandrea L. Harmell, and Philip D. Harvey

INTRODUCTION and the role of such measures in gauging the impact of new
treatments.
Bipolar disorder (BD), by any measure, produces substan- A number of lessons can be drawn from these efforts as
tial disability. Historical depictions of BD as a relapsing they apply to BD (Harvey, Wingo, Burdick, & Baldessarini,
illness with inter-episode recovery of function now appear 2010), yet there are several unique aspects of BD that neces-
to be inaccurate for most patients. Longitudinal data sug- sitate bipolar-specific data. In particular, cognitive impair-
gests that the majority of patients spend most of their time ments in BD are less common and severe than among
experiencing some amount of disability. Even though up patients with schizophrenia, with a meta-analysis of 24
to 90% of patients recover from a first episode of mania, studies estimating that BD is associated with approximately
only one-third experience a return to premorbid function- a 0.5 standard deviation better performance on neuropsy-
ing (Tohen et al., 2003). Globally, according to the World chological testing than schizophrenia (Krabbendam, Arts,
Health Organization’s worldwide burden of disease study, Van Os, & Aleman, 2005). Bipolar patients additionally
BD is the seventh leading cause of years of life lost to dis- appear to have higher levels of premorbid functioning than
ability among all illnesses. These functional impairments do people with schizophrenia (Reichenberg et  al., 2002),
translate to massive indirect costs, such as from lost produc- and, unlike the generalized neurocognitive deficits seen in
tivity and absenteeism, which were estimated in 2009 to schizophrenia, the deficits associated with BD are thought
total $150 million in the United States per year (Dilsaver, to be more selective (Goldberg & Chengappa, 2009). Such
2011). This cost estimate is about double that of schizophre- heterogeneity in level and domain of impairment also
nia, possibly because of the higher levels of functioning of appears to be present in the realm of disability. In this chap-
employed patients with bipolar illness. Historically, much ter, we draw from lessons learned in schizophrenia research,
of the focus on the cause of these functional deficits has as well as the accumulating data in bipolar samples on the
been on the affective symptoms experienced by patients. impact of cognitive impairments on functional abilities.
However, more recent work has suggested that cognitive We also provide directions for future research and discuss
impairments play a substantial role in producing disabil- implications for cognitive and functional rehabilitation.
ity, perhaps even more so than affective symptoms over the
long term.
The intersection of cognitive ability and functional dis- R AT E S OF   F U NC T ION A L
ability has been better documented in other psychiatric I M PA I R M E N T I N
illnesses, particularly in schizophrenia (Bowie & Harvey, B I P OL A R DI S OR DE R
2005). Coordinated and federally funded efforts such
as the MATRICS initiative (Nuechterlein et  al., 2008)  Functional outcome is a multidimensional construct with
and VALERO (Leifker, Patterson, Heaton, & Harvey, no single indicator, spanning the impact of illness on
2011) have delineated the impact and measurement issues physical, emotional, social, and productive life domains.
in functional outcome in schizophrenia. There is now a Moreover, there is considerable variation in the measure-
strong evidence base in schizophrenia guiding data collec- ment strategies used to quantify disability. Research on
tion strategies and modalities on functional measurement functional impairment in BD has largely focused on

191
measuring the impact of the illness on vocational outcomes disability in BD is highly important, perhaps more so than
and subjective quality of life, with relatively less work on in schizophrenia.
social, residential, and independent living skills. We sum-
marize briefly the prevalence of disability across measure-
S TA N DA R DI Z E D C L I N IC I A N R AT E D
ment strategies in BD, prior to describing the impact of
M E A S U R E S OF F U NC T ION I NG A N D
cognitive ability on these disabilities.
QUA L I T Y OF L I F E
A range of clinical and self-reported measures have been
M I L E S TON E S used to assess functional outcome in BD. Nonetheless,
We refer to lifespan achievements such as the attainment general (e.g., Medical Outcomes Study), psychiatric-
of employment, residential independence, or marriage as illness focused (e.g., Specific Level of Functioning Scale;
functional milestones, which can be differentiated from Schneider & Struening, 1983), and disease-specific (e.g.,
functional outcome measures that typically involve sum- Functional Assessment Short Test) instruments have been
mative clinician or patient ratings on standardized instru- used to assess functioning and quality of life and patients
ments that also address “subthreshold” achievements, with BD evidence global impairments of approximately 0.5
such as seeking work. In regard to employment, a recent standard deviations below age-matched comparison sub-
quantitative review indicated that about 40% to 60% of jects. Examples of common functional outcome instruments
patients were currently participating in paid employment. are provided in Box 16.1. Of note, it is clear from work in
Followed over time, over 50% of patients experienced a schizophrenia that these scales frequently suffer from suspect
downward change in the level of employment (Marwaha, validity when indexed to neurocognitive ability or even to
Durrani, & Singh, 2013). At the population level, patients functional milestones. It is also apparent that a considerable
with BD experienced 1.7 disability days per month, higher proportion of patients over- or underestimate their own func-
than most physical disorders, and 50 days of lost work due tioning (Bowie et al., 2007), with cognitive impairments and
to illness per year (Dean, Gerner, & Gerner, 2004). Half poor levels of functioning often associated with overestima-
of patients with BD reported missing at least one week tion (Sabbag et al., 2011) and significant depression associ-
of work in the past month (Stang, Frank, Ulcickas Yood, ated with underestimation (Bowie et al., 2007). Interestingly,
Wells, & Burch, 2007). In a Norwegian sample, 48% of mild to moderate levels of depression are associated with
patients were receiving disability compensation associated accurate assessment of current functioning (Sabbag et  al.,
with work dysfunction (Schoeyen et al., 2011). In regard to 2012). It is not known whether these same biases are evident
marriage, in a multinational survey, patients with BD were in BD, although there is reason to suspect that such prob-
reported to be 1.6 times more likely to be divorced com- lems may exist. This is likely to be important, given findings
pared to population controls (Breslau et al., 2011). As far as suggesting that many patients with bipolar illness experience
residential independence, in a relatively high-functioning stable mild depression (Judd & Akiskal, 2003; Judd et  al.,
sample, 18% of patients were not independently resid- 2012) in the context of fluctuations between depression and
ing (i.e., residing in sheltered housing; Depp, Mausbach, euphoria. Thus fluctuations in the ability to evaluate current
Bowie, et  al., 2012). Finally, incarceration is a different functioning may have a changeable course, much like every-
sort of milestone, and in a large sample of persons who day functioning, in response to mood states.
were serving criminal sentences, BD was associated with a
three-fold higher risk of four of more incarcerations. This
elevated level of risk of repeated incarcerations was higher Box 16.1.  FUNCTIONAL OUTCOME
than that associated with any other mental illness and may INSTRUMENTS USED IN BIPOLAR
also be associated with additional comorbidities such as DISORDER
substance abuse.
Medical Outcomes Study (SF-36)
Two somewhat divergent conclusions can be drawn
Specific Level of Functioning Scale (SLOF)
from these figures. On the one hand, BD produces remark-
Multidimensional Scale for Independent Functioning 
able global burden, second only to schizophrenia among
(MSIF)
mental illnesses per person affected. On the other hand, it
Functional Assessment Short Test (FAST)
is apparent that a substantial proportion of people with BD
Social Adjustment Scale (SAS)
are married, employed, and residing independently. Thus
LIFE Range of Impaired Functioning Tool (LIFE-RIFT)
understanding the factors that account for heterogeneity in

19 2   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
Moreover, a key consideration is the extent to which cognitive ability on various indicators of functioning. Our
these impairments are state-dependent (e.g., exacerbated recent meta-analysis of 22 studies estimated the aggre-
during manic or depressive episodes). The general finding gate strength of association between specific neurocogni-
is that quality of life is diminished during both manic and tive abilities and everyday functioning by examining the
depressive episodes but is also considerably diminished pooled correlation coefficient reported in these studies.
compared to healthy controls during euthymic periods. As The modal study was cross-sectional, small (mean n = 61,
such, functional impairment is likely to be both a state and SD = 23) and used a convenience clinical sample of out-
a trait in BD. In addition, mood state may produce biases in patients. One of the key limitations of this meta-analysis
the self-evaluation of functioning, which may be particu- was the measurement of disability, as the majority of stud-
larly evident in BD and are discussed later in this chapter. ies (14 out of 22) measured functional outcome with the
Global Assessment of Functioning (Endicott, Spitzer,
Fleiss, & Cohen, 1976; Jones, Thornicroft, Coffey, &
F U NC T IONA L C A PACI T Y
Dunn, 1995), a single-item clinical rating that is con-
Only recently have objective performance-based mea- founded with symptom severity(Soderberg, Tungstrom, &
sures of functional capacity been used to quantify func- Armelius, 2005).
tional impairment in BD. These instruments, which have Notwithstanding the limitations of this body of
become commonplace as proximal or secondary mea- research, in regard to the estimated impact of cognitive
sures in schizophrenia treatment trials, typically involve ability on functioning, the mean correlation coefficient
observed performance on ecologically valid real-world was r  =  0.22, which is identical to that found in a simi-
tasks such as managing medication dosing schedules or lar meta-analysis in schizophrenia by Fett and colleagues
navigating public transportation. In the few studies that (2011). Notably, there were no moderating influences of
have employed performance based measures, such as the the mean age of the sample or sample size. There were also
UCSD Performance-Based Skills Assessment, Social no moderating influences of whether or not the sample was
Skills Performance Assessment, Performance Assessment restricted to euthymic patients. As such, even though the
of Self-Care Skills, and Medication Management Ability average cognitive performance in BD is substantially better
Assessment, patients with BD evidence impairment com- than that in schizophrenia, the proportion of variation in
pared to normal controls on independent living skills, med- predicting functional outcome is the same. The significant
ication management ability, and social competence (Depp association between cognitive and functional measures
et al., 2009; Gildengers et al., 2013). Patients across clini- seemed to be fairly robust across unique samples with vary-
cal states perform worse than healthy comparators, with ing restrictions based on demographic characteristics of
relatively little influence of symptom severity on perfor- symptom severity.
mance on these measures. Moreover, performance on the
Performance-Based Skills Assessment predicts real-world
L ONG I T U DI NA L S T U DI E S
performance in employment and residential status in out-
OF T H E I M PAC T OF CO G N I T I V E
patients with BD. Thus disability in BD may arise from
I M PA I R M E N T ON DI S A BI L I T Y
skill deficits, which are more stable and closely linked to
cognitive abilities than to mood symptoms. A smaller number of longitudinal studies have examined
the long-term impact of cognitive impairment on func-
tional outcome. Specifically, in a group of recently diag-
I M PAC T OF   C O G N I T I V E nosed patients with BD I, Torres et  al. (2011) found a
F U NC T ION I NG ON   DI S A B I L I T Y significant association between memory and six-month
functional outcome using the Multidimensional Scale of
Independent Functioning, even after partialing out the
M E TA-A NA LY S I S OF T H E I M PAC T
effects of mood symptomatology. Another study examined
OF CO G N I T I V E I M PA I R M E N T
a sample of euthymic patients with BD and found that
ON DI SA BI L I T Y
impairments in the areas of attention, executive function,
The recognition of cognitive impairments in BD is rela- and verbal memory at baseline independently predicted
tively recent in the literature, as is the persistence of level of functional outcome one year later (as assessed
disability across mood states. Nonetheless, a number by Global Assessment of Functioning scores and scores
of cross-sectional studies have examined the impact of on the Functioning Assessment Short Test). Significant

C ogniti v e I mpairments and E v eryday D isability in B ipolar I llness   •   19 3


associations have also been reported between cognitive performance-based measurements of functional capacity
impairment and functional outcomes at 2- (Gildengers being more convergent with performance-based assess-
et  al., 2013), 6- (Mora, Portella, Forcada, Vieta, & Mur, ments of cognitive ability than clinician or self-reports.
2013), and 15-year follow-up visits (Burdick, Goldberg, & The strongest correlations, pooling effects across cog-
Harrow, 2010). Despite the fact that there has been very nitive domains, were seen with Performance-Based
little research exploring whether or not cognitive function- (r  =  0.32) and Functional Milestone (r  =  0.33) mea-
ing predicts future functional outcomes, of the studies that sures of functioning. Weaker correlations were evident
do evaluate this relationship, there appear to be robust find- in Clinician-Reported (r  =  0.23) and Self-Reported
ings showing that cognitive impairment among patients Functioning (r = 0.20). It is likely that functional capac-
with BD is significantly associated with poorer long-term ity measures are more apt to assess what patients “can do,”
functioning. which is more specifically determined by cognitive abil-
ity, rather than ratings on functional outcome measures,
which also reflect motivation and social-environmental
E V I DE NC E FOR S PECI F ICI T Y
opportunities. Moreover, Zarate, Tohen, Land, and
I N CO G N I T I V E D OM A I NS A N D
Cavanagh (2000) concluded that individuals with
DI SA BI L I T Y M E A S U R E S
BD may perform well in one area of functioning but
A number of individual studies have implicated impaired poorly in others. In examining the relative determina-
executive functioning in particular as a predictor of func- tion of functional outcome domains using multivariate
tional deficits in BD (Bonnin et al., 2010). Other studies models, Bowie et  al. (2010) found that cognitive abil-
have identified selective impact of attention and verbal ity accounted for a higher proportion of variation in
memory deficits (Andreou & Bozikas, 2013; Martino predicting instrumental activities than interpersonal
et  al., 2009)  or processing speed (Burdick et  al., 2010; functioning, with work skills being an intermediate of
Mur, Portella, Martinez-Aran, Pifarre, & Vieta, 2008) as the two.
being among the most significant predictors of functional
outcomes. In the more developed literature in schizophre-
COM PA R I S ON OF F I N DI NG S
nia, there does not appear to be substantial evidence that
TO S CH I Z OPH R E N I A
any one cognitive domain is associated with greater func-
tional impairment (Dickinson & Harvey, 2009; Jabben, In addition to having a similar strength of association
Arts, van Os, & Krabbendam, 2010); To help evaluate between cognitive and functional domains as schizophre-
whether or not a similar global versus specific pattern of nia, the structure of the association between cognition,
cognitive impairment is representative in BD, individual functional capacity, and functional outcome is quite similar.
cognitive domains were assessed for their mean correlation Indeed, Bowie et al. (2010) found that the same structural
with functional outcome in the meta-analysis described relationships were evident in BD, with objective functional
previously (Depp, Mausbach, Harmell, et  al., 2012). The capacity measures mediating cognitive-functional outcome
proportion of variance accounted for by cognition in pre- associations in work abilities, interpersonal functioning,
dicting various measures of functioning was significant and independent living abilities. However, there were more
for all cognitive domains measured (lowest correlation subtle but potentially important differences. First, direct
was r  =  0.21 for Visual Learning and Memory domain and unmediated associations with functional outcome
and the highest was r = 0.29 for Working Memory). There were evident with depressive and manic symptoms, whereas
was no statistical evidence for significant variation among symptoms had little direct impact in schizophrenia. As such,
these domains. Thus there does not appear to be compel- the contributors to disability in BD are possibly more com-
ling evidence that any one cognitive domain produces plex than in schizophrenia, driven additionally by dynamic
more disability than others, combining evidence from the mood symptoms. Second, a greater proportion of functional
meta-analysis, individual studies, and the larger body of outcome in BD is (currently) unexplained—the prediction
literature in schizophrenia. of work, social, and instrumental outcomes was roughly
In contrast, there may be greater variation in the cor- two-thirds (R 2  =  .30 – .46) of that seen in schizophrenia
relation between cognition and outcomes depending (R 2 = .40 – .55). In Table 16.1 we summarize some of the
on the method by which disability is measured. As in challenges in extrapolating functional assessment methods
schizophrenia, there does seem to be a discrepancy with from schizophrenia to BD, along with potential solutions.

19 4   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
Table 16 .1.  CHALLENGES IN EXTR APOLATING FUNCTIONAL OUTCOME R ESEARCH FROM 
SCHIZOPHR ENIA TO BIPOLAR DISOR DER

BIPOLAR DISOR DER IN COMPAR ISON PSYCHOMETR IC CH ALLENGES METHODOLOGICAL SOLUTIONS


TO SCHIZOPHR ENI A

1. More subtle and selective deficits in •  Possible ceiling effects • Use of more difficult tests of functional
cognition and functioning capacity
2.  G
 reater cross-sectional impact of • Relationship between neuropsychological • Longitudinal analysis of impact of
symptoms, particularly depression, (NP) and functioning may differ by mood depression construct validity
on ability and real world (RW) states • Measure divergence between observer
functioning • Self-assessment may diverge from and patient reports
observation and both may be affected • Inclusion of measures of social cognitive
by current mood state deficits

3.  G
 reater between- and within-person • Instability of estimates Use of within-subjects analyses to quantify
variability • Practice effects stability across phases

4.  H
 igher rates of achievement of func- Capacity measures and rating scales may Validation against functional milestones
tional milestones not be sensitive to milestones such as employment and independence

COM PA R I S ON OF F I N DI NG S COGNITI V E A BILIT Y A ND


TO U N I P OL A R DE PR E S S ION I T S A S S O C I AT ION W I T H R I S K
FAC TOR S F OR   DI S A B I L I T Y
Although both patients with unipolar depression and
patients with BD have been reported to perform below
the level of nonpsychiatric controls on cognitive tasks, BIOL O G IC A L M ECH A N I SM S
when the two groups are directly compared to one OF BI P OL A R DI S OR DE R
another, BD has been shown to be related to worse
Although it is becoming increasingly recognized that cog-
overall performance in areas such as delayed memory,
nitive impairment is a core feature of BD, the exact etiol-
executive functioning, language, and visuomotor abili-
ogy of this impairment has not been well defined. Different
ties (Gildengers et  al., 2012). Moreover, in a separate
mechanisms have been hypothesized as potential origins
study investigating BD versus depression-specific mark-
including, but not limited to, dysregulation of dopaminer-
ers of affective disorders, impaired sustained attention
gic and glutamatergic systems, the presence of mitochon-
was shown to be more specific to BD (irrespective of
drial dysfunction and oxidative stress, and inflammation of
euthymic or depressive state), and executive function-
cytokines and neurotrophins (Berk et al., 2010). Individuals
ing was reported to be a potential marker more specific
experiencing recurrent episodes have also been shown to
to depression as a whole (as this domain was found to be
have progressive ventricular enlargement (Strakowski et al.,
impaired in both unipolar depression and depressed BD
2002) and loss of grey-matter thickness (Lyoo et al., 2006),
patients; Maalouf et  al., 2010). Among patients with
both of which may contribute to similarly progressive
jobs, BD has also been associated with twice the absen-
changes in neurocognition. Compounding the difficulty in
teeism as unipolar depression (Laxman, Lovibond, &
finding specific causes of cognitive impairment is the fact
Hassan, 2008). One study assessing work impairment
that circumscribed cognitive deficits may have iatrogenic
over 15 years, reported that BD-I patients were unable
origins, as many of the medications used to treat BD may
to carry out work-role functions during 30% of assessed
have secondary effects on cognition and functioning.
months, a value significantly higher than patients with
unipolar depression (21%; Judd et al., 2008). Therefore,
although both affective disorders have wide-ranging
ACC E L E R AT E D AG I NG
clinical and economic implications, there are key fea-
tures that distinguish them from one another in that In the small number of samples restricted to older adults
BD appears to be associated with more persistent with BD, the association between cognitive ability and
and pronounced patterns of cognitive and functional functional outcome appears to be somewhat stronger.
impairment. Normal aging is, of course, associated with decreased

C ogniti v e I mpairments and E v eryday D isability in B ipolar I llness   •   19 5


cognitive performance and reduced everyday functional with BD have at least one other Axis I diagnosis, it is par-
abilities. There is some short-term evidence that cognitive ticularly challenging to determine the extent to which any
impairments worsen at a faster rate in middle-aged and one cluster of symptoms interacts with functional outcome.
older patients with BD than age-matched healthy com-
parators (Dhingra & Rabins, 1991; Gildengers et al., 2009),
M E DIC A L COMOR BI DI T Y
with a number of authors describing BD as akin to a neu-
rodegenerative illness. Because of the design of these stud- In BD, the presence of comorbid medical illnesses that
ies, it is unclear whether functional deficits accumulate at a may affect cognition are ubiquitous, yet is often underrec-
faster rate than in normal aging. Concurrently, there is also ognized. The strong connection between BD and medical
evidence that the severity and frequency of manic symp- comorbidity can largely be attributed to common risk fac-
toms are diminished in older patients with BD, and there is tors inherent to both, such as insufficient access to health
also strong indication that substance abuse comorbidity is care, lower socioeconomic status, poor medical adher-
about half as common as among younger adults. Thus, given ence, higher rates of smoking and substance abuse, lack of
the lessening of some of the other risk factors for functional exercise, and poor diet. As such, the prevalence of chronic
impairment with aging, limited evidence suggests that fatigue syndrome, migraines, asthma, chronic bronchitis,
older age may be associated with a stronger link between hypertension, and gastric ulcer have all been reported to
cognitive impairment and functional outcome in BD. be significantly higher in BD groups and have also been
shown to have deleterious and reciprocal effects on func-
tional outcomes, including less successful employment,
MO OD A N D P S YCHOT IC S Y M P TOM S
greater dependency on others for assistance, and higher
The National Institute of Mental Health Collaborative utilization of mental health care services (McIntyre et al.,
Depression Study and others have shown that at least 40% 2006). Additional common medical conditions found in
of patients’ time is spent experiencing depressive symptoms patients with BD include hyperlipidemia (23%), type 2
and that these symptoms are a stronger driver of functional diabetes (17%), obesity (12%), and infectious diseases such
impairments than are manic symptoms (which are present as hepatitis C (6%; Kilbourne et  al., 2004). The interac-
only 10% of the time; Judd et al., 2005). There is somewhat tions between medical comorbidity, decrements in cog-
inconsistent evidence that cognitive impairments are wors- nition, and functional domain trajectories will continue
ened during depressive phases of the illnesses, with perhaps to be an important area of further investigation, as there
more evidence for impairment in selected domains during is currently a large gap in studies that specifically exam-
the manic phase. ine the reciprocal relationships among these three closely
A key question is whether the impact of cognition and related factors.
affective symptoms are additive or interactive; few studies
are available to help address this question. Our group has
previously shown by using multivariate path analytic models L I M I TAT ION S OF   E X TA N T
that depression appears to exert an independent unmediated R E S E A RC H A N D S UG G E S T ION S
effect on interpersonal and work skills. In addition, the rela-
tive impact of these associations may not be monotonic. For
S A M PL E S I Z E A N D
example, in a sample of outpatients with BD, neurocognitive
R E S E A RCH DE S IG N
ability was the strongest predictor of obtaining any work,
whereas depression was most associated with poor work Most of the extant knowledge on cognitive ability and func-
performance among those who do work (Depp, Mausbach, tioning in BD derives from cross-sectional studies with a
Bowie et al., 2012). The implication of this finding is that modest sample size (n ~60). As such, there is currently lim-
cognitive enhancement would not improve work perfor- ited information about moderators, such as diagnostic het-
mance and reducing depression would not increase the erogeneity or comorbid factors. There is also little known
likelihood of finding work. Additional symptom and comor- about the short- or long-term trajectories of cognitive
bidity drivers of functional impairment include comorbidi- impairment and functional disability or the mechanisms by
ties such as anxiety (lifetime diagnosis ~55%; Simon et al., which cognitive abilities impact functioning. Study designs
2004), substance abuse (lifetime diagnosis ~50%; Cassidy, that model dynamic interactions between symptoms and
Ahearn, & Carroll, 2001), and psychotic features (lifetime cognitive impairments may account for greater variation in
diagnosis ~50%; Keck, 2003). Given that 90% of patients functioning (Bowie et al. 2010).

19 6   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
S TA N DA R DI Z AT ION OF  life-long unemployed patients performed significantly
CO G N I T I V E A S S E S SM E N T more poorly than the employed patients on direct assess-
ments of cognitive and functional skills, suggesting that
Improvement in the measures used to quantify cognitive
their inflated self-assessments may be related to cognitive
impairment and everyday functioning would also con-
impairments.
tribute to this literature. Unlike in schizophrenia and the
In fact, in other neuropsychiatric conditions, poor
MATRICS battery, there is no corollary consistent bat-
scores on the index of interest, either everyday functioning
tery of cognitive tests for BD. Recent data and consensus
or performance-based measures, is typically the best pre-
has indicated that the MATRICS Consensus Cognitive
dictor of overestimation of functioning. Individuals with
Battery is largely appropriate for BD, with possible inclu-
disinhibition syndromes in multiple sclerosis report that
sion of additional executive functioning tasks and the use
they are more capable in domains of cognitive functioning
of the more challenging California Verbal Learning Test.
as compared to their actual functioning (Carone, Benedict,
In addition, the inclusion of instruments that bridge cog-
Munschauer, Fishman, & Weinstock-Guttman, 2005), with
nitive and affective domains (e.g., impulsivity, theory of
similar findings in traumatic brain injury (Spikman &
mind) may predict greater proportion of variation in every-
van der Naalt, 2010). This is a potentially trivial finding, in
day functioning, although there are considerable reserva-
that if performance is very poor, the only error in estima-
tions about the measurement properties of many of these
tion that can be made is overestimation. However, multiple
measures (Pinkham et al., 2013).
other predictive factors have emerged. In particular, both
psychotic and negative symptoms predict overestimation
E M E RG I NG DI R E C T ION S I N   T H E of functioning in people with schizophrenia (Sabbag et al.,
M E A S U R E M E N T OF   DI S A B I L I T Y 2012). Across all of these conditions, and particularly rel-
evant to bipolar illness, the presence of symptoms of depres-
There are a number of advances in the measurement of sion has a substantial role in the accuracy of self-assessment
everyday functioning that have been introduced initially of abilities and potential across an array of neuropsychiatric
in research on aging and other severe mental illnesses such conditions and in the healthy population.
as schizophrenia. These include considerations of the valid- Healthy people routinely overestimate their capabili-
ity of different reports of functioning, such as self-reports ties and the likelihood of positive outcomes in ambiguous
and the reports of other informants such as caregivers. In situations, referred to as the “optimistic” bias (Ehrlinger &
addition, remote assessment has advanced to a considerable Dunning, 2003). In contrast, the presence of mild depres-
extent in these other areas, including both Internet- and sion is associated with increased accuracy of self assessment,
smartphone-based assessments. These areas of investigation with this phenomenon referred to as “depressive realism”
are just beginning to develop in bipolar illness, and some (Dunning & Story, 1991). In healthy people, deflating feed-
preliminary research has suggested that some of the same back increases the extent to which self-assessments become
factors are operative in BD. more congruent with objective reality (Ehrlinger, Johnson,
Banner, Dunning, & Kruger, 2008). Bipolar disorder is
perhaps the condition most likely to see variable ability to
I M PA I R E D S E L F -A S S E S SM E N T
perform adequate self-assessments across the course of ill-
Neuropsychiatric conditions accompanied by impairments ness, as individuals with both BD I and II experience mood
on performance-based cognitive assessments also appear shifts ranging from elevated to depressed mood.
to include substantial deficits in the ability to accurately Although there has been little research on this topic,
evaluate current levels of functioning. These impairments the results in patients with BD seem quite consistent with
include self-assessments of cognitive abilities (Burdick, results from other conditions. Burdick et al. (2005) evalu-
Endick, & Goldberg, 2005; Keefe, Poe, Walker, Kang, & ated 37 BD patients using three different ratings of cognitive
Harvey, 2006) and everyday functional skills. For instance, impairment by patient self-report, as well as a short objective
in a recent study of patients with schizophrenia, individu- battery of neuropsychological tests. Patients’ self-reports
als who had never worked in any type of employment at correlated poorly with objectively measured cognitive defi-
any time in their lives rated themselves as more competent cits, and, although not statistically significant, there was a
at vocational skills than people with schizophrenia with clear trend toward mood-state effect on self-report measures.
current full-time employment (Gould, Sabbag, Durand, Specifically, depression ratings were positively correlated
Patterson, & Harvey, 2013). At the same time, these with cognitive complaints while mania ratings coincided

C ogniti v e I mpairments and E v eryday D isability in B ipolar I llness   •   19 7


with fewer deficits by patient report, suggesting a possible individuals who have never written a check or completed
affective bias in self-reports. Another study of 60 euthymic a deposit slip who are actually quite skilled at banking
BD subjects found that a subset of BD patients were able through the use of ATMs, computerized account manage-
to accurately identify their cognitive problems by self-report ment, and even smartphone banking apps.
and that these patients who had better insight into their A solution to the problem of outdated and indirect
deficits had higher levels of subsyndromal depression, with assessments is the development of realistic simulations of
an increased number of prior episodes and an earlier age of functionally skilled tasks. Such simulations have been used
onset as compared with the BD patients who were unaware previously for both assessment, such as to examine the abil-
of their objective impairment (Martinez-Aran et al., 2005). ity of older healthy people to manage ATM and Internet
Consistent with this, BD patients’ own assessment of func- technology, and for treatment, such as through the develop-
tional capacity and quality of life may also be influenced ment of virtual reality simulations of classrooms and com-
by their affective state. In 90 BD subjects, Piccinni et  al. bat situations. These simulations have the advantages of
(2007) reported that partially remitted manic patients often being a complete replica of real-world demands such as
reported a significantly better quality of life across all in banking, telephone voice menus, or Internet bill-paying
domains of function compared with the partially remitted scenarios. They can be placed into the homes of individuals
BD depressed group, despite nearly identical clinician-based who do not have a computer of their own, and they can be
ratings of global functioning. These cross-sectional stud- administered in laboratory settings. The disadvantages of
ies suggest that the changes in mood state that are associ- the method are that some individuals may have never had
ated with the lifetime course of BD may not only influence computer interactions, meaning that the results may not be
real-world functioning but also may influence the accuracy applicable for individuals who have avoided computer inter-
of self-assessments of current functioning, highlighting the action to date. This concern may be somewhat overstated,
importance of formal evaluation. as survey results indicated that 91% of individuals under 50
and 53% of individuals over 65 were regular Internet users
as of 2012. In individuals over age 65, regular Internet use
increased from 10% in 2000 to the current level in 2012
COM PU T E R I Z E D A S S E S SM E N T
(Zickuhr & Madden, 2012). Thus use of realistic simula-
A N D ECOL O G IC A L
tion technology to assess functional capacity in people with
MOM E N TA RY A S S E S SM E N T
bipolar illness is clearly on the horizon.
Given the questions raised about the validity of self-reports Ecological momentary assessment (EMA) is an addi-
of cognitive functioning and disability, it is clearly important tional technology with high potential. EMA technology
to utilize direct assessments of ability to accurately estimate currently employs smartphones, which can contact the
future potential. At the same time, even performance-based participant, obtain answers to a series of questions, and
assessments may have limitations. Neuropsychological tests directly transmit the information to a central database. In
have been shown in BD to be indirectly related to real-world contrast to older monitoring strategies, such as diaries, the
functioning when functional capacity measures are also col- EMA technology allows for time stamping of information
lected (Bowie et al., 2010; Mausbach et al., 2010). This may and delivery of reminders to increase adherence. Current
not be surprising because neuropsychological tests measure smartphones also have GPS features, which allow for moni-
relatively general skills that may be required for the per- toring of activity. These technologies can also be used to
formance of skilled acts in the environment, although not collect multidimensional information, including mood
specifically overlapping with the content of those skills. state, current activity, others present, and level of assistance
Even current functional capacity measures have some limi- being received or provided. As a result, their direct and
tations. For instance, given the rapid development of tech- momentary information far surpasses self-reported infor-
nology, many of the tasks in commonly used functional mation, particularly over extended periods such as the typi-
capacity measures are now less commonly performed. For cal one-month period between clinical visits.
instance, the UCSD Performance-Based Skills Assessment EMA assessment is feasible, in that in our previous
has a subtest that measures the ability to write a check, and research we found that more than 75% of bipolar patients
the Advanced Finances subtest of the Everyday function- already had a smartphone. Moreover, adherence to moni-
ing battery requires wiring checks, making bank deposits toring requests was in the 90% range, and smartphones
with a deposit slip, and maintaining a balance. With the provided to research participants were returned over 95%
changes in banking technology in the past decade, there are of the time.

19 8   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
F U NC T ION A L A N D C O G N I T I V E as a determinant of functional outcome can be transported
R E H A B I L I TAT ION from schizophrenia to BD. As in schizophrenia, there does
not appear to be strong evidence that any one cognitive
From these facts it stands to reason that cognitive deficits domain causes greater impairment, functional measurement
are a potential treatment target for functional rehabilita- with objective performance-based measure appears to more
tion in BD (Green, 2006; Harvey et  al., 2010). It is also strongly associated with cognition than other measurement
clear that functional outcome should be a secondary out- modalities, and cognitive rehabilitation may have a distal
come in efforts to improve cognitive functioning. In schizo- positive impact on functioning. However, the influence of
phrenia, the Food and Drug Administration (Buchanan self-assessment, mood symptoms, and comorbidities add to
et al., 2010) has provided strong guidance that functional the complexity of cognitive and functional impairments in
outcome should be an end-point in clinical trials of agents BD, and so it is clear that longitudinal research would be
targeting cognitive enhancement; it is likely that people needed to understand the dynamic and lifespan influences
with BD will be provided cognitive enhancing medica- among cognitive and functional markers. Finally, it does not
tions approved for schizophrenia, and it will be essential seem premature to develop and evaluate cognitive remedia-
to determine if the benefits of these extend to functional tion strategies in order to improve functional outcome in
outcome in BD. BD. Ultimately, these adjunctive therapies targeting neu-
There have been few attempts at improving cognitive rocognitive impairment may lead to long-lasting improve-
functioning in BD. One of the few studies implementing ments in the everyday functioning and quality of life for the
cognitive remediation treatment in an effort to improve millions of people suffering from BD.
neurocognitive abilities (i.e., difficulties with organiza-
Disclosure statement: Alexandrea Harmell has no conflicts
tion, planning attention, and memory), as well as simulta-
of interest. Over the past three years Dr. Harvey has served
neously treat residual depressive symptoms in BD, found
as a consultant to Abbvie, Boeheringer-Ingelheim, Forest
decreased residual symptoms and increased occupational
Labs, Forum (Formerly En Vivo), Genentech, Lundbeck,
and overall psychosocial functioning following treatment
Otsuka America, Roche Pharma, Sunovion, Takeda, and
and at three-month follow-up (Deckersbach et  al., 2010).
Teva Pharma. He received contract research support from
Another study looking within patients with BD in the
Genentech. Dr. Colin A. Depp has not disclosed any con-
depressive phase of the disorder found similar improve-
flicts of interest.
ments in depression and reports of less difficulty in every-
day coping following an eight-week computer-based online
cognitive training program (Preiss, Shatil, Cermakova,
Cimermanova, & Ram, 2013). Consistent with these R E F E R E NC E S
results, a large multicenter, randomized, rater-blind clinical
trial comparing a novel functional remediation interven- Andreou, C., & Bozikas, V.  P. (2013). The predictive significance of
neurocognitive factors for functional outcome in bipolar disorder.
tion with psychoeducation and treatment as usual found Current Opinions in Psychiatry, 26(1), 54–59.
that the functional remediation intervention showed supe- Berk, M., Conus, P., Kapczinski, F., Andreazza, A. C., Yucel, M., Wood,
riority over treatment as usual in improving both occupa- S. J., . . . McGorry, P. D. (2010). From neuroprogression to neuropro-
tection: implications for clinical care. Medical Journal of Australia,
tional and interpersonal functioning (Torrent et al., 2013). 193(4 Suppl.), S36–S40.
Given these promising findings supporting the connection Bonnin, C.  M., Martinez-Aran, A., Torrent, C., Pacchiarotti, I.,
between improved cognition and functional outcomes, Rosa, A.  R., Franco, C., . . . Vieta, E. (2010). Clinical and neuro-
cognitive predictors of functional outcome in bipolar euthymic
efforts to continue to develop and test functional and cog- patients:  A  long-term, follow-up study. Journal of Affective Disor-
nitive rehabilitation programs are very much warranted. ders, 121(1–2), 156–160.
Bowie, C. R., Depp, C., McGrath, J. A., Wolyniec, P., Mausbach, B. T.,
Thornquist, M. H., . . . Pulver, A. E. (2010). Prediction of real-world
functional disability in chronic mental disorders: A comparison of
C ONC LUS ION schizophrenia and bipolar disorder. American Journal of Psychiatry,
167(9), 1116–1124.
Bowie, C.  R., & Harvey, P.  D. (2005). Cognition in schizophre-
Bipolar disorder is a highly disabling illness for most patients, nia:  Impairments, determinants, and functional importance. Psy-
and it is becoming clear that cognitive impairments are a chiatric Clinics of North America, 28(3), 613–633, 626.
major source of this disability. Even though BD produces Bowie, C. R., Twamley, E. W., Anderson, H., Halpern, B., Patterson,
T. L., & Harvey, P. D. (2007). Self-assessment of functional status in
less disability, on average, than schizophrenia, many find- schizophrenia. Journal of Psychiatric Research, 41(12), 1012–1018.
ings of the more established body of literature on cognition doi:10.1016/j.jpsychires.2006.08.003

C ogniti v e I mpairments and E v eryday D isability in B ipolar I llness   •   19 9


Breslau, J., Miller, E., Jin, R., Sampson, N.  A., Alonso, J., Andrade, Fett, A. K., Viechtbauer, W., Dominguez, M. D., Penn, D. L., van Os,
L.  H., . . . Kessler, R.  C. (2011). A multinational study of mental J., & Krabbendam, L. (2011). The relationship between neurocog-
disorders, marriage, and divorce. Acta Psychiatrica Scandinavica, nition and social cognition with functional outcomes in schizo-
124(6), 474–486. phrenia: A meta-analysis. Neuroscience and Biobehavioral Reviews,
Buchanan, R. W., Keefe, R. S., Umbricht, D., Green, M. F., Laughren, 35(3), 573–588. doi:S0149-7634(10)00114-4 [pii] 10.1016/j.neu-
T., & Marder, S.  R. (2010). The FDA-NIMH-MATRICS guide- biorev.2010.07.001
lines for clinical trial design of cognitive-enhancing drugs:  What Gildengers, A. G., Butters, M. A., Chisholm, D., Anderson, S. J., Begley,
do we know 5 years later? Schizophrenia Bulletin, 37(6), 1209–1217. A., Holm, M., . . . Mulsant, B. H. (2012). Cognition in older adults
doi:10.1093/schbul/sbq038 with bipolar disorder versus major depressive disorder. Bipolar Dis-
Burdick, K. E., Endick, C. J., & Goldberg, J. F. (2005). Assessing cog- orders, 14(2), 198–205.
nitive deficits in bipolar disorder: Are self-reports valid? Psychiatry Gildengers, A. G., Chisholm, D., Butters, M. A., Anderson, S. J., Beg-
Research, 136(1), 43–50. ley, A., Holm, M., . . . Mulsant, B. H. (2013). Two-year course of cog-
Burdick, K.  E., Goldberg, J.  F., & Harrow, M. (2010). Neurocogni- nitive function and instrumental activities of daily living in older
tive dysfunction and psychosocial outcome in patients with bipo- adults with bipolar disorder: evidence for neuroprogression? Psycho-
lar I disorder at 15-year follow-up. Acta Psychiatrica Scandinavica, logical Medicine, 43(4), 801–811.
122(6), 499–506. Gildengers, A. G., Mulsant, B. H., Begley, A., Mazumdar, S., Hyams,
Carone, D. A., Benedict, R. H., Munschauer, F. E. III, Fishman, I., & A. V., Reynolds Iii, C. F., . . . Butters, M. A. (2009). The longitudi-
Weinstock-Guttman, B. (2005). Interpreting patient/informant nal course of cognition in older adults with bipolar disorder. Bipolar
discrepancies of reported cognitive symptoms in MS. Journal of the Disorders, 11(7), 744–752.
International Neuropsychological Society, 11(5), 574–583. Goldberg, J.  F., & Chengappa, K.  N. (2009). Identifying and treat-
Cassidy, F., Ahearn, E. P., & Carroll, B. J. (2001). Substance abuse in ing cognitive impairment in bipolar disorder. Bipolar Disorders, 2,
bipolar disorder. Bipolar Disorders, 3(4), 181–188. 123–137.
Dean, B. B., Gerner, D, & Gerner, R. H. (2004). A systematic review Gould, F., Sabbag, S., Durand, D., Patterson, T.  L., & Harvey, P.  D.
evaluating health-related quality of life, work impairment, and (2013). Self-assessment of functional ability in schizophrenia: mile-
healthcare costs and utilization in bipolar disorder. Current Medi- stone achievement and its relationship to accuracy of self-evaluation.
cal Research and Opinion, 20, 139–154. Psychiatry Research, 207(1–2), 19–24.
Deckersbach, T., Nierenberg, A. A., Kessler, R., Lund, H. G., Ametrano, Green, M. F. (2006). Cognitive impairment and functional outcome in
R.  M., Sachs, G., . . . Dougherty, D. (2010). RESEARCH:  Cogni- schizophrenia and bipolar disorder. Journal of Clinical Psychiatry,
tive rehabilitation for bipolar disorder: An open trial for employed 67(10), e12.
patients with residual depressive symptoms. CNS Neuroscience & Harvey, P. D., Wingo, A. P., Burdick, K. E., & Baldessarini, R. J. (2010).
Therapeutics, 16(5), 298–307. Cognition and disability in bipolar disorder:  lessons from schizo-
Depp, C.  A., Mausbach, B.  T., Bowie, C., Wolyniec, P., Thornquist, phrenia research. Bipolar Disorders, 12(4), 364–375.
M. H., Luke, J. R., . . . Patterson, T. L. (2012). Determinants of occu- Jabben, N., Arts, B., van Os, J., & Krabbendam, L. (2010). Neurocogni-
pational and residential functioning in bipolar disorder. Journal of tive functioning as intermediary phenotype and predictor of psy-
Affective Disorders, 136(3), 812–818. chosocial functioning across the psychosis continuum:  Studies in
Depp, C. A., Mausbach, B. T., Eyler, L. T., Palmer, B. W., Cain, A. E., schizophrenia and bipolar disorder. Journal of Clinical Psychiatry,
Lebowitz, B. D., . . . Jeste, D. V. (2009). Performance-based and sub- 71(6), 764–774.
jective measures of functioning in middle-aged and older adults Jones, S. H., Thornicroft, G., Coffey, M., & Dunn, G. (1995). A brief
with bipolar disorder. Journal of Nervous and Mental Disease, mental health outcome scale—Reliability and validity of the Global
197(7), 471–475. Assessment of Functioning (GAF). The British Journal of Psychiatry,
Depp, C.  A., Mausbach, B.  T., Harmell, A.  L., Savla, G.  N., Bowie, 166(5), 654.
C. R., Harvey, P. D., & Patterson, T. L. (2012). Meta-analysis of the Judd, L. L., & Akiskal, H. S. (2003). Depressive episodes and symptoms
association between cognitive abilities and everyday functioning in dominate the longitudinal course of bipolar disorder. Current Psy-
bipolar disorder. Bipolar Disorders, 14(3), 217–226. doi:10.1111/ chiatry Reports, 5(6), 417–418.
j.1399-5618.2012.01011.x Judd, L. L., Akiskal, H. S., Schettler, P. J., Endicott, J., Leon, A. C.,
Dhingra, U., & Rabins, P. V. (1991). Mania in the elderly: A 5-7 year Solomon, D.  A., . . . Keller, M.  B. (2005). Psychosocial disability
follow-up. Journal of the American Geriatrics Society, 39(6), in the course of bipolar I  and II disorders:  A  prospective, com-
581–583. parative, longitudinal study. Archives of General Psychiatry, 62(12),
Dickinson, D., & Harvey, P. D. (2009). Systemic hypotheses for gen- 1322–1330.
eralized cognitive deficits in schizophrenia: A new take on an old Judd, L.  L., Schettler, P.  J., Akiskal, H., Coryell, W., Fawcett, J., Fie-
problem. Schizophrenia Bulletin, 35(2), 403–414. dorowicz, J. G., . . . Keller, M. B. (2012). Prevalence and clinical sig-
Dilsaver, S. C. (2011). An estimate of the minimum economic burden nificance of subsyndromal manic symptoms, including irritability
of bipolar I and II disorders in the United States: 2009. Journal of and psychomotor agitation, during bipolar major depressive epi-
Affective Disorders, 129(1–3), 79–83. doi:10.1016/j.jad.2010.08.030 sodes. Journal of Affective Disorders, 138(3), 440–448.
Dunning, D., & Story, A. L. (1991). Depression, realism, and the overcon- Judd, L. L., Schettler, P. J., Solomon, D. A., Maser, J. D., Coryell, W.,
fidence effect: Are the sadder wiser when predicting future actions and Endicott, J., & Akiskal, H.  S. (2008). Psychosocial disability and
events? Journal of Personality and Social Psychology, 61(4), 521–532. work role function compared across the long-term course of bipo-
Ehrlinger, J., & Dunning, D. (2003). How chronic self-views influence lar I, bipolar II and unipolar major depressive disorders. Journal of
(and potentially mislead) estimates of performance. Journal of Per- Affective Disorders, 108(1–2), 49–58.
sonality and Social Psychology, 84(1), 5–17. Keck, P. E. (2003). Psychosis in bipolar disorder: Phenomenology and
Ehrlinger, J., Johnson, K., Banner, M., Dunning, D., & Kruger, J. impact on morbidity and course of illness. Comprehensive Psychia-
(2008). Why the unskilled are unaware:  Further explorations of try, 44(4), 263–269.
(absent) self-insight among the incompetent. Organizational Behav- Keefe, R. S., Poe, M., Walker, T. M., Kang, J. W., & Harvey, P. D. (2006).
ior and Human Decision Processes, 105(1), 98–121. The Schizophrenia Cognition Rating Scale:  An interview-based
Endicott, J., Spitzer, R. L., Fleiss, J. L., & Cohen, J. (1976). The Global assessment and its relationship to cognition, real-world function-
Assessment Scale: A procedure for measuring overall severity of psy- ing, and functional capacity. American Journal of Psychiatry, 163(3),
chiatric disturbance. Archives of General Psychiatry, 33(6), 766. 426–432.

2 0 0   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
Kilbourne, A. M., Cornelius, J. R., Han, X., Pincus, H. A., Shad, M., disorder:  A  study of cognitive functioning. Frontiers in Human
Salloum, I., . . . Haas, G. L. (2004). Burden of general medical con- Neuroscience, 7, 108.
ditions among individuals with bipolar disorder. Bipolar Disorders, Reichenberg, A., Weiser, M., Rabinowitz, J., Caspi, A., Schmeidler,
6(5), 368–373. J., Mark, M., . . . Davidson, M. (2002). A population-based cohort
Krabbendam, L., Arts, B. M., Van Os, J., & Aleman, A. (2005). Cog- study of premorbid intellectual, language, and behavioral func-
nitive functioning in patients with schizophrenia and bipolar tioning in patients with schizophrenia, schizoaffective disorder,
disorder:  A  quantitative review. Schizophrenia Research, 80(2–3), and nonpsychotic bipolar disorder. American Journal of Psychiatry,
137–149. 159(12), 2027–2035. doi:10.1176/appi.ajp.159.12.2027
Laxman, K. E., Lovibond, K. S., & Hassan, M. K. (2008). Impact of Sabbag, S., Twamley, E. M., Vella, L., Heaton, R. K., Patterson, T. L., &
bipolar disorder in employed populations. American Journal of Harvey, P. D. (2011). Assessing everyday functioning in schizophre-
Managed Care, 14(11), 757–764. nia:  Not all informants seem equally informative. Schizophrenia
Leifker, F. R., Patterson, T. L., Heaton, R. K., & Harvey, P. D. (2011). Research, 131(1–3), 250–255.
Validating measures of real-world outcome:  The results of the Sabbag, S., Twamley, E. W., Vella, L., Heaton, R. K., Patterson, T. L., &
VALERO expert survey and RAND panel. Schizophrenia Bulletin, Harvey, P. D. (2012). Predictors of the accuracy of self assessment of
37(2), 334–343. doi:10.1093/schbul/sbp044 everyday functioning in people with schizophrenia. Schizophrenia
Lyoo, I. K., Sung, Y. H., Dager, S. R., Friedman, S. D., Lee, J. Y., Kim, Research, 137(1–3), 190–195. doi:10.1016/j.schres.2012.02.002
S. J., . . . Renshaw, P. F. (2006). Regional cerebral cortical thinning in Schneider, L. C., & Struening, E. L. (1983). SLOF: A behavioral rat-
bipolar disorder. Bipolar Disorders, 8(1), 65–74. ing scale for assessing the mentally ill. Social Work Research and
Maalouf, F. T., Klein, C., Clark, L., Sahakian, B. J., Labarbara, E. J., Ver- Abstracts, 19(3), 9–21.
sace, A., . . . Phillips, M. L. (2010). Impaired sustained attention and Schoeyen, H. K., Birkenaes, A. B., Vaaler, A. E., Auestad, B. H., Malt,
executive dysfunction:  bipolar disorder versus depression-specific U.  F., Andreassen, O.  A., & Morken, G. (2011). Bipolar disorder
markers of affective disorders. Neuropsychologia, 48(6), 1862–1868. patients have similar levels of education but lower socio-economic
Martinez-Aran, A., Vieta, E., Colom, F., Torrent, C., Reinares, M., status than the general population. Journal of Affective Disorders,
Goikolea, J.  M., . . . Sanchez-Moreno, J. (2005). Do cognitive com- 129(1–3), 68–74. doi:10.1016/j.jad.2010.08.012
plaints in euthymic bipolar patients reflect objective cognitive Simon, N. M., Otto, M. W., Wisniewski, S. R., Fossey, M., Sagduyu, K.,
impairment? Psychotherapy and Psychosomatics, 74(5), 295–302. Frank, E., . . . Pollack, M. H. (2004). Anxiety disorder comorbidity
Martino, D. J., Marengo, E., Igoa, A., Scapola, M., Ais, E. D., Perinot, in bipolar disorder patients: Data from the first 500 participants in
L., & Strejilevich, S.  A. (2009). Neurocognitive and symptom- the Systematic Treatment Enhancement Program for Bipolar Disor-
atic predictors of functional outcome in bipolar disorders:  A  pro- der (STEP-BD). The American Journal of Psychiatry, 161(12), 2222.
spective 1  year follow-up study. Journal of Affective Disorders, Soderberg, P., Tungstrom, S., & Armelius, B. A. (2005). Special section
116(1–2), 37–42. on the GAF: Reliability of Global Assessment of Functioning rat-
Marwaha, S., Durrani, A., & Singh, S. (2013). Employment outcomes ings made by clinical psychiatric staff. Psychiatric Services, 56(4),
in people with bipolar disorder: A systematic review. Acta Psychiat- 434–438. doi:10.1176/appi.ps.56.4.434
rica Scandinavica, 4(10), 12087. Spikman, J.  M., & van der Naalt, J. (2010). Indices of impaired
Mausbach, B. T., Harvey, P. D., Pulver, A. E., Depp, C. A., Wolyniec, self-awareness in traumatic brain injury patients with focal frontal
P. S., Thornquist, M. H., . . . Patterson, T. L. (2010). Relationship of lesions and executive deficits:  Implications for outcome measure-
the Brief UCSD Performance-based Skills Assessment (UPSA-B) ment. Journal of Neurotrauma, 27(7), 1195–1202.
to multiple indicators of functioning in people with schizophrenia Stang, P., Frank, C., Ulcickas Yood, M., Wells, K., & Burch, S.
and bipolar disorder. Bipolar Disorders, 12(1), 45–55. (2007). Impact of bipolar disorder: Results from a screening study.
McIntyre, R.  S., Konarski, J.  Z., Soczynska, J.  K., Wilkins, K., Pan- Primary Care Companion to the Journal of Clinical Psychiatry,
jwani, G., Bouffard, B., . . . Kennedy, S. H. (2006). Medical comor- 9(1), 42–47.
bidity in bipolar disorder: Implications for functional outcomes and Strakowski, S. M., DelBello, M. P., Zimmerman, M. E., Getz, G. E.,
health service utilization. Psychiatric Services, 57(8), 1140–1144. Mills, N. P., Ret, J., . . . Adler, C. M. (2002). Ventricular and periven-
Mora, E., Portella, M.  J., Forcada, I., Vieta, E., & Mur, M. (2013). tricular structural volumes in first- versus multiple-episode bipolar
Persistence of cognitive impairment and its negative impact on disorder. The American Journal of Psychiatry, 159(11), 1841–1847.
psychosocial functioning in lithium-treated, euthymic bipolar Tohen, M., Zarate, C.  A. Jr., Hennen, J., Khalsa, H.  M., Stra-
patients:  A  6-year follow-up study. Psychological Medicine, 43(6), kowski, S.  M., Gebre-Medhin, P., . . . Baldessarini, R.  J. (2003).
1187–1196. The McLean-Harvard First-Episode Mania Study:  Prediction of
Mur, M., Portella, M.  J., Martinez-Aran, A., Pifarre, J., & Vieta, E. recovery and first recurrence. The American Journal of Psychiatry,
(2008). Long-term stability of cognitive impairment in bipolar dis- 160(12), 2099–2107.
order: A 2-year follow-up study of lithium-treated euthymic bipolar Torrent, C., Del Mar Bonnin, C., Martinez-Aran, A., Valle, J.,
patients. Journal of Clinical Psychiatry, 69(5), 712–719. Amann, B. L., Gonzalez-Pinto, A., . . . Vieta, E. (2013). Efficacy of
Nuechterlein, K.  H., Green, M.  F., Kern, R.  S., Baade, L.  E., Barch, functional remediation in bipolar disorder: A multicenter random-
D. M., Cohen, J. D., . . . Marder, S. R. (2008). The MATRICS Con- ized controlled study. The American Journal of Psychiatry, 20(10),
sensus Cognitive Battery, Part  1:  Test selection, reliability, and 12070971.
validity. American Journal of Psychiatry, 165(2), 203–213. Torres, I. J., DeFreitas, C. M., DeFreitas, V. G., Bond, D. J., Kunz, M.,
Piccinni, A., Catena, M., Del Debbio, A., Marazziti, D., Monje, C., Honer, W. G., . . . Yatham, L. N. (2011). Relationship between cog-
Schiavi, E., . . . Dell’Osso, L. (2007). Health-related quality of life nitive functioning and 6-month clinical and functional outcome in
and functioning in remitted bipolar I  outpatients. Comprehensive patients with first manic episode bipolar I  disorder. Psychological
Psychiatry, 48(4), 323–328. Medicine, 41(5), 971–982.
Pinkham, A. E., Penn, D. L., Green, M. F., Buck, B., Healey, K., & Har- Zarate, C. A. Jr., Tohen, M., Land, M., & Cavanagh, S. (2000). Func-
vey, P.  D. (2013). The Social Cognition Psychometric Evaluation tional impairment and cognition in bipolar disorder. Psychiatric
Study: Results of the expert survey and RAND panel. Schizophrenia Quarterly, 71(4), 309–329.
Bulletin, 31, 31. Zickuhr, K., & Madden, M. (June 6, 2012). Older adults and Internet
Preiss, M., Shatil, E., Cermakova, R., Cimermanova, D., & Ram, I. use. Retrieved from http://www.pewinternet.org/Reports/2012/
(2013). Personalized cognitive training in unipolar and bipolar Older-adults-and-internet-use.aspx

C ogniti v e I mpairments and E v eryday D isability in B ipolar I llness   •   2 01


PA R T   I V

N EU ROI M AG I NG
17.
STRUCTUR A L A ND FU NCTIONA L M AGNETIC
R ESONA NCE IM AGING FINDINGS A ND
THEIR TR EATMENT IMPACT
Marguerite Reid Schneider, David E. Fleck, James C. Eliassen,

and Stephen M. Strakowski

INTRODUCTION S T RUC T U R A L M AG N E T IC
R E S ON A NC E I M AG I NG A PPROACH E S
Magnetic resonance imaging (MRI) has been instrumen-
Structural MRI (sMRI) consists of a set of techniques to
tal in helping to define the neuropathology of bipolar
examine the anatomy of the body, including the brain.
disorder at the level of brain networks. This work has pro-
Unlike computed tomography (CT), sMRI does not
vided potential biological markers to potentially predict
use ionizing radiation. Instead it relies on the different
the onset of disease, disease chronicity, and the likeli-
magnetic properties of various tissue types to generate a
hood of treatment response. Moreover, new MRI-based
three-dimensional static image. There are two primary
imaging techniques, including diffusion tensor imaging
approaches to analyzing sMRI images of the brain: region
(DTI), and new analysis approaches, such as functional
of interest (ROI) and voxel based morphometry (VBM),
connectivity (FC), promise to further characterize bipo-
although others are being developed (Fleck et  al., 2008).
lar neuropathology in the interest of informing treat-
The use of an ROI or VBM approach depends, in part, on
ment. However, the validity of neuroimaging effects
the research question at hand, the study design, and the
identified in bipolar disorder to date is almost universally
inherent advantages and disadvantages of each approach.
compromised by confounding factors such as medica-
tion exposure. Untangling the independent influences
of genetics, symptomatology, and medication exposure Region of interest analysis
on brain structure and function is a primary goal of psy- Historically, sMRI image processing involved isolating
chiatric imaging research at the beginning of this new macroscopic brain morphology (specific brain regions or
millennium. Toward this end, this chapter reviews MRI structures of interest) by manually “tracing” them on the
techniques, approaches to image analysis, current struc- computer screen. The measurements are obtained by manu-
tural and functional findings, treatment effects, and ally tracing each region/structure in consecutive image
predictors of treatment response as they relate to bipolar “slices” based on anatomical landmarks and then averaging
disorder at the beginning of the 21st century. It is hoped voxels (volumetric pixels). The total volume of each ROI
that this information will provide a solid foundation for is then calculated and compared among different experi-
rapid neuroimaging-based medical advances in the years mental groups or across baseline and treatment conditions.
to come, including a more definitive diagnosis of bipolar As one might imagine, manual ROI analysis is very time
disorder, a better understanding of its development and consuming and painstaking and, as such, can generally be
progression, identification of clinically useful biological used only to examine a relatively few a priori defined brain
treatment targets, and more successful evaluation of treat- regions. In addition to time constraints for tracing each
ment effects. ROI manually, the number of regions examined is further

205
constrained by the time and training required to obtain cerebrospinal fluid, and non-brain tissues). Recently, seg-
acceptable inter- and intrarater reliability for all ROI trac- mentation algorithms have been combined with metods
ings. More recently, technological advances allow semi- or for intersubject alignment (normalization) and averaging
fully automated computer algorithms in sMRI measure- to create new approaches, including volume- and shape-
ments. The use of these newer automated tracing proce- based morphometry methods (e.g., mesh mapping).
dures bypasses many time and training limitations but is Surface-based approaches also exist in which surface-of-
restricted to structures with highly demarcated edge lines interest or point-by-point comparisons can be made among
that are easily detected and, as such, is not applicable to individuals on measures such as gray matter thickness.
all possible ROIs. As an artifact of manual and automated These automated techniques provide high throughput in
tracing limitations, ROI analysis is typically employed as a single experiment, but they may assess features of the
a hypothesis-driven technique in which brain regions are MRI signal that have an inexact correspondence with spe-
prespecified and hypothesized to increase or decrease in cific structural quantities such as tissue volume or thick-
volume as a group or treatment effect. ness. They also rely heavily on the accuracy of the digital
ROI approaches offer the advantage of construct and image processing technique employed and on intersubject
face validity because they are directly linked to neuroanat- comparisons for validity. Indeed, a primary disadvantage
omy and provide absolute volumes as outcome measures. of VBM is that it is subject to errors based on the assump-
Generally, ROI assessments can be conducted accurately tion that all brains can be similarly transformed into ste-
and provide robust results. The primary drawback of this reotactic space. For instance, overall brain size differences
approach is the substantial time, effort, and cost required to identified in bipolar samples suggest that the removal of
insure reliable and valid measurement as discussed. Other overall brain size information during normalization may
limitations include variability in how regions are anatomi- be an invalid assumption. Another disadvantage is that this
cally defined (subregions of frontal cortex, for instance that method is potentially less sensitive to regional differences in
are better defined functionally than anatomically), tracer small subcortical structures relative to ROI analysis.
drift (which requires periodic recalibration training), and
individual heterogeneity of brain structures, especially in
Diffusion tensor imaging
individuals with neuropathology such as in bipolar disorder.
Diffusion tensor imaging (DTI) is an MRI-based anatomi-
cal connectivity technique that uses measurements of in
Voxel-based morphometry analysis
vivo water diffusion to examine white matter integrity.
Unlike ROI analysis, VBM examines sMRI images on a Water diffusion within white matter tracts is strongly influ-
voxel-by-voxel basis for differences in signal intensity, vol- enced by specific structural features including the myelin
ume, and shape across the whole brain. The newer VBM sheath and the tightly packed nature of axonal bundles. In
approach to data processing requires that every voxel in a the absence of any constraints on diffusion, the volume over
brain image be examined through an automated process. which water molecules diffuse is isotropic; that is, it forms
This technique detects volume differences throughout the a sphere. The structure of normal white matter, however,
brain very rapidly, despite the large quantity of data input. results in a more rapid diffusion along the axis of the tract.
Unlike the hypothesis-driven approach of ROI analysis, This relative freedom of movement in only one direction
VBM is an exploratory, hypothesis-generating technique gives rise to a nonisotropic (anisotropic) diffusion pattern,
for examining each voxel in a data array and allowing upon which various DTI measurements are built.
comparisons among corresponding voxels. Rather than Fractional anisotropy (FA) is one DTI measure derived
generating absolute volumes, as in ROI analysis, the VBM from the ratio between water molecule movements parallel
approach requires data be “normalized” into stereotactic with and perpendicular to the axonal tract; this approach
space, and, as such, between-group differences represent has been widely used by bipolar disorder researchers.
relative volumes within each voxel. Decreased FA may represent axonal pathology such as
The VBM approach has a number of distinct advantages neuropathic changes, a loss of bundle coherence (i.e., less
and disadvantages relative to ROI analysis. Its most impor- tightly packed fibers), or a disruption in axonal organiza-
tant advantage is the objective and rapid measurements tion (Foong et al., 2002). Mean diffusivity is another mea-
of whole-brain structure. It also requires relatively little sure derived from DTI of white matter tracts that has been
experimenter effort since it is highly automated (e.g., image widely employed. Increased diffusivity is thought to reflect
segmentation that distinguishes gray matter, white matter, a loss of impediments to water movement in and around

2 0 6   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
white matter tracts that has been observed with edema and Functional connectivity analysis
axonal demyelization and loss (Beaulieu, 2002).
Functional connectivity is a type of fMRI analysis that
Diffusion weighted imaging techniques, such as DTI,
looks at statistical associations between BOLD intensity
have limited spatial resolution but are useful as whole brain
time series for different brain regions. The FC approach
in vivo markers of temporal changes in fiber tracts. The
is useful to explore the brain’s functional organization
physical pattern of anatomical connections is relatively sta-
and to determine if it is altered in brain disorders such as
ble over short time periods (seconds to hours). Over longer
bipolar disorder. The term “functional connectivity” refers
periods (days to weeks), structural connectivity patterns are
to connections among brain regions that share functional
likely subject to morphological change and plasticity, which
properties. More specifically, it can be defined as temporal
may make DTI useful in treatment monitoring.
correlations between spatially remote neurophysiologi-
cal events (fMRI brain activation in this case) expressed
F U NC T IONA L M AG N E T IC as a deviation from statistical independence across these
R E S ONA NC E I M AG I NG A PPROACH E S events (Biswal, Van Kylen, & Hyde, 1997). FC analysis is
In contrast to structural brain imaging techniques, func- conducted with fMRI data collected while an individual
tional imaging allows examinations of online brain activ- is either actively performing a task (an event-related or
ity. Functional MRI (fMRI) builds on the basic principles blocked working memory task for instance) or in a rest-
of sMRI and adds a step allowing brain activation to be ing state to evaluate regional brain interactions that occur
inferred from changes in the blood-oxygenation-level when not actively engaged in an explicit task. Because brain
dependent (BOLD) response. The BOLD response occurs activity is present even in the absence of external stimuli,
because the magnetic properties of hemoglobin differ in resting state FC, any brain region can have seemingly
depending on whether the molecule is oxygenated or deox- spontaneous fluctuations in the BOLD response that may
ygenated. Because active brain regions have an increased represent aspects of the individual’s internal state (mood,
rate of neural firing, they recruit more blood flow, which is future planning, daydreaming, etc.).
detected by the BOLD response. Resting-state FC research has revealed a number of net-
Functional MRI has the advantage over positron emis- works with patterns of synchronous activity that are con-
sion tomography (PET) and single photon emission com- sistently found in healthy individuals. The default mode
puted tomography (SPECT) of not requiring the use of network (DMN) is a network that is active when an indi-
ionizing radiation. Functional MRI also provides better vidual is awake and at rest. The DMN is an anatomically
spatial resolution than PET and SPECT. An additional defined brain system that preferentially activates when
advantage is that fMRI requires little special equipment one’s focus is on internal experiences such as daydreaming,
beyond a standard MRI console as used to perform struc- planning, or memory retrieval. It is negatively correlated
tural scans. Disadvantages of fMRI include a slow temporal with brain systems responsible for external visual or audi-
resolution, on the order of seconds, and a potential loss of tory signal processing and is one of the most studied net-
signal along air/tissue interfaces such as the sinuses. Similar works of resting state brain activity, and one of the most
to sMRI, fMRI data can also be analyzed using either easily visualized. Depending on the method of resting state
ROI or voxel-wise (whole brain) approaches. However, analysis, FC researchers have reported a number of neural
brain network approaches to fMRI analysis (e.g., FC) have networks that appear to be functionally connected during
recently become especially popular due to their ability to rest. The key networks include, but are not limited to, the
identify functional relationships between brain structures. DMN, a sensory/motor component, an executive control
Other network-based approaches also exist, such as effec- component, numerous visual components, frontal/parietal
tive connectivity, which examines information flow and components, an auditory component, and a temporal/pari-
can be compared and contrasted with the FC approach. etal component.
However, for the illustrative rather than technical purposes
of this chapter we restrict the current exposition to a brief
overview of FC analysis and direct the reader to Eickhoff OV E RV I E W OF   s M R I F I N DI NG S
and Grefkes (2011) and Ramnani et al. (2004) for compre- I N   B I P OL A R DI S OR DE R
hensive reviews of FC and effective connectivity analyses
(Eickhoff & Grefkes, 2011; Ramnani, Behrens, Penny, & The neural substrates of bipolar disorder are incom-
Matthews, 2004). pletely defined. However, sMRI provides a detailed in vivo

S tructural and F unctional M agnetic R esonance I maging F indings   •   2 0 7


examination of neuroanatomy that is beginning to differen- 2008), posterior cingulate (Wang, Jackowski, et al., 2008),
tiate individuals with bipolar disorder from other psychiatric and subcortical white matter (Haznedar et  al., 2005). By
groups and from healthy individuals. Structural MRI studies contrast, mean diffusivity changes have not been widely
have been useful not only to define the neural substrates of identified outside of frontosubcortical white matter tracts
bipolar disorder but also in guiding fMRI and neurochemi- (Beyer et al., 2005; Houenou et al., 2007; Yurgelun-Todd
cal studies using magnetic resonance spectroscopy (discussed et al., 2007). In general, the greater reliability of frontosub-
in Chapter 18). cortical DTI findings, and the neurodevelopmental impli-
cation that frontosubcortical white matter abnormalities
underlie a behavioral “disconnection syndrome” involv-
W H I T E M AT T E R A BNOR M A L I T I E S ing disinhibition and risk-taking behaviors, has driven the
White matter hyperintensities focus on these circuits in bipolar disorder.
Some investigators have reported reduced FA in ante-
One of the most replicated anatomic abnormalities in rior prefrontal white matter tracts, suggesting neuropathic
bipolar disorder is a greater incidence of T2-signal hyper- changes (Adler et al., 2004; Wang, Jackowski, et al., 2008),
intensities in white matter tracts relative to healthy and while others have reported either a lack of FA differences
psychiatric comparison samples (Altshuler et  al., 1995). in prefrontal white matter (Beyer et al., 2005; Bruno et al.,
These small lesions appear on T2-weighted MRI scans as 2008; Haznedar et al., 2005; Yurgelun-Todd et al., 2007) or
areas of higher signal intensity relative to surrounding tis- increased FA (Versace et al., 2008). The areas involved, how-
sue. It should be noted that white matter hyperintensities ever, do not entirely overlap, and it is still unclear to what
are likely only secondarily related to bipolar disorder, with extent discrepant findings between studies reflect meth-
primary relations to cardiovascular risk factors, which are odological differences. Similarly, frontosubcortical diffu-
more common in individuals with bipolar disorder and due sivity has been found to be both increased (the direction
to advancing age and to a positive treatment history. The of pathology) or unchanged in adult patients with bipolar
etiology of these hyperintense lesions, and their specificity disorder (Adler et al., 2004; Beyer et al., 2005; Bruno et al.,
to bipolar disorder, is an issue that has yet to be resolved. 2008; Yurgelun-Todd et al., 2007). Nonetheless, a pattern
It has been speculated that they adversely affect signaling of structural frontosubcortical network disruption is the
between frontosubcortical and limbic structures, thereby most consistent finding in adults with bipolar disorder.
influencing mood and cognition (Norris, Krishnan, & Only a few DTI studies have been conducted examining
Ahearn, 1997; Strakowski, DelBello, Adler, Cecil, & Sax, children and adolescents with bipolar disorder, and find-
2000). This possibility is consistent with DTI studies ings here have also been inconsistent. Adler et  al. (2006)
implicating abnormalities in white matter tract composi- studied a small group of bipolar and healthy youth ranging
tion in bipolar disorder. from 10 to 18 years of age. Individuals with bipolar disorder
were experiencing their first manic or mixed episode when
they were scanned. Findings mirrored the results obtained
Diffusion tensor imaging
in older patients; FA was reduced in anterior white matter
Diffusion tensor imaging has only been used to examine tracts while diffusivity was unchanged. In an additional
white matter tracts in individuals with bipolar disorder analysis of posterior white matter tracts, no between-group
for approximately a decade, and findings have been mixed. differences in either FA or mean diffusivity were found
For the most part, bipolar disorder research has been con- (Adler et al., 2006). Frazier et al. (2007) examined a some-
cerned with white matter tracts within and connecting to what younger group of bipolar children. Decreased FA
frontal cortex. Elsewhere in the brain, DTI findings have was again observed in superior-frontal white matter tracts
been either inconsistent or lack replication. One excep- in addition to orbitofrontal white matter and corpus cal-
tion is the corpus callosum, in which Wang, Kalmar, et al. losum. Moreover, FA values did not correlate with age.
(2008) identified reduced FA in the anterior corpus callo- They extended their study to include a group of high-risk
sum in adults with bipolar disorder. This finding is consis- children that showed similar but smaller differences in FA
tent with another report of reduced corpus callosum FA in compared with healthy individuals (Frazier et  al., 2007).
adult patients, possibly indicating altered interhemispheric The combination of reduced FA values in acute and at-risk
connectivity (Yurgelun-Todd, Silveri, Gruber, Rohan, & children independent of age suggests that these findings are
Pimentel, 2007). Additionally, reduced FA has been risk markers for bipolar disorder rather than developmental
reported in temporal cortex (Bruno, Cercignani, & Ron, in nature. The combination of reduced FA without changes

2 0 8   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
in mean diffusivity suggests that frontal white matter suggest that brain abnormalities in bipolar disorder are pri-
abnormalities in youth with bipolar disorder may represent marily regional, as global measures of total cerebral volumes,
axonal disorganization rather than frank axonal loss (Adler and gray and white matter volumes are frequently unaffected
et  al., 2006). However, these findings are not pathogno- (Haldane & Frangou, 2004). This contention is supported
monic for bipolar disorder. Changes in anisotropy and by meta-analyses (Hallahan et al., 2011; Hoge, Friedman, &
diffusivity have also been observed in patients with other Schulz, 1999; McDonald et al., 2004). Consistent with
affective and psychotic disorders, including major depres- findings in major depressive disorder, Hoge et al. concluded
sive disorder and schizophrenia (Kyriakopoulos, Bargiotas, that there are no cerebral volume differences in bipolar dis-
Barker, & Frangou, 2008). Many of these changes suggest order based on a meta-analysis. In a larger meta-analysis,
similar pathology in frontosubcortical tracts consistent McDonald et al. also found no difference in total brain
with elements of overlapping symptomatology observed volume. However, despite a lack of significant group dif-
among bipolar disorder, major depressive disorder, and ferences in this regard, Hallahan et al. found that cerebral
schizophrenia. volume reduction was associated with longer illness dura-
tions in individuals with bipolar disorder. In light of often-
reported null mean effects for global cerebral volume, many
V E N T R ICU L A R S Y S T E M CH A NG E S
investigators have chosen to focus on volume changes in
Despite the fact that DTI, and white matter hyperinten- specific brain regions related to mood and cognitive regu-
sity counts and confluence ratings, indicate white matter lation, primarily structures within frontosubcortical net-
pathology in bipolar disorder, brain tissue segmentation works and their intersection with medial temporal circuits.
(into white matter, gray matter, and cerebral spinal fluid)
studies often demonstrate little loss of white matter volume
Frontal cortex
(e.g., Haldane & Frangou, 2004). This conclusion stands
in contrast to tissue segmentation and postmortem stud- Although abnormalities in cerebral brain regions are
ies involving the ventricular system in which abnormalities reported inconsistently in bipolar disorder, when present
are readily apparent. Indeed, one of the earliest and most they typically consist of decreased frontal or prefrontal cor-
replicated findings in the psychiatric literature in general is tical volumes (Strakowski, Adler, & DelBello, 2002). Such
ventricular enlargement in schizophrenia (e.g., see Shenton, findings are consistent with histological reports of glial and
Dickey, Frumin, & McCarley, 2001, for a review). Although neuronal cell loss in prefrontal and anterior cingulate cor-
enlargement of the lateral and third ventricles is less con- tex in patients with bipolar disorder. Decreased volumes
sistently observed in bipolar disorder (Norris et al., 1997), have been observed particularly in cortical regions encom-
ventricular enlargement in bipolar disorder, when present, passing ventrolateral and dorsolateral prefrontal cortex
may represent underdevelopment or degeneration of peri- (Harvey, Persaud, Ron, Baker, & Murray, 1994; Strakowski
ventricular structures rather than a ventricular abnormal- et  al., 1993), which help regulate emotion and cognition.
ity per se (Strakowski, DelBello, et  al., 2002; Strakowski For instance, Sax et al. (1999) noted that decreased prefron-
et al., 1993). This suggestion is supported by a large litera- tal volume correlated with decrements in sustained atten-
ture reporting both periventricular and other regional gray tion on a continuous performance task. Drevets et al. (1997)
matter abnormalities in bipolar disorder. specifically examined the subgenual prefrontal cortex and
noted that not only was the volume decreased in individu-
als with bipolar disorder but the decrease was also associ-
R E G IONA L G R AY M AT T E R
ated with decreased metabolic activity as well (Drevets
A BNOR M A L I T I E S
et  al., 1997). However, McDonald et  al. (2004), who
Advances in MRI resolution have made research into examined prefrontal and subgenual prefrontal cortex in a
regional brain volumes common and increasingly accu- meta-analysis, found no volume differences in these brain
rate. ROI analysis and VBM are the state-of-the-art image regions, thereby highlighting the inconsistency of sMRI
analysis techniques most often used today, although newer findings in frontal cortex. Measurements of cortical thick-
approaches such as cortical thickness analysis and deforma- ness, on the other hand, have consistently demonstrated
tion based morphometry are continually being developed. thinning of prefrontal cortex in bipolar I  (Foland-Ross
Consistent with findings in major depressive disorder et al., 2011; Lyoo et al., 2006; Rimol et al., 2010) and bipo-
(Coffey et al., 1993; Husain et al., 1991; Krishnan et al., lar II disorder (Elvsashagen et  al., 2013); similar findings
1992; Kumar, Bilker, Jin, & Udupa, 2000), sMRI studies have been reported in other cortical regions including

S tructural and F unctional M agnetic R esonance I maging F indings   •   2 0 9


temporal (Elvsashagen et al., 2013; Rimol et al., 2010) and Strakowski, 2008), so the use of reduced amygdala size as
anterior cingulate cortex (Foland-Ross et  al., 2011; Lyoo a risk biomarker of incipient mania prior to illness onset is
et  al., 2006). Unlike cortical thinning, however, cortical still tenuous. However, despite a lack of mean group volu-
surface area may not be affected in bipolar I or II disorders metric differences, Chen et al. observed that amygdala size
(Elvsashagen et al., 2013; Rimol et al., 2012). was positively correlated with age in bipolar adolescents
but negatively correlated with age in healthy adolescents,
suggesting abnormal amygdala development in at-risk indi-
Medial temporal structure: Amygdala
viduals. This result provides an indication that the devel-
In addition to frontal lobe volume deficits, medial tempo- opmental trajectory of the amygdala is different in bipolar
ral lobe structures may also be abnormal in bipolar disorder disorder than healthy individuals, consistent with a recent
(Schlaepfer et al., 1994). Many prior research studies identi- finding that adolescents experiencing a first manic episode
fied volume changes in the amygdala (Altshuler et al., 2000; do not show increases in amygdala volume over 12 months
Altshuler, Bartzokis, Grieder, Curran, & Mintz, 1998; seen in healthy adolescents and adolescents with atten-
Pearlson et al., 1997; Strakowski et al., 1999). Strakowski tion deficit hyperactivity disorder (Bitter, Mills, Adler,
et al. found bilaterally increased amygdala volume in bipo- Strakowski, & DelBello, 2011). Indeed, reduced amygdala
lar relative to health individuals, and Altshuler et al. found volumes may represent a disease-specific finding, as this
increased amygdala volumes in individuals with bipolar effect has not been associated with other childhood-onset
disorder relative to both schizophrenic and healthy indi- disorders such as attention deficit hyperactivity disorder
viduals. Pearlson et al., however, reported that individuals and autism (Filipek et al., 1997; Schumann et al., 2004).
with bipolar disorder had decreased left amygdala volumes
relative to healthy individuals and no volume differences
Medial temporal structure: Hippocampus
relative to individuals with schizophrenia.
Indeed, one of the most important neurodevelop- Unlike amygdala changes, and contrary to findings in major
mental distinctions in the sMRI literature to date has depression, there appears to be little change in hippocam-
been the finding of reduced amygdala volume in bipolar pal volume in bipolar disorder. Videbech and Ravnkilde
youth relative to normal or enlarged amygdala volume in (2004) combined data from six studies reporting hippo-
adults with bipolar disorder (Fleck et  al., 2008; Pfeifer, campal volumes in bipolar relative to healthy individuals,
Welge, Strakowski, Adler, & DelBello, 2008). In stud- only one of which showed changes in bipolar disorder. No
ies involving pediatric bipolar samples, reduced amygdala differences in hippocampal volume were identified in the
volume has been the most consistent neuroanatomical resulting meta-analysis either, consistent with a separate
finding (Blumberg, Kaufman, et  al., 2003; Chen et  al., meta-analysis by McDonald et al. (2004) conducted in the
2004; DelBello, Zimmerman, Mills, Getz, & Strakowski, same year. Based on such findings, Strakowski, Adler, et al.
2004; Dickstein et al., 2005; Frazier et al., 2005). In con- (2002) suggested that affective illness may involve under-
trast, studies of individuals with adult-onset bipolar dis- developed or atrophied prefrontal regions that lead to a
order have found either increased (Altshuler et  al., 2000; loss of cortical modulation of emotional circuits involving
Brambilla et  al., 2003; Frangou, Donaldson, Hadjulis, the amygdala. Possible reasons for a preferential impact on
Landau, & Goldstein, 2005; Sax et  al., 1999; Strakowski amygdala volume, to a greater degree than hippocampal vol-
et al., 1999) or normal (Strakowski, DelBello, et al., 2002; ume, are unclear considering the functional and anatomi-
Swayze, Andreasen, Alliger, Yuh, & Ehrhardt, 1992) amyg- cal proximity of these structures and the fact that bipolar
dala volumes relative to healthy individuals. Smaller amyg- disorder is characterized by declarative memory deficits in
dala volumes have even been reported in first-episode addition to mood dysregulation.
bipolar samples (Rosso et al., 2007), suggesting that amyg- Bearden et  al., (2008), demonstrated that total hip-
dala volume deficits are apparent early in the illness course pocampal volume was significantly smaller in adolescents
and might be useful as an illness biomarker. Smaller amyg- with bipolar disorder relative to healthy individuals.
dala volumes have also been noted in bipolar youth who Using a mesh mapping approach, the authors were able
have a parent with bipolar disorder (Chang et al., 2005), but to identify localized deformations in the head and tail
studies of children at familial risk for developing bipolar of the left hippocampus in adolescents with bipolar dis-
disorder do not always demonstrate a significant decrease order relative to their healthy counterparts, suggesting
in amygdala volume (Blumberg, Kaufman, et  al., 2003; that by probing the microstructure of common ROIs new
Chen et  al., 2004; Singh, Delbello, Adler, Stanford, & insights regarding the impact of bipolar disorder on brain

210   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
structure are possible. Additionally, in the same study, numerous factors, most significantly the episodic nature
there was a significant positive correlation between hippo- of the illness and the effects of medications. In adults,
campal size and age in individuals with bipolar disorder, the majority of functional studies have used euthymic
whereas healthy adolescents showed an inverse relation; patients, and only a smaller body of research explores
associations similar to those previously demonstrated for functional activation patterns during episodes of depres-
the amygdala in bipolar and healthy adolescents, respec- sion or mania, despite the fact that mania is the defining
tively (Chen et al., 2004). feature of the illness. In addition, the majority of func-
tional imaging studies are confounded by current or prior
psychotropic medication exposure among participants,
Basal ganglia and thalamus with only a small minority of studies assessing medica-
Several studies reported morphological differences between tion naïve individuals.
bipolar and healthy individuals in the basal ganglia (Aylward
et al., 1994; Strakowski et al., 1999), as well as the thalamus E MOT ION A L PRO C E S S I NG S T U DI E S
(McIntosh et al., 2004; Strakowski et al., 1999). Consistent
with the field as a whole, however, sMRI studies of the basal Emotional processing studies in bipolar disorder have
ganglia and thalamus have been mixed in bipolar disorder, employed a variety of behavioral paradigms, including both
with most showing no differences compared with healthy implicit and explicit emotional processing and regulation.
individuals (Dupont, Jernigan, et al., 1995; Harvey et al., The brain regions most commonly studied in emotional
1994; Sax et al., 1999; Strakowski et al., 1993; Swayze et al., processing and regulation paradigms overlap significantly
1992). However, Strakowski et al. (1999) found bilaterally with those that have been the focus of sMRI studies in
increased striatal volumes in individuals with bipolar dis- bipolar individuals and include the amygdala, the ante-
order and Aylward et al. (1994) also found increased basal rior cingulate cortex, and the medial and ventrolateral
ganglia volumes but in males with bipolar disorder only. prefrontal cortices. In healthy individuals, these regions
Similarly, although certain investigators report no abnor- interact as a network that functions to identify and process
malities in thalamic volume in bipolar disorder (Caetano emotional stimuli and to regulate emotional responses. In
et al., 2001; Dupont, Jernigan, et al., 1995; Sax et al., 1999; this model, the amygdala functions to detect and signal
Strakowski et al., 1993), Strakowski et al. and Dupont, the salience of potentially emotional stimuli. Amygdala
Butters, et al. (1995) found increased thalamic volumes. activation in response to such stimuli triggers physiologi-
cal and behavioral responses via connections between this
region and both the autonomic nervous system and the
OV E RV I E W OF  f M R I F I N DI NG S hypothalamic-pituitary-adrenal axis. The prefrontal regions
I N   B I P OL A R DI S OR DE R in this network are thought to play a regulatory or inhibi-
tory role through reciprocal connections with amygdala
Bipolar disorder has long been considered primarily a dis- and other limbic regions. For example, the medial prefron-
order of mood regulation. While abnormal emotional pro- tal cortex and the ventrolateral prefrontal cortex may act to
cessing is a core feature of the illness, neuropsychological dampen response to threatening stimuli both by reducing
research has demonstrated that individuals with bipolar the activation of amygdala and via inhibitory connections
disorder also demonstrate deficits in a number of cogni- between the ventrolateral prefrontal cortex and the auto-
tive domains, including attention, executive function, nomic nervous system. In tasks involving healthy partici-
and verbal memory (Fleck et al., 2009). These deficits are pants, successful emotional regulation is often associated
reviewed in detail in Chapter 16. Functional MRI stud- with increased activation of ventrolateral and medial pre-
ies in bipolar disorder can be largely divided along these frontal cortices and reductions in amygdala activation. The
emotional-cognitive lines. The majority of fMRI studies anterior cingulate cortex is thought to function in healthy
of the bipolar brain have attempted to detect functional individuals to manage attentional processing, in part by
abnormalities associated with emotional processing and modulating attentional commitment to emotional stimuli.
regulation. However, there is also a significant body of The anterior cingulate may also be involved in regulation
research that has sought to assess the functional altera- of reciprocal interactions between emotional and cognitive
tions associated with purely cognitive tasks and executive processing systems (Yamasaki, LaBar, & McCarthy, 2002).
function in bipolar disorder. Similar to sMRI studies, For more through reviews of healthy human circuitry for
fMRI research in bipolar disorder is complicated by

S tructural and F unctional M agnetic R esonance I maging F indings   •   211


emotional perception and regulation see Ochsner and emotional regulation, hypoactivation in this region may
Gross (2005) or Phillips, Drevets, Rauch, and Lane (2003). contribute to the increased emotional reactivity seen in
As might be expected, given the prominence of emo- bipolar patients, even those not currently experiencing an
tional dysregulation in individuals with bipolar disorder, acute mood episode. Further research in this area is needed,
studies of emotional processing in this population have as very few fMRI studies have directly compared the func-
often found altered patterns of functional activation in this tional correlates of emotional processing in bipolar individ-
emotional processing network. The abnormalities detected uals in different mood states, and even fewer have done so
align with the hypothesis that bipolar disorder is associ- by following the same patients longitudinally.
ated with heightened emotional reactivity and decreased
emotional regulation. The most frequently replicated find-
CO G N I T I V E PRO C E S S I NG S T U DI E S
ings in both youth and adults are hyperactivation of the
amygdala and hypoactivation in ventrolateral prefrontal While emotional processing alterations are core features
cortex, although the striatum is also frequently involved of bipolar disorder, these individuals also demonstrate
(Townsend & Altshuler, 2012). A  recent meta-analysis significant deficits in purely cognitive tasks, particularly
found that during emotional processing tasks, individuals those associated with domains of executive function.
with bipolar disorder show an overall pattern of increased Understanding the functional underpinnings of these defi-
activation in the medial temporal lobe, including the amyg- cits is essential because executive function deficits are asso-
dala, hippocampus, and parahippocampal gyrus, and in ciated with significant functional impairment and poor
putamen, along with decreased activation in the inferior outcomes in both youth and adults with bipolar disorder
frontal gyrus (Chen, Suckling, Lennox, Ooi, & Bullmore, (Biederman et al., 2011; Pavuluri et al., 2006). Studies of
2011). Interestingly, in bipolar youth hyperactivity of the functional correlates of cognitive tasks in individuals with
amygdala in response to emotional stimuli has been found bipolar disorder have employed a variety of behavioral
to correlate with decreases in volume (Kalmar et al., 2009). paradigms, including sustained attention, response inhibi-
The degree of hyperactivity of amygdala during emotional tion, interference control, and working-memory tasks. This
tasks has also been found to correlate with symptom severity diversity of paradigms generally means that relatively few
as measured by the Young Mania Rating Scale (Bermpohl studies have been conducted for each executive function
et al., 2009). domain. However, there are some commonalities across
A comprehensive understanding of emotional process- studies of cognitive processing in bipolar disorder.
ing in bipolar disorder is complicated by the multiple mood In the past, cognitive and emotional processing systems
states that bipolar individuals experience over the course were thought of as separate, both anatomically and func-
of their illness. While the majority of imaging studies in tionally. In such models, cognitive processing was thought
adults have included only subjects that were euthymic at to be “top-down,” primarily involving dorsal regions of
the time of the scan, several studies have included indi- the prefrontal cortex, while emotional systems were more
viduals during manic or depressive episodes. When these “bottom-up,” relying on limbic system and ventral pre-
studies are considered together, there is evidence that the frontal regions. Such hypotheses separating emotion from
functional alterations associated with emotional processing cognition have largely fallen out of favor, as recent find-
in bipolar disorder have both state- and trait-related com- ings suggest that the classically “cognitive” brain regions,
ponents. Specifically, amygdala hyperactivation during the including the anterior cingulate and dorsolateral prefrontal
processing of emotional stimuli is seen in the majority of cortex, and classically “emotional” regions interact signifi-
studies performed with participants in the manic state and cantly across a variety of tasks. Some of these interactions
is less frequently replicated in depressed or euthymic sam- reflect the idea of reciprocal interactions between emotional
ples (Townsend & Altshuler, 2012). Therefore, amygdala and cognitive processing. That is to say, “purely cognitive”
hyperactivity, at least in response to emotional stimuli, may processing involves suppression of emotional networks,
be a state-related finding exclusive to mania. In contrast, whereas strong emotional activation interferes with cogni-
hypoactivation in the ventrolateral prefrontal cortex has tive processing (Yamasaki et al., 2002). For a review of the
been detected in the majority of studies of emotional pro- literature on the interactions between cognitive and emo-
cessing in bipolar disorder, regardless of mood state, sug- tional processing systems in healthy individuals, see Pessoa
gesting that altered function in that region may be a trait (2008).
abnormality in bipolar disorder (Townsend & Altshuler, It is not surprising, then, that the alterations seen
2012). Given the ventrolateral prefrontal cortex’s role in in fMRI studies of cognitive processing paradigms in

21 2   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
individuals with bipolar disorder often involve the same to clarify relationships between genes associated with bipo-
regions found to be dysfunctional in tasks of emotional lar disorder and functional alterations.
processing. In a meta-analysis, the performance of purely
cognitive tasks by individuals with bipolar disorder across
multiple studies was associated with decreased activation M E DIC AT ION E F F E C T S
in the inferior frontal gyrus, the lingual gyrus, and the I N   I M AG I NG S T U DI E S
putamen (Chen et  al., 2011). Dysregulation in regions of OF   B I P OL A R DI S OR DE R
the ventrolateral prefrontal cortex has been detected dur-
ing sustained attention, response inhibition, and interfer- Bipolar disorder is a serious psychiatric illness that is most
ence control in individuals with bipolar disorder (Altshuler often treated with psychotropic medications, including
et al., 2005; Blumberg, Leung, et al., 2003; Schneider et al., atypical antipsychotics, lithium, and other mood stabiliz-
2012). Indeed, a recent meta-analysis of functional imaging ers. As ethical limitations prevent delaying or terminating
studies in bipolar disorder reported decreased activation effective treatment for the purpose of research, the major-
in the inferior frontal gyrus during cognitive processing ity of MRI studies in bipolar patients are confounded by
tasks (Chen et al., 2011). Similarly, hyperactivation of the the potential effects of medications on brain structure
amygdala has been detected in cognitive paradigms, even and function. Very limited research has attempted to
in the absence of emotionally valenced stimuli (Fleck et al., directly assess the effects of medication on the brains of
2012). However, limbic hyperactivation is less reliably bipolar individuals. Rather, the majority of studies that
detected in purely cognitive tasks than in tasks using emo- explore the effects of medications on imaging findings do
tionally valenced stimuli (Chen et  al., 2011). In addition, so as a secondary analysis, for the most part in an attempt
striatal, parietal, and cingulate abnormalities appear to be to identify the influence of medication exposure as a con-
more prominent during tasks requiring cognitive process- found on the interpretation of results. The majority of
ing than during emotional processing, and dorsal prefron- studies that have taken this approach have found either no
tal alterations have been detected, most prominently in significant effects of medication on either brain structure
studies employing memory paradigms (Kupferschmidt & or function in bipolar individuals or have found medica-
Zakzanis, 2011). Interestingly, alterations in dorsal prefron- tion exposure to have a normalizing effect. This pattern
tal regions include both hypoactivity and hyperactivity. It is suggests that medication exposure confounds may be
possible that, in some situations, processing is impaired by increasing risk of type II error but not introducing spuri-
insufficient recruitment of dorsal prefrontal regions, while ous findings not associated with the pathophysiology of
in others the dorsolateral prefrontal cortex is dispropor- the disease.
tionally recruited, perhaps as a compensatory control.
LITHIUM
T H E ROL E OF G E N E T IC S
The medication most frequently associated with signifi-
Recently, several studies have explored relationships among cant volumetric changes is lithium; therefore, the effect
single nucleotide polymorphisms in genes associated with of lithium on brain volume has received considerable
bipolar disorder and MRI findings. While genetic asso- attention relative to other mood stabilizers. An early
ciations with structural alterations have been rare, several study demonstrated that increases in T1-weighted signal
studies in both healthy and bipolar samples have found values in frontal white matter normalized with lithium
an association between the risk alleles at several gene loci, treatment (Rangel-Guerra et al., 1983). More recently,
including CACNA1C, ANK3, and ODZ4, and activation Brambilla, et al., (2001) found no effect of lithium on
patterns associated with bipolar disorder, especially limbic total brain volume in individuals with bipolar disorder,
hyperactivity during emotional and reward tasks (Dima but most extant studies suggest an effect. For instance,
et al., 2013; Heinrich et al., 2013; Jogia et al., 2011; Tesli Moore et al. (2000) found increased total brain volume
et al., 2013). However, several of these studies also suggest in bipolar disorder after four weeks of lithium treatment.
a complicated picture, with significant gene-by-diagnosis Drevets (2000) found that individuals taking either lith-
interactions, such that individuals with bipolar disorder ium or valproic acid do not exhibit the same reductions
show an opposite pattern of activation change associated in subgenual prefrontal cortex volumes seen in those not
with risk allele load than healthy subjects (Dima et  al., taking these medications.
2013; Jogia et al., 2011). More research in this area is needed

S tructural and F unctional M agnetic R esonance I maging F indings   •   213


The preponderance of evidence points to lithium’s lithium. As can be seen in Figure 17.1, mesh mapping results
influence on the amygdala-hippocampal complex in par- from this study revealed localized increases in regions of
ticular, in both pediatric and adult bipolar samples. It was the hippocampus in lithium-treated individuals with bipo-
reported that amygdala size is negatively associated with lar disorder compared with healthy individuals. Observed
antidepressant exposure and duration of illness (DelBello reductions in subregions of the hippocampus in unmedi-
et al., 2004), but positively associated with lithium or dival- cated bipolar disorder were proposed as a possible neural
proex exposure in pediatric samples (Chang et  al., 2005), correlate of the declarative memory deficits reported in the
possibly indicating that antidepressants (and increased ill- illness. (Bearden et al., 2008). Beyond the possible impact
ness duration) potentate the risk of amygdala reductions, of these results on treatment models of bipolar disorder,
while lithium and mood stabilizers protect against volume they also highlight the ability of sMRI to probe the micro-
loss in pediatric bipolar disorder. structure of even small brain volumes. Finally, in a recent
Foland et al. (2008) assessed whether lithium treatment meta-analysis of 321 individuals with bipolar I  disorder
was associated with volume differences in the amygdala and 442 healthy individuals, Hallahan et al. (2011) demon-
and hippocampus in adults with bipolar disorder using strated that bipolar disorder is characterized by increased
tensor-based morphometry, which derives volumes from right lateral ventricular, left temporal lobe, and right puta-
the amount of warping required to match structures to a men volumes. More important, individuals treated with
common template. Group comparisons of deformation lithium displayed significantly increased hippocampal and
maps showed that lithium-treated individuals exhibited amygdala volumes compared with untreated and healthy
significantly increased volumes of amygdala and hippo- individuals. Therefore lithium exposure appears to be a
campus compared with individuals not receiving lithium. source of considerable heterogeneity in sMRI results and
Bearden et al., (2008) also reported that total hippocampal to have a possible neuroprotective effect, as suggested by its
volumes were significantly larger in lithium-treated indi- association with increased amygdala-hippocampal complex
viduals with bipolar disorder compared with healthy and volumes (Hallahan et al., 2011).
unmedicated individuals with bipolar disorder, indicat- A small prospective study by Selek and colleagues
ing that the greater volume reflects a neurotropic effect of (2013) directly assessed the impact of lithium treatment

Li+ BPD vs. Healthy Control: Increase in Li+ Group

(a)

Percent Difference

–30% –20% –10% +5%

(b)

Significance

<0.01 0.02 0.03 0.04 >0.05

Figure 17.1 
Statistical 3-D maps showing local differences in hippocampal structures between lithium-treated participants with bipolar disorder
and healthy comparison participants in terms of percent difference (a) and statistical significance (b). The right hippocampus is on the left side of
the figure (from Bearden et al., 2008).

214   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
on brain volumes in bipolar individuals longitudinally PRO S PEC T I V E M E DIC AT ION
over a four-week treatment period. This study used auto- S T U DI E S
mated ROI techniques to test treatment effects on vol-
ume of the hippocampus, amygdala, anterior cingulate, A much smaller body of research has prospectively assessed
and dorsolateral prefrontal cortex, as well as prefrontal the effects of medication directly through repeated struc-
cortex overall. The results suggested that the neurotropic tural or functional scanning of individuals before and after
effects of lithium on brain volumes may be region-specific specific medication regimens. The majority of these studies
and related to the level of treatment response. Specifically, were conducted in youth with bipolar disorder. Some sig-
individuals who responded to lithium treatment showed nificant limitations are common to much of the published
increased volumes with treatment in left dorsolateral pre- research in this area. First, these studies have almost uni-
frontal cortex and prefrontal cortex overall that was not versally employed treatment courses of 14 weeks or less and
seen in those who did not respond to treatment. Those in some cases as few as 4 weeks. Second, due in part to the
who did not respond to lithium showed volume decreases ethical issues associated with withholding treatment from
in the left hippocampus and the right anterior cingulate. acutely ill individuals for research purposes, the majority of
Healthy controls not treated with medications and bipo- studies do not include a placebo group and instead employ
lar individuals who were euthymic at baseline showed no only healthy youth as comparison subjects. To date, only
volumetric changes (Selek et  al., 2013). Interpretation of one published report has included a placebo-treated group
these results, particularly for the nonresponder group, of youth with bipolar disorder followed longitudinally to
is limited by the very small sample size, as only six indi- control for activation changes associated with untreated
viduals failed to respond. However, complex interactions disease course (Schneider et  al., 2012). Third, these stud-
between treatment response and volumetric effects may ies employed a variety of different cognitive tasks, which
explain some of the heterogeneity in correlational analyses. complicates interpretation of results across studies. Finally,
Interestingly, this study also found that those who did not the majority of studies included only youth in manic,
respond to lithium had smaller right amygdala volumes at hypomanic, or mixed states. Only two prospective stud-
baseline, suggesting that regional volumes may also be use- ies explored treatment effects in the depressed phase of ill-
ful in predicting response prior to the onset of treatment. ness, one using lamotrigine monotherapy (Chang, Wagner,
Despite the fact that lithium effects on sMRI findings Garrett, Howe, & Reiss, 2008) and another using a natu-
have frequently been explored, fMRI effects have been ralistic treatment protocol (Diler et al., 2013). Despite these
explored only as secondary analyses attempting to con- limitations, published research in this area has significant
trol for medication effects. Longitudinal studies exploring potential to inform our understanding of the neurofunc-
the functional effects of lithium treatment are needed to tional correlates of treatment and response in individuals
determine if the structural changes seen in patients treated with bipolar disorder.
with lithium are associated with functional normalization Two published studies prospectively looked at treat-
in the same regions. ment effects in adults with bipolar disorder. Both employed
a 12-week course of lamotrigine in adults who were stable at
baseline. In an emotional processing task (sad affect recog-
A N T I P S YCHOT IC S
nition), lamotrigine treatment was associated with increases
Even short-term treatment with conventional antipsychot- in activation, including both dorsomedial and ventrolat-
ics has been associated with increased basal ganglia vol- eral prefrontal cortex. These increases were interpreted
umes, especially caudate volume, and a variety of decreased as normalizing, as bipolar individuals showed decreased
cortical gray matter volumes. By contrast, atypical antipsy- activation relative to healthy subjects at baseline prior to
chotics seem to have little influence on the caudate, pos- lamotrigine treatment (Jogia, Haldane, Cobb, Kumari, &
sibly consistent with their relative lack of extrapyramidal Frangou, 2008). Another study employed both emotional
side effects, and have been related to increased cortical gray (angry affect recognition) and cognitive (n-back working
matter volume (Kumari & Cooke, 2006; Scherk & Falkai, memory) tasks. In this study, treatment with lamotrigine
2004), suggesting that atypicals may ameliorate structural was associated with increased activation in bilateral supe-
changes produced by psychopathology and/or typical anti- rior frontal and cingulate cortices during the n-back task
psychotic use. In fact, switching from typical to atypical and increased activation in the right middle frontal gyrus
antipsychotics may reduce basal ganglia volumes to values and left medial frontal cortex (Haldane et  al., 2008). As
comparable to healthy individuals (Scherk & Falkai, 2006). this study did not employ a comparison group of healthy

S tructural and F unctional M agnetic R esonance I maging F indings   •   215


participants, it is difficult to determine if these changes medication effects may differ between treatments. While
were normalizing. One significant limitation of these stud- symptoms improved for individuals in both treatment
ies is that they included only participants who were stable at groups, the two medications differed in their effects on
baseline, and participants showed little change in symptom activation during emotional working memory, affective
severity with treatment. Given that there are state-related color matching, and response inhibition (Pavuluri, Ellis,
differences in functional activation patterns in individuals Wegbreit, Passarotti, & Stevens, 2012; Pavuluri, Passarotti,
with bipolar disorder, these findings might not extend to Fitzgerald, Wegbreit, & Sweeney, 2012; Pavuluri,
individuals experiencing manic or depressive episodes. Passarotti, Lu, Carbray, & Sweeney, 2011). For example, in
Longitudinal studies of medication effects have been an emotional working-memory task, risperidone was asso-
more common in youth with bipolar disorder. In particular, ciated with increases in activation in the anterior cingulate
several medications have been studied longitudinally using and striatal regions, while divalproex was associated with
functional imaging paradigms in manic youth, includ- increased activation in the inferior frontal gyrus (Pavuluri,
ing lamotrigine, risperidone, divalproex, and ziprasidone. Passarotti, et al., 2012). In a response-inhibition task, the
Lamotrigine treatment has been associated with normal- medications were also found to have differential effects on
ization of prefrontal activation patterns in several cognitive FC patterns, with risperidone showing greater increases in
tasks, and this change in activation in several key regions FC of the insula in an affective circuit, while divalproex was
is correlated with treatment response. During a response associated with increases in subgenual cingulate connectiv-
inhibition task, youth with bipolar disorder showed ity (Pavuluri, Ellis, et al., 2012). The difference in findings
decreased activation in ventral prefrontal regions relative between treatment groups supports the hypothesis that the
to healthy subjects. Activation increased with lamotrigine functional effects were related to the actions of the medica-
treatment, and the increased activation in the ventrolat- tions themselves as opposed to treatment response, despite
eral prefrontal cortex was correlated with improvements in the absence of a placebo control groups.
manic symptoms (Pavuluri, Passarotti, Harral, & Sweeney, To date, the only placebo-controlled study of the neu-
2010). In an affective color-matching task, prior to treat- rofunctional correlates of pharmacological treatment in
ment bipolar individuals showed increased activation in manic youth with bipolar disorder employed ziprasidone
both dorsolateral and medial prefrontal regions relative to in a four-week, randomized, double-blind trial, which
healthy subjects. Following lamotrigine treatment, activ- employed a sustained attention task. Treatment with zipra-
ity in both of these regions normalized, and the increase sidone was associated with greater increases in activation in
in medial prefrontal cortex activity was correlated with a the right ventral prefrontal cortex. In addition, individu-
decrease in manic symptoms. Similar results of prefron- als with lower baseline activation in right BA 47 were more
tal normalization with lamotrigine treatment correlat- likely to respond to ziprasidone treatment, and baseline
ing with symptomatic improvement were obtained in an activation in this region was significantly correlated with
affective working-memory task. However, for the affective symptomatic improvement as measured by reduction in
working-memory task, significant amygdala hyperactiva- Young Mania Rating Scale scores (Schneider et al., 2012).
tion relative to healthy youth was detected at baseline, and Only a single, small study has explored the effects of
while there was some decrease in activation with treat- a specific medication on functional activation patterns in
ment, the amygdala remained hyperactive at the end of youth during the depressed phase of bipolar illness. Chang
the study period (Passarotti, Sweeney, & Pavuluri, 2011). and colleagues (2008) scanned eight depressed bipolar
Unfortunately, none of the studies of lamotrigine treat- youth during an emotional stimulus rating task before and
ment employed a placebo-treated control group. As these after eight weeks of lamotrigine therapy. Analysis focused
studies all had a very high response rate, it is difficult to on ROIs in the dorsolateral prefrontal cortex and amyg-
determine if the activation changes detected are associated dala bilaterally. All patients responded to treatment, and
with lamotrigine treatment specifically or with symptom- decreases in right amygdala activation between baseline and
atic improvement more broadly. endpoint were significantly correlated with symptomatic
The functional effects of risperidone and divalproex improvement, as measured by the Children’s Depression
on functional activity in manic youth with bipolar disor- Rating Scale, Revised (Chang et al., 2008). A more recent
der have been studied in six-week, double-blind, random- study looked at treatment effects more generally, scanning
ized studies. In these studies participants were randomized 10 depressed adolescents with bipolar disorder on a purely
to treatment with either risperidone or divalproex, and cognitive go/no-go task before and after six weeks of natu-
treatment groups were compared to assess directly how ralistic treatment, which included psychotherapy and either

216   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
medication changes or dose increases for most participants. in bipolar disorder. Further research is needed to better
When posttreatment scans were compared to baseline, assess the long-term effects of treatment with specific medi-
increased activation was seen in the right hippocampus cations on both structure and function of the brain in both
and left thalamus, and the decrease in depression scores pediatric and adult samples of bipolar disorder patients.
was correlated with the increase in left thalamus activity Interestingly, it appears that nonpharmacological inter-
in the six individuals who responded to the treatment regi- ventions for bipolar disorder may also have effects on func-
mens. Interestingly, the treatment effects in this study do tional activation patterns. In a small study, Ives-Delperi and
not appear to reflect normalization, as both of these regions colleagues (2013) scanned 16 bipolar individuals who were
of increased activation were hyperactive relative to healthy mildly symptomatic at enrollment before and after eight
subjects. Additionally, abnormally increased activation in weeks of mindfulness-based cognitive therapy (MBCT).
ventrolateral prefrontal cortex relative to healthy subjects A  comparison group of 7 bipolar individuals who were
persisted following treatment. The small sample and com- wait-listed for the mindfulness training and 10 healthy
plicated treatment protocol, including uncontrolled psy- subjects were scanned at the same intervals. Following the
chotherapy, make it difficult to interpret these findings in MBCT program, participants reported increased emotional
the context of specific medication effects. regulation and decreased anxiety and showed improved
The limitation of almost all prospective studies is that performance on neurocognitive testing. Relative to healthy
they assess the effects of monotherapy over a very brief subjects, bipolar individuals overall showed decreased acti-
timeframe—14 weeks or less—in a small number of sub- vation in the medial prefrontal cortex, a region associated
jects. In contrast, individuals with bipolar disorder often with emotional regulation. Treatment with MBCT was
remain on medication regimens for years, and many are associated with an increase in activation in this region, as
treated with multiple medications and combinations of well as the right posterior cingulate. Among those who
medications over the course of their illness. One recent underwent the treatment, the change in activation in the
study, by Yang and colleagues (2013), looked at the longer medial prefrontal cortex was correlated with self-reported
term effects of treatment in a small pediatric bipolar disor- increases in mindfulness (Ives-Deliperi, Howells, Stein,
der sample. In this study, unmedicated youth were scanned Meintjes, & Horn, 2013). The cognitive task used in this
at baseline during the performance of the pediatric affec- study was a block design mindfulness meditation task. It is
tive color-matching task. Individuals were then treated not clear if this functional normalization extends to emo-
following a standard treatment algorithm and underwent tional regulation in other contexts. However, mindfulness
follow-up scans at both 16 weeks and 3 years. A comparison strategies have been shown to be effective in attenuating
group of typically developing youth was scanned at the same emotional responsiveness, although the mechanism may
time points. The results demonstrate that youth with bipo- differ between beginning and very experienced meditators
lar disorder showed abnormal hyperactivation of right dor- (Taylor et al., 2011).
solateral prefrontal cortex, striatum, amygdala, and anterior
cingulate and that the activation in these regions normal-
ized over the three-year course of treatment. Interestingly, C ONC LUS ION
the right dorsolateral prefrontal cortex had significantly
normalized by 16 weeks, while the other regions took lon- Structural MRI has been instrumental in uncovering pat-
ger to do so, suggesting that cognitive circuitry may recover terns of brain abnormalities in bipolar disorder. Although
more quickly with symptom resolution than limbic areas. global volumetrics are often unaffected, it is now clear that
There are significant limitations to this study:  the large bipolar disorder is characterized by regional changes in both
span between the 16-week and three-year time points pre- white and gray matter. Increased numbers of white matter
vented pinpointing the time to recovery in later recover- hyperintensities have been observed (Aylward et al., 1994;
ing regions; moreover, the subjects did not all receive the Bearden, Hoffman, & Cannon, 2001; Norris et al., 1997),
same treatment but were treated using a flexible protocol and, although DTI data are still sparse, results suggest that
that allowed for multiple treatment options, medication white matter pathology in frontosubcortical tracts and the
changes, and medication combinations. Finally, the sample corpus callosum is present across bipolar age groups (Adler
size was small, with only 13 youth in the bipolar group and et al., 2004; Agarwal, Port, Bazzocchi, & Renshaw, 2010;
10 healthy subjects. Nonetheless, the study is significant in Wang, Kalmar, et  al., 2008). Furthermore, white matter
that it is the first to explore prospectively treatment effects abnormalities in these pathways may be intrinsic to bipolar
over the long courses of medication exposure typically used disorder and result in changes to the neurodevelopmental

S tructural and F unctional M agnetic R esonance I maging F indings   •   217


trajectory of networked structures (Bruno et  al., 2008; individual level, with varying accuracy (Grotegerd et  al.,
Frazier et al., 2007). 2013; Rocha-Rego et al., 2013; Schnack et al., 2013). The
The most consistent regionally defined structural practicality and clinical utility of such techniques remain
abnormalities identified in bipolar studies to date have been to be seen but offer promise.
in the prefrontal cortex (Drevets et al., 1997; Strakowski, Recent reviews of treatment effects indicate that psycho-
Adler, et  al., 2002), subgenual anterior cingulate cortex tropic medications have a primary influence on brain volu-
(Agarwal et al., 2010), basal ganglia (Aylward et al., 1994; metrics, possibly to a greater extent than DTI and fMRI, in
Strakowski et al., 1999), thalamus (McIntosh et al., 2004; both child and adolescent (Singh & Chang, 2012) and adult
Strakowski et  al., 1999), and especially the amygdala samples (Hafeman, Chang, Garrett, Sanders, & Phillips,
(Altshuler et  al., 2000; Pearlson et  al., 1997; Strakowski 2012). Studies suggest that lithium, in particular, is neuro-
et  al., 1999). As noted, although hypothesis-driven ROI protective and is associated with increased volumes in cor-
studies are often limited to frontosubcortical regions based ticolimbic regions subserving mood regulation (Hafeman
on the behavioral manifestation of bipolar disorder (call- et al., 2012), including the amygdala/hippocampal complex
ing into question the regional specificity of ROI results), (Hallahan et al., 2011). In general, psychotropic medications
brain-wide VBM results have helped confirm the impor- appear to either normalize or have no effect on sMRI indi-
tant role that these regions play in the neuropathology of ces, suggesting that they likely do not explain MRI findings
bipolar disorder. in bipolar disorder. Similarly, the effects of medication on
With respect to brain function, bipolar disorder is often functional activation appear to be primarily normalizing.
thought of as resulting from a disruption in emotional pro- However, different pharmacological treatments appear to
cessing systems. This hypothesis is supported by the frequent be associated with disparate functional effects, although the
replication of functional alterations in regions associated absence of a placebo control group complicates the interpre-
with emotional processing when participants with bipolar tation of most prospective medication studies using fMRI.
are exposed to emotionally valenced stimuli. Some altera- Since measures of symptomatic improvement often corre-
tions may be specific to mood state, for example, amygdala late with changes in activation patterns, these studies suggest
hyperactiviation appears to be most prominent during that there may be promise for functional imaging patterns as
mania, while other functional abnormalities, in particular predictors of treatment response; however, far more research
ventral prefrontal hypoactivation, may be associated with is needed in this area before this suggestion could yield clini-
bipolar disorder across all mood presentations (Townsend & cally actionable results. Finally, almost all of the prospec-
Altshuler, 2012). Abnormal interactions among these ventral tive studies that explore the effects of medication have been
prefrontal–limbic emotional networks and more dorsal cog- conducted in pediatric samples. It is likely that early-onset
nitive processing systems may be partially responsible for the bipolar disorder differs from adult-onset presentations, and
neurocognitive deficits associated with bipolar disorder. therefore those results might not extend to adults.
A growing body of research has combined structural Several factors need to be considered when attempting
and functional imaging techniques within the same study to differentiate patients with mood disorders on the basis of
(e.g., DTI and fMRI) to distinguish bipolar disorder from MRI. First, volume and activation changes may occur only as
other common psychiatric illnesses that have significant a cumulative effect in proportion to the number of affective
symptomatic overlap with bipolar disorder, including episodes (DelBello, Strakowski, Zimmerman, Hawkins,
depression and schizophrenia. For reviews of these direct & Sax, 1999; Mills, Delbello, Adler, & Strakowski, 2005).
comparison studies, see Whalley et al. (2012) and Cardaso Alternatively, these changes may reflect the cumulative
de Almeida and Phillips (2013). These studies found both effects of medications or illness severity may influence mea-
distinct and overlapping patterns of structural and func- surements. An understanding of these, and other, caveats
tional alterations associated with these illnesses. This with respect to the use of MRI in bipolar disorder will be
approach of combining structural and functional tech- essential to move the field forward. However, MRI findings
niques will likely prove essential to future progress in the may be useful to predict and monitor treatment response.
field. Other methodological and technical advances have In the future, the use of serial MRI to measure the rate
come from advances in computer science. Very recent of brain changes might provide a method for quantifying
studies have shown that computer analysis algorithms not only disease progression but also treatment efficacy in
that use imaging parameters may be useful in distinguish- bipolar disorder. Structural MRI studies are already being
ing individuals with bipolar disorder from healthy indi- used to predict treatment response in major depressive dis-
viduals and those with depression or schizophrenia on an order (Videbech & Ravnkilde, 2004). In one prospective

218   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
study, Chen et  al. (2007) correlated sMRI findings with This resolution provides superior boundary demarcation
treatment in depressed individuals before eight weeks of between gray and white matter, which will be essential for
fluoxetine treatment. Individuals with increased volume continued improvement in automatic VBM gray/white
of the anterior cingulate cortex, insula, and right temporo- segmentation algorithms. The use of multicoil receivers
parietal cortex showed faster rates of recovery from depres- can also improve resolution and/or reduce imaging time
sion (Chen et al., 2007). In bipolar disorder, Schneider and to allow visualization of more and more subtle structural
colleagues (2012) reported that youth with lower activity changes, reduce costs, and improve patient comfort. This
in right Brodmann area 47 at baseline were more likely to technology will allow a more fine-grained analysis of the
respond to ziprasidone treatment for mania. These findings microstructure of the brain using newer mesh mapping
support the extended use of MRI to help assign treatment techniques. On the macroscopic level, newer analysis tech-
for bipolar disorder in the future. niques, such as FC analysis, have provided insight into
On the technological front, the ongoing development network function and dysfunction in bipolar disorder, set-
and more routine use of higher field-strength magnets and ting the stage for examining connections among multiple
multicoil receivers will continue to advance the field. With regions using density mapping, an example of which is
all other variables being equal (e.g., gradient strength), depicted in Figure 17.2, structural equation modeling, and
high field magnets allow superior quality of neuroanatomi- graph theory approaches.
cal images with resolution in the submillimeter range by The ability of any single brain region to predict or
delivering images with improved contrast-to-noise ratio. track the development of bipolar disorder with diagnostic

A
A
C
C

B
B D
B

Figure 17.2 
These images show local and remote functional connectivity density (LFCD/RFCD) mapping results conducted as reported by Dardo
Tomasi and Nora D. Volkow [Functional connectivity density mapping, PNAS 2010; 107:9885-9890.] using the fMRI data from a single run of
an attentional task. Working successively through all voxels in the brain and beginning with the data from one subject, the number of functional
connections between a given voxel and all other voxels was computed. Using a voxel’s time series as the regressor of interest, we calculated the
correlation between that voxel and all other brain voxels. This image was then thresholded at r > 0.5 and the number significant voxels that were
direct neighbors of the given voxel or its significant neighbors, i.e, the number of “local connections,” were counted. This count was then placed
into a new dataset at the same location, and the whole process was repeated for each voxel. The resulting dataset contained the number of local
connections (significant correlations) for each voxel, a LFCD map. We also counted the number of remote connections, i.e., non-neighboring
significant “remote connections” and placed that count into an RFCD map. In total we included the data from 31 manic patients and 28 healthy
comparison subjects. The FCD maps were compared between groups using a t-test with a p-threshold of p < 0.01 and 25 or more voxels in a cluster
for LFCD (Top Row) and RFCD (bottom row). The regions that exhibited significantly greater LFCD in manic patients than healthy subjects
included (A) right medial superior frontal gyrus, BA 8, 6, and 9, and (C) Left Lingual Gyrus, BA 30. The regions that exhibited significantly greater
RFCD in manic patients that healthy subjects included (B) bilateral insula, BA 13, and (D) bilateral posterior cingulate cortex, BA 23.

S tructural and F unctional M agnetic R esonance I maging F indings   •   219


reliability remains fairly limited and has led to the sugges- Disclosure statement: Dr.  Strakowski serves as data-safety
tion that deterioration in interregional communication and monitoring board chair for Sunovion, Inc. and
processes (i.e., connectivity) may underlie the development Novartis, Inc. He is an EAP consultant to Procter &
of adolescent bipolar disorder (Rich et  al., 2008). Basic Gamble and has performed grant reviews for the National
seed-region FC methods have been used since the very ear- Institutes of Health. He has previously published with
liest days of functional MRI (Biswal, Yetkin, Haughton, & Oxford University Press.
Hyde, 1995) and can identify interregional low-frequency
BOLD correlations. As imaging research broadened to
include psychiatry, these methods have been implemented R E F E R E NC E S
to understand FC in mental illness. Seed-region FC suf-
Adler, C.  M., Adams, J., DelBello, M.  P., Holland, S.  K., Schmi-
fers from the problem inherent in many imaging studies thorst, V., Levine, A., . . . Strakowski, S.  M. (2006). Evidence of
of mental illnesses, namely that the correct region must be white matter pathology in bipolar disorder adolescents experienc-
chosen in order to identify meaningful connectivity dif- ing their first episode of mania: A diffusion tensor imaging study.
American Journal of Psychiatry, 163(2), 322–324. doi:10.1176/
ferences. Characterization of multiple seed regions might appi.ajp.163.2.322
address this concern, but few studies examine a large set of Adler, C. M., Holland, S. K., Schmithorst, V., Wilke, M., Weiss, K. L.,
regions. Also, the results of seed-region FC analyses provide Pan, H., & Strakowski, S. M. (2004). Abnormal frontal white mat-
ter tracts in bipolar disorder: A diffusion tensor imaging study. Bipo-
a fairly limited assessment of connectivity, only between a lar Disorders, 6(3), 197–203. doi:10.1111/j.1399-56004.00108.x
single brain region and elsewhere (i.e., first-order connec- Agarwal, N., Port, J.  D., Bazzocchi, M., & Renshaw, P.  F. (2010).
tions). The relationships among the other brain regions (i.e., Update on the use of MR for assessment and diagnosis of psychiatric
diseases. Radiology, 255(1), 23–41. doi:10.1148/radiol.09090339
second-order connections to the seed region) remain com- Altshuler, L.  L., Bartzokis, G., Grieder, T., Curran, J., Jimenez, T.,
pletely undescribed by this method. As a result, seed-region Leight, K., . . . Mintz, J. (2000). An MRI study of temporal lobe
FC does not provide a quantitative view of complex net- structures in men with bipolar disorder or schizophrenia. Biological
Psychiatry, 48(2), 147–162.
work structure in the brain. Altshuler, L.  L., Bartzokis, G., Grieder, T., Curran, J., & Mintz, J.
Graph theory methods have been adapted over the (1998). Amygdala enlargement in bipolar disorder and hippocampal
past decade or so to the study of brain network organiza- reduction in schizophrenia: An MRI study demonstrating neuro-
anatomic specificity. Archives of General Psychiatry, 55(7), 663–664.
tion (Bullmore & Sporns, 2009). These methods rely on Altshuler, L.  L., Bookheimer, S.  Y., Townsend, J., Proenza, M.  A.,
seed-region FC calculations as the starting point for assess- Eisenberger, N., Sabb, F., . . . Cohen, M.  S. (2005). Blunted activa-
tion in orbitofrontal cortex during mania:  A  functional magnetic
ing network structure. Unlike seed-region FC, interregional resonance imaging study. Biological Psychiatry, 58(10), 763–769.
correlations are calculated among every pair of regions (or doi:10.1016/j.biopsych.2005.09.012
sometimes voxels). This approach allows for characterization Altshuler, L.  L., Curran, J.  G., Hauser, P., Mintz, J., Denicoff, K., &
Post, R. (1995). T2 hyperintensities in bipolar disorder: Magnetic
of a wide variety of network measures, such as node degree resonance imaging comparison and literature meta-analysis. Ameri-
(the number of “significant” connections to a seed region can Journal of Psychiatry, 152(8), 1139–1144.
or seed voxel, i.e., node, but now calculated for every node). Aylward, E. H., Roberts-Twillie, J. V., Barta, P. E., Kumar, A. J., Har-
ris, G. J., Geer, M., . . . Pearlson, G. D. (1994). Basal ganglia volumes
Other network characteristics include degree distribution and white matter hyperintensities in patients with bipolar disorder.
(variation in degree from node to node), clustering coeffi- American Journal of Psychiatry, 151(5), 687–693.
cients (the amount of connectivity among nearest neighbors Bearden, C. E., Hoffman, K. M., & Cannon, T. D. (2001). The neuro-
psychology and neuroanatomy of bipolar affective disorder: a criti-
of a particular node, which is a characteristic of second-order cal review. Bipolar Disorders, 3(3), 106–150; discussion 151–103.
connections to a node), and path length and its inverse, effi- Bearden, C. E., Soares, J. C., Klunder, A. D., Nicoletti, M., Dierschke,
ciency (average distance between two nodes). These mea- N., Hayashi, K. M., . . . Thompson, P. M. (2008). Three-dimensional
mapping of hippocampal anatomy in adolescents with bipolar dis-
sures allow for characterization of whole-brain features of a order. Journal of the American Child & Adolescent Psychiatry, 47(5),
network. Recent research has found that connection density 515–525. doi:10.1097/CHI.0b013e31816765ab
(the spatial clustering of high-degree nodes) is predictive of Beaulieu, C. (2002). The basis of anisotropic water diffusion in the ner-
vous system—A technical review. NMR in Biomedicine, 15(7–8),
intelligence. The extensive and largely untapped possibilities 435–455. doi:10.1002/nbm.782
for examining whole-network connectivity in bipolar disor- Bermpohl, F., Dalanay, U., Kahnt, T., Sajonz, B., Heimann, H., Ricken,
der and the successful identification of relationships among R., . . . Bauer, M. (2009). A preliminary study of increased amygdala
activation to positive affective stimuli in mania. Bipolar Disorders,
behavior and connectivity measures (e.g., IQ; van den 11(1), 70–75. doi:10.1111/j.1399-5618.2008.00648.x
Heuvel, Stam, Kahn, & Hulshoff Pol, 2009) suggests that Beyer, J.  L., Taylor, W.  D., MacFall, J.  R., Kuchibhatla, M., Payne,
graph theory approaches could significantly advance under- M. E., Provenzale, J. M., . . . Krishnan, K. R. (2005). Cortical white
matter microstructural abnormalities in bipolar disorder. Neuropsy-
standing of the effects of bipolar disorder on brain structure chopharmacology, 30(12), 2225–2229. doi:10.1038/sj.npp.1300802
and function (Guye, Bettus, Bartolomei, & Cozzone, 2010).

2 2 0   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
Biederman, J., Petty, C. R., Wozniak, J., Wilens, T. E., Fried, R., Doyle, amygdala development in adolescents and young adults with bipolar
A., . . . Faraone, S.  V. (2011). Impact of executive function deficits disorder. Biological Psychiatry, 56(6), 399–405. doi:10.1016/j.bio-
in youth with bipolar I  disorder:  A  controlled study. Psychiatry psych.2004.06.024
Research, 186(1), 58–64. doi:10.1016/j.psychres.2010.08.029 Chen, C.  H., Suckling, J., Lennox, B.  R., Ooi, C., & Bullmore,
Biswal, B. B., Van Kylen, J., & Hyde, J. S. (1997). Simultaneous assess- E.  T. (2011). A quantitative meta-analysis of fMRI studies in
ment of flow and BOLD signals in resting-state functional connec- bipolar disorder. Bipolar Disorders, 13(1), 1–15. doi:10.1111/
tivity maps. NMR Biomedicine, 10(4–5), 165–170. j.1399-5618.2011.00893.x
Biswal, B., Yetkin, F. Z., Haughton, V. M., & Hyde, J. S. (1995). Func- Coffey, C. E., Wilkinson, W. E., Weiner, R. D., Parashos, I. A., Djang,
tional connectivity in the motor cortex of resting human brain W. T., Webb, M. C., . . . Spritzer, C. E. (1993). Quantitative cerebral
using echo-planar MRI. Journal of Magnetic Resonance Imaging, anatomy in depression:  A  controlled magnetic resonance imaging
34(4), 537–541. study. Archives of General Psychiatry, 50(1), 7–16.
Bitter, S. M., Mills, N. P., Adler, C. M., Strakowski, S. M., & DelBello, DelBello, M.  P., Strakowski, S.  M., Zimmerman, M.  E., Hawkins,
M. P. (2011). Progression of amygdala volumetric abnormalities in J. M., & Sax, K. W. (1999). MRI analysis of the cerebellum in bipo-
adolescents after their first manic episode. Journal of the American lar disorder: a pilot study. Neuropsychopharmacology, 21(1), 63–68.
Child & Adolescent Psychiatry, 50(10), 1017–1026. doi:10.1016/j. doi:10.1016/S0893-133X(99)00026-3
jaac.2011.07.001 DelBello, M. P., Zimmerman, M. E., Mills, N. P., Getz, G. E., & Stra-
Blumberg, H.  P., Kaufman, J., Martin, A., Whiteman, R., Zhang, kowski, S.  M. (2004). Magnetic resonance imaging analysis of
J.  H., Gore, J.  C., . . . Peterson, B.  S. (2003). Amygdala and hippo- amygdala and other subcortical brain regions in adolescents with
campal volumes in adolescents and adults with bipolar disorder. bipolar disorder. Bipolar Disorders, 6(1), 43–52.
Archives of General Psychiatry, 60(12), 1201–1208. doi:10.1001/ Dickstein, D. P., Milham, M. P., Nugent, A. C., Drevets, W. C., Char-
archpsyc.60.12.1201 ney, D.  S., Pine, D.  S., & Leibenluft, E. (2005). Frontotemporal
Blumberg, H. P., Leung, H. C., Skudlarski, P., Lacadie, C. M., Fred- alterations in pediatric bipolar disorder:  Results of a voxel-based
ericks, C. A., Harris, B. C., . . . Peterson, B. S. (2003). A functional morphometry study. Archives of General Psychiatry, 62(7), 734–741.
magnetic resonance imaging study of bipolar disorder: State- and doi:10.1001/archpsyc.62.7.734
trait-related dysfunction in ventral prefrontal cortices. Archives Diler, R. S., Segreti, A. M., Ladouceur, C. D., Almeida, J. R., Birmaher,
of General Psychiatry, 60(6), 601–609. doi:10.1001/arch- B., Axelson, D. A., . . . Pan, L. (2013). Neural correlates of treatment
psyc.60.6.601 in adolescents with bipolar depression during response inhibi-
Brambilla, P., Harenski, K., Nicoletti, M., Mallinger, A. G., Frank, E., tion. Journal of Child and Adolescent Psychopharmacology, 23(3),
Kupfer, D. J., . . . Soares, J. C. (2001). MRI study of posterior fossa 214–221. doi:10.1089/cap.2012.0054
structures and brain ventricles in bipolar patients. Journal of Psychi- Dima, D., Jogia, J., Collier, D., Vassos, E., Burdick, K. E., & Frangou,
atric Research, 35(6), 313–322. S. (2013). Independent modulation of engagement and connectivity
Brambilla, P., Harenski, K., Nicoletti, M., Sassi, R. B., Mallinger, A. G., of the facial network during affect processing by CACNA1C and
Frank, E., . . . Soares, J.  C. (2003). MRI investigation of temporal ANK3 risk genes for bipolar disorder. JAMA Psychiatry, 70(12),
lobe structures in bipolar patients. Journal of Psychiatric Research, 1303–1311. doi:10.1001/jamapsychiatry.2013.2099
37(4), 287–295. Drevets, W. C. (2000). Neuroimaging studies of mood disorders. Bio-
Bruno, S., Cercignani, M., & Ron, M.  A. (2008). White matter logical Psychiatry, 48(8), 813–829.
abnormalities in bipolar disorder:  A  voxel-based diffusion tensor Drevets, W. C., Price, J. L., Simpson, J. R., Jr., Todd, R. D., Reich, T.,
imaging study. Bipolar Disorders, 10(4), 460–468. doi:10.1111/ Vannier, M., & Raichle, M. E. (1997). Subgenual prefrontal cortex
j.1399-5618.2007.00552.x abnormalities in mood disorders. Nature, 386(6627), 824–827.
Bullmore, E., & Sporns, O. (2009). Complex brain networks:  graph doi:10.1038/386824a0
theoretical analysis of structural and functional systems. Nature Dupont, R.  M., Butters, N., Schafer, K., Wilson, T., Hesselink, J., &
Reviews Neuroscience, 10(3), 186–198. doi:10.1038/nrn2575 Gillin, J.  C. (1995). Diagnostic specificity of focal white matter
Caetano, S.  C., Sassi, R., Brambilla, P., Harenski, K., Nicoletti, M., abnormalities in bipolar and unipolar mood disorder. Biological
Mallinger, A.  G., . . . Soares, J.  C. (2001). MRI study of thalamic Psychiatry, 38(7), 482–486.
volumes in bipolar and unipolar patients and healthy individuals. Dupont, R.  M., Jernigan, T.  L., Heindel, W., Butters, N., Shafer, K.,
Psychiatry Research, 108(3), 161–168. Wilson, T., . . . Gillin, J. C. (1995). Magnetic resonance imaging and
Cardaso de Almeida, J.  R., & Phillips M.  L. (2013). Distinguishing mood disorders. Localization of white matter and other subcortical
between unipolar depression and bipolar depression: Current and abnormalities. Archives of General Psychiatry, 52(9), 747–755.
future clinical and neuroimaging perspectives. Biological Psychiatry, Eickhoff, S.  B., & Grefkes, C. (2011). Approaches for the integrated
73(2), 111–118. doi:10.1016/j.biopsych.2012.06.010. analysis of structure, function and connectivity of the human brain.
Chang, K., Karchemskiy, A., Barnea-Goraly, N., Garrett, A., Sime- Clinical EEG Neuroscience, 42(2), 107–121.
onova, D.  I., & Reiss, A. (2005). Reduced amygdalar gray matter Elvsashagen, T., Westlye, L. T., Boen, E., Hol, P. K., Andreassen, O. A.,
volume in familial pediatric bipolar disorder. Journal of the Ameri- Boye, B., & Malt, U. F. (2013). Bipolar II disorder is associated with
can Child & Adolescent Psychiatry, 44(6), 565–573. doi:10.1097/01. thinning of prefrontal and temporal cortices involved in affect regu-
chi.0000159948.75136.0d lation. Bipolar Disorders, 15(8), 855–864. doi:10.1111/bdi.12117
Chang, K. D., Wagner, C., Garrett, A., Howe, M., & Reiss, A. (2008). Filipek, P. A., Semrud-Clikeman, M., Steingard, R. J., Renshaw, P. F.,
A preliminary functional magnetic resonance imaging study of Kennedy, D. N., & Biederman, J. (1997). Volumetric MRI analysis
prefrontal-amygdalar activation changes in adolescents with bipo- comparing subjects having attention-deficit hyperactivity disorder
lar depression treated with lamotrigine. Bipolar Disorders, 10(3), with normal controls. Neurology, 48(3), 589–601.
426–431. doi:10.1111/j.1399-5618.2007.00576.x Fleck, D.  E., Eliassen, J.  C., Durling, M., Lamy, M., Adler, C.  M.,
Chen, C.  H., Ridler, K., Suckling, J., Williams, S., Fu, C.  H., DelBello, M.  P., . . . Strakowski, S.  M. (2012). Functional MRI of
Merlo-Pich, E., & Bullmore, E. (2007). Brain imaging correlates of sustained attention in bipolar mania. Molecular Psychiatry, 17(3),
depressive symptom severity and predictors of symptom improve- 325–336. doi:10.1038/mp.2010.108
ment after antidepressant treatment. Biological Psychiatry, 62(5), Fleck, D.  E., Nandagopal, J., Cerullo, M.  A., Eliassen, J.  C., Del-
407–414. doi:10.1016/j.biopsych.2006.09.018 Bello, M.  P., & Adler, C.  M., & Strakowski, S.  M. (2008).
Chen, B.  K., Sassi, R., Axelson, D., Hatch, J.  P., Sanches, M., Nico- Morphometric magnetic resonance imaging in psychiatry.
letti, M., . . . Soares, J. C. (2004). Cross-sectional study of abnormal

S tructural and F unctional M agnetic R esonance I maging F indings   •   2 21


Topics in Magnetic Resonance Imaging, 19(2), 131–142. doi:10.1097/ Biological Psychiatry, 57(7), 733–742. doi:10.1016/j.bio-
RMR.0b013e3181808152 psych.2005.01.002
Fleck, D. E., Shear, P. K., & Strakowski, S. M. (2009). Manic distract- Heinrich, A., Lourdusamy, A., Tzschoppe, J., Vollstadt-Klein, S.,
ibility and processing efficiency in bipolar disorder. In S. J. Woods, Buhler, M., Steiner, S., . . . Nees, F. (2013). The risk variant in
N. B. Allen, & C. Pantelis (Eds.), The neuropsychology of mental ill- ODZ4 for bipolar disorder impacts on amygdala activation during
ness. New York: Cambridge University Press. reward processing. Bipolar Disorders, 15(4), 440–445. doi:10.1111/
Foland, L. C., Altshuler, L. L., Sugar, C. A., Lee, A. D., Leow, A. D., bdi.12068
Townsend, J., . . . Thompson, P. M. (2008). Increased volume of the Hoge, E. A., Friedman, L., & Schulz, S. C. (1999). Meta-analysis of brain
amygdala and hippocampus in bipolar patients treated with lith- size in bipolar disorder. Schizophrenia Research, 37(2), 177–181.
ium. Neuroreport, 19(2), 221–224. Houenou, J., Wessa, M., Douaud, G., Leboyer, M., Chanraud, S., Perrin,
Foland-Ross, L. C., Thompson, P. M., Sugar, C. A., Madsen, S. K., Shen, M., . . . Paillere-Martinot, M. L. (2007). Increased white matter con-
J. K., Penfold, C., . . . Altshuler, L. L. (2011). Investigation of cortical nectivity in euthymic bipolar patients: diffusion tensor tractography
thickness abnormalities in lithium-free adults with bipolar I disor- between the subgenual cingulate and the amygdalo-hippocampal
der using cortical pattern matching. American Journal of Psychiatry, complex. Molecular Psychiatry, 12(11), 1001–1010. doi:10.1038/
168(5), 530–539. doi:10.1176/appi.ajp.2010.10060896 sj.mp.4002010
Foong, J., Symms, M. R., Barker, G. J., Maier, M., Miller, D. H., & Ron, Husain, M. M., McDonald, W. M., Doraiswamy, P. M., Figiel, G. S.,
M. A. (2002). Investigating regional white matter in schizophrenia Na, C., Escalona, P. R., . . . Krishnan, K. R. (1991). A magnetic reso-
using diffusion tensor imaging. NeuroReport, 13(3), 333–336. nance imaging study of putamen nuclei in major depression. Psy-
Frangou, S., Donaldson, S., Hadjulis, M., Landau, S., & Goldstein, chiatry Research, 40(2), 95–99.
L.  H. (2005). The Maudsley Bipolar Disorder Project:  Execu- Ives-Deliperi, V. L., Howells, F., Stein, D. J., Meintjes, E. M., & Horn,
tive dysfunction in bipolar disorder I  and its clinical corre- N. (2013). The effects of mindfulness-based cognitive therapy
lates. Biological Psychiatry, 58(11), 859–864. doi:10.1016/j. in patients with bipolar disorder:  A  controlled functional MRI
biopsych.2005.04.056 investigation. Journal of Affective Disorders, 150(3), 1152–1157.
Frazier, J. A., Breeze, J. L., Papadimitriou, G., Kennedy, D. N., Hodge, doi:10.1016/j.jad.2013.05.074
S.  M., Moore, C.  M., . . . Makris, N. (2007). White matter abnor- Jogia, J., Haldane, M., Cobb, A., Kumari, V., & Frangou, S. (2008).
malities in children with and at risk for bipolar disorder. Bipolar Pilot investigation of the changes in cortical activation during facial
Disorders, 9(8), 799–809. doi:10.1111/j.1399-5618.2007.00482.x affect recognition with lamotrigine monotherapy in bipolar disor-
Frazier, J. A., Chiu, S., Breeze, J. L., Makris, N., Lange, N., Kennedy, der. British Journal of Psychiatry, 192(3), 197–201. doi:10.1192/bjp.
D.  N., . . . Biederman, J. (2005). Structural brain magnetic reso- bp.107.037960
nance imaging of limbic and thalamic volumes in pediatric bipo- Jogia, J., Ruberto, G., Lelli-Chiesa, G., Vassos, E., Maieru, M., Tatarelli,
lar disorder. American Journal of Psychiatry, 162(7), 1256–1265. R., . . . Frangou, S. (2011). The impact of the CACNA1C gene poly-
doi:10.1176/appi.ajp.162.7.1256 morphism on frontolimbic function in bipolar disorder. Molecular
Grotegerd, D., Suslow, T., Bauer, J., Ohrmann, P., Arolt, V., Stuhrmann, Psychiatry, 16(11), 1070–1071. doi:10.1038/mp.2011.49
A., . . . Dannlowski, U. (2013). Discriminating unipolar and bipolar Kalmar, J. H., Wang, F., Chepenik, L. G., Womer, F. Y., Jones, M. M.,
depression by means of fMRI and pattern classification:  a pilot Pittman, B., . . . Blumberg, H. P. (2009). Relation between amygdala
study. European Archives of Psychiatry and Clinical Neuroscience, structure and function in adolescents with bipolar disorder. Jour-
263(2), 119–131. doi:10.1007/s00406-012-0329-4 nal of Academy of Child & Adolescent Psychiatry, 48(6), 636–642.
Guye, M., Bettus, G., Bartolomei, F., & Cozzone, P. J. (2010). Graph doi:10.1097/CHI.0b013e31819f6fbc
theoretical analysis of structural and functional connectivity MRI Krishnan, K.  R., McDonald, W.  M., Escalona, P.  R., Doraiswamy,
in normal and pathological brain networks. MAGMA, 23(5–6), P. M., Na, C., Husain, M. M., . . . Nemeroff, C. B. (1992). Magnetic
409–421. doi:10.1007/s10334-010-0205-z resonance imaging of the caudate nuclei in depression: Preliminary
Hafeman, D. M., Chang, K. D., Garrett, A. S., Sanders, E. M., & Phil- observations. Archives of General Psychiatry, 49(7), 553–557.
lips, M. L. (2012). Effects of medication on neuroimaging findings Kumar, A., Bilker, W., Jin, Z., & Udupa, J. (2000). Atrophy and high
in bipolar disorder:  An updated review. Bipolar Disorders, 14(4), intensity lesions:  Complementary neurobiological mechanisms
375–410. doi:10.1111/j.1399-5618.2012.01023.x in late-life major depression. Neuropsychopharmacology, 22(3),
Haldane, M., & Frangou, S. (2004). New insights help define the patho- 264–274. doi:10.1016/S0893-133X(99)00124-4
physiology of bipolar affective disorder: neuroimaging and neuropa- Kumari, V., & Cooke, M. (2006). Use of magnetic resonance imag-
thology findings. Progress in Neuro-Psychopharmacology & Biological ing in tracking the course and treatment of schizophrenia. Expert
Psychiatry, 28(6), 943–960. doi:10.1016/j.pnpbp.2004.05.040 Review Neurotherapy, 6(7), 1005–1016.
Haldane, M., Jogia, J., Cobb, A., Kozuch, E., Kumari, V., & Frangou, Kupferschmidt, D. A., & Zakzanis, K. K. (2011). Toward a functional
S. (2008). Changes in brain activation during working memory neuroanatomical signature of bipolar disorder:  Quantitative evi-
and facial recognition tasks in patients with bipolar disorder with dence from the neuroimaging literature. Psychiatry Research, 193(2),
Lamotrigine monotherapy. European Neuropsychopharmacology, 71–79. doi:10.1016/j.pscychresns.2011.02.011
18(1), 48–54. doi:10.1016/j.euroneuro.2007.05.009 Kyriakopoulos, M., Bargiotas, T., Barker, G. J., & Frangou, S. (2008).
Hallahan, B., Newell, J., Soares, J. C., Brambilla, P., Strakowski, S. M., Diffusion tensor imaging in schizophrenia. European Psychiatry,
Fleck, D.  E., . . . McDonald, C. (2011). Structural magnetic reso- 23(4), 255–273. doi:10.1016/j.eurpsy.2007.12.004
nance imaging in bipolar disorder:  An international collaborative Lyoo, I. K., Sung, Y. H., Dager, S. R., Friedman, S. D., Lee, J. Y., Kim,
mega-analysis of individual adult patient data. Biological Psychiatry, S. J., . . . Renshaw, P. F. (2006). Regional cerebral cortical thinning
69(4), 326–335. doi:10.1016/j.biopsych.2010.08.029 in bipolar disorder. Bipolar Disorders, 8(1), 65–74. doi:10.1111/
Harvey, I., Persaud, R., Ron, M.  A., Baker, G., & Murray, R.  M. j.1399-5618.2006.00284.x
(1994). Volumetric MRI measurements in bipolars compared with McDonald, C., Zanelli, J., Rabe-Hesketh, S., Ellison-Wright, I., Sham,
schizophrenics and healthy controls. Psychological Medicine, 24(3), P., Kalidindi, S., . . . Kennedy, N. (2004). Meta-analysis of mag-
689–699. netic resonance imaging brain morphometry studies in bipolar
Haznedar, M.  M., Roversi, F., Pallanti, S., Baldini-Rossi, N., disorder. Biological Psychiatry, 56(6), 411–417. doi:10.1016/j.bio-
Schnur, D.  B., Licalzi, E.  M., . . . Buchsbaum, M.  S. (2005). psych.2004.06.021
Fronto-thalamo-striatal gray and white matter volumes and McIntosh, A. M., Job, D. E., Moorhead, T. W., Harrison, L. K., For-
anisotropy of their connections in bipolar spectrum illnesses. rester, K., Lawrie, S.  M., & Johnstone, E.  C. (2004). Voxel-based

2 2 2   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
morphometry of patients with schizophrenia or bipolar disorder Rangel-Guerra, R.  A., Perez-Payan, H., Minkoff, L., & Todd, L.  E.
and their unaffected relatives. Biological Psychiatry, 56(8), 544–552. (1983). Nuclear magnetic resonance in bipolar affective disorders.
doi:10.1016/j.biopsych.2004.07.020 AJNR American Journal of Neuroradiology, 4(3), 229–231.
Mills, N. P., Delbello, M. P., Adler, C. M., & Strakowski, S. M. (2005). Rich, B. A., Fromm, S. J., Berghorst, L. H., Dickstein, D. P., Brotman,
MRI analysis of cerebellar vermal abnormalities in bipolar disorder. M. A., Pine, D. S., & Leibenluft, E. (2008). Neural connectivity in
American Journal of Psychiatry, 162(8), 1530–1532. doi:10.1176/ children with bipolar disorder:  Impairment in the face emotion
appi.ajp.162.8.1530 processing circuit. Journal of Child Psychology and Psychiatry, 49(1),
Moore, G. J., Bebchuk, J. M., Wilds, I. B., Chen, G., & Manji, H. K. 88–96. doi:10.1111/j.1469-7610.2007.01819.x
(2000). Lithium-induced increase in human brain grey matter. Lan- Rimol, L.  M., Hartberg, C.  B., Nesvag, R., Fennema-Notestine, C.,
cet, 356(9237), 1241–1242. Hagler, D. J., Jr., Pung, C. J., . . . Agartz, I. (2010). Cortical thickness
Norris, S. D., Krishnan, K. R ., & Ahearn, E. (1997). Structural changes in and subcortical volumes in schizophrenia and bipolar disorder. Bio-
the brain of patients with bipolar affective disorder by MRI: A review logical Psychiatry, 68(1), 41–50. doi:10.1016/j.biopsych.2010.03.036
of the literature. Progress in Neuro-Psychopharmacology & Biological Rimol, L.  M., Nesvag, R., Hagler, D.  J., Jr., Bergmann, O.,
Psychiatry, 21(8), 1323–1337. Fennema-Notestine, C., Hartberg, C.  B., . . . Dale, A.  M. (2012).
Ochsner, K.  N., & Gross, J.  J. (2005). The cognitive control of emo- Cortical volume, surface area, and thickness in schizophre-
tion. Trends in Cognitive Science, 9(5), 242–249. doi:10.1016/j. nia and bipolar disorder. Biological Psychiatry, 71(6), 552–560.
tics.2005.03.010 doi:10.1016/j.biopsych.2011.11.026
Passarotti, A.  M., Sweeney, J.  A., & Pavuluri, M.  N. (2011). Rocha-Rego, V., Jogia, J., Marquand, A. F., Mourao-Miranda, J., Sim-
Fronto-limbic dysfunction in mania pre-treatment and persistent mons, A., & Frangou, S. (2013). Examination of the predictive value
amygdala over-activity post-treatment in pediatric bipolar disor- of structural magnetic resonance scans in bipolar disorder: A pat-
der. Psychopharmacology (Berlin), 216(4), 485–499. doi:10.1007/ tern classification approach. Psychological Medicine, 44(3), 519–532.
s00213-011-2243-2 doi:10.1017/s0033291713001013
Pavuluri, M. N., Ellis, J. A., Wegbreit, E., Passarotti, A. M., & Stevens, Rosso, I.  M., Killgore, W.  D., Cintron, C.  M., Gruber, S.  A., Tohen,
M.  C. (2012). Pharmacotherapy impacts functional connectivity M., & Yurgelun-Todd, D.  A. (2007). Reduced amygdala volumes
among affective circuits during response inhibition in pediatric in first-episode bipolar disorder and correlation with cerebral white
mania. Behavioural Brain Research, 226(2), 493–503. doi:10.1016/j. matter. Biological Psychiatry, 61(6), 743–749. doi:10.1016/j.bio-
bbr.2011.10.003 psych.2006.07.035
Pavuluri, M.  N., Passarotti, A.  M., Fitzgerald, J.  M., Wegbreit, E., Sax, K.  W., Strakowski, S.  M., Zimmerman, M.  E., DelBello, M.  P.,
& Sweeney, J.  A. (2012). Risperidone and divalproex differen- Keck, P. E. Jr., & Hawkins, J. M. (1999). Frontosubcortical neuro-
tially engage the fronto-striato-temporal circuitry in pediatric anatomy and the continuous performance test in mania. American
mania: A pharmacological functional magnetic resonance imaging Journal of Psychiatry, 156(1), 139–141.
study. Journal of Academy of Child & Adolescent Psychiatry, 51(2), Scherk, H., & Falkai, P. (2004). Changes in brain structure caused by
157–170. doi:10.1016/j.jaac.2011.10.019 neuroleptic medication. Nervenarzt, 75(11), 1112–1117.
Pavuluri, M.  N., Passarotti, A.  M., Harral, E.  M., & Sweeney, J.  A. Scherk, H., & Falkai, P. (2006). Effects of antipsychotics on brain struc-
(2010). Enhanced prefrontal function with pharmacotherapy ture. Current Opinion in Psychiatry, 19(2), 145–150.
on a response inhibition task in adolescent bipolar disorder. Schlaepfer, T. E., Harris, G. J., Tien, A. Y., Peng, L. W., Lee, S., Feder-
Journal of Clinical Psychiatry, 71(11), 1526–1534. doi:10.4088/ man, E.  B., . . . Pearlson, G.  D. (1994). Decreased regional cortical
JCP.09m05504yel gray matter volume in schizophrenia. American Journal of Psychia-
Pavuluri, M. N., Passarotti, A. M., Lu, L. H., Carbray, J. A., & Sweeney, try, 151(6), 842–848.
J.  A. (2011). Double-blind randomized trial of risperidone versus Schnack, H.  G., Nieuwenhuis, M., van Haren, N.  E., Abramovic,
divalproex in pediatric bipolar disorder: fMRI outcomes. Psychiatry L., Scheewe, T.  W., Brouwer, R.  M., . . . Kahn, R.  S. (2013). Can
Research, 193(1), 28–37. doi:10.1016/j.pscychresns.2011.01.005 structural MRI aid in clinical classification? A  machine learning
Pavuluri, M. N., Schenkel, L. S., Aryal, S., Harral, E. M., Hill, S. K., study in two independent samples of patients with schizophrenia,
Herbener, E.  S., & Sweeney, J.  A. (2006). Neurocognitive func- bipolar disorder and healthy subjects. Neuroimage, 84c, 299–306.
tion in unmedicated manic and medicated euthymic pediatric doi:10.1016/j.neuroimage.2013.08.053
bipolar patients. American Journal of Psychiatry, 163(2), 286–293. Schneider, M. R., Adler, C. M., Whitsel, R., Weber, W., Mills, N. P.,
doi:10.1176/appi.ajp.163.2.286 Bitter, S. M., . . . Delbello, M. P. (2012). The effects of ziprasidone on
Pearlson, G.  D., Barta, P.  E., Powers, R.  E., Menon, R.  R., Richards, prefrontal and amygdalar activation in manic youth with bipolar
S.  S., Aylward, E.  H., . . . Tien, A.  Y. (1997). Ziskind-Somerfeld disorder. The Israel Journal of Psychiatry and Related Sciences, 49(2),
Research Award 1996: Medial and superior temporal gyral volumes 112–120.
and cerebral asymmetry in schizophrenia versus bipolar disorder. Schumann, C. M., Hamstra, J., Goodlin-Jones, B. L., Lotspeich, L. J.,
Biological Psychiatry, 41(1), 1–14. Kwon, H., Buonocore, M. H., . . . Amaral, D. G. (2004). The amyg-
Pessoa, L. (2008). On the relationship between emotion and cognition. dala is enlarged in children but not adolescents with autism:  The
Nature Reviews Neuroscience, 9(2), 148–158. doi:10.1038/nrn2317 hippocampus is enlarged at all ages. Journal of Neuroscience, 24(28),
Pfeifer, J. C., Welge, J., Strakowski, S. M., Adler, C. M., & DelBello, 6392–6401. doi:10.1523/JNEUROSCI.1297-04.2004
M.  P. (2008). Meta-analysis of amygdala volumes in children Selek, S., Nicoletti, M., Zunta-Soares, G.  B., Hatch, J.  P., Nery,
and adolescents with bipolar disorder. Journal of Academy of F.  G., Matsuo, K., . . . Soares, J.  C. (2013). A longitudinal study of
Child & Adolescent Psychiatry, 47(11), 1289–1298. doi:10.1097/ fronto-limbic brain structures in patients with bipolar I  disorder
CHI.0b013e318185d299 during lithium treatment. Journal of Affective Disorders, 150(2),
Phillips, M. L., Drevets, W. C., Rauch, S. L., & Lane, R. (2003). Neu- 629–633. doi:10.1016/j.jad.2013.04.020
robiology of emotion perception: I. The neural basis of normal emo- Shenton, M. E., Dickey, C. C., Frumin, M., & McCarley, R. W. (2001).
tion perception. Biological Psychiatry, 54(5), 504–514. A review of MRI findings in schizophrenia. Schizophrenic Research,
Ramnani, N., Behrens, T.  E., Penny, W., & Matthews, P.  M. (2004). 49(1–2), 1–52.
New approaches for exploring anatomical and functional connec- Singh, M.  K., & Chang, K.  D. (2012). The neural effects of psycho-
tivity in the human brain. Biological Psychiatry, 56(9), 613–619. tropic medications in children and adolescents. Adolescent Psy-
doi:10.1016/j.biopsych.2004.02.004 chiatric Clinics of North America., 21(4), 753–771. doi:10.1016/j.
chc.2012.07.010

S tructural and F unctional M agnetic R esonance I maging F indings   •   2 2 3


Singh, M. K., Delbello, M. P., Adler, C. M., Stanford, K. E., & Stra- intellectual performance. Journal of Neuroscience, 29(23),
kowski, S.  M. (2008). Neuroanatomical characterization of child 7619–7624. doi:10.1523/JNEUROSCI.1443-09.2009
offspring of bipolar parents. J Am Journal of American Academy Versace, A., Almeida, J.  R., Hassel, S., Walsh, N.  D., Novelli, M.,
of Child & Adolescent Psychiatry, 47(5), 526–531. doi:10.1097/ Klein, C. R., . . . Phillips, M. L. (2008). Elevated left and reduced
CHI.0b013e318167655a right orbitomedial prefrontal fractional anisotropy in adults
Strakowski, S. M., Adler, C. M., & DelBello, M. P. (2002). Volumetric with bipolar disorder revealed by tract-based spatial statistics.
MRI studies of mood disorders: Do they distinguish unipolar and Archives of General Psychiatry, 65(9), 1041–1052. doi:10.1001/
bipolar disorder? Bipolar Disorders, 4(2), 80–88. archpsyc.65.9.1041
Strakowski, S.  M., DelBello, M.  P., Adler, C., Cecil, D.  M., & Sax, Videbech, P., & Ravnkilde, B. (2004). Hippocampal volume and
K. W. (2000). Neuroimaging in bipolar disorder. Bipolar Disorders, depression: A meta-analysis of MRI studies. American Journal of
2(3 Pt. 1), 148–164. Psychiatry, 161(11), 1957–1966. doi:10.1176/appi.ajp.161.11.1957
Strakowski, S.  M., DelBello, M.  P., Sax, K.  W., Zimmerman, M.  E., Wang, F., Jackowski, M., Kalmar, J. H., Chepenik, L. G., Tie, K., Qiu,
Shear, P. K., Hawkins, J. M., & Larson, E. R. (1999). Brain magnetic M., . . . Blumberg, H. P. (2008). Abnormal anterior cingulum integ-
resonance imaging of structural abnormalities in bipolar disorder. rity in bipolar disorder determined through diffusion tensor imag-
Arch Gen Psychiatry, 56(3), 254–260. ing. British Journal of Psychiatry, 193(2), 126–129. doi:10.1192/bjp.
Strakowski, S. M., DelBello, M. P., Zimmerman, M. E., Getz, G. E., bp.107.048793
Mills, N. P., Ret, J., . . . Adler, C. M. (2002). Ventricular and peri- Wang, F., Kalmar, J.  H., Edmiston, E., Chepenik, L.  G., Bhagwagar,
ventricular structural volumes in first- versus multiple-episode Z., Spencer, L., . . . Blumberg, H.  P. (2008). Abnormal corpus cal-
bipolar disorder. American Journal of Psychiatry, 159(11), losum integrity in bipolar disorder:  A  diffusion tensor imag-
1841–1847. ing study. Biological Psychiatry, 64(8), 730–733. doi:10.1016/j.
Strakowski, S.  M., Wilson, D.  R., Tohen, M., Woods, B.  T., Doug- biopsych.2008.06.001
lass, A. W., & Stoll, A. L. (1993). Structural brain abnormalities in Whalley, H.  C., Papmeyer, M., Sprooten, E., Lawrie, S.  M., Suss-
first-episode mania. Biological Psychiatry, 33(8–9), 602–609. mann, J.  E., & McIntosh, A.  M. (2012). Review of functional
Swayze, V.  W. II, Andreasen, N.  C., Alliger, R.  J., Yuh, W.  T., & magnetic resonance imaging studies comparing bipolar disorder
Ehrhardt, J. C. (1992). Subcortical and temporal structures in affec- and schizophrenia. Bipolar Disorders, 14(4), 411–431. doi:10.1111/
tive disorder and schizophrenia:  A  magnetic resonance imaging j.1399-5618.2012.01016.x
study. Biological Psychiatry, 31(3), 221–240. Yamasaki, H., LaBar, K. S., & McCarthy, G. (2002). Dissociable pre-
Taylor, V.  A., Grant, J., Daneault, V., Scavone, G., Breton, E., frontal brain systems for attention and emotion. Proceedings of
Roffe-Vidal, S., . . . Beauregard, M. (2011). Impact of mindfulness the American Academy of Sciences USA, 99(17), 11447–11451.
on the neural responses to emotional pictures in experienced and doi:10.1073/pnas.182176499
beginner meditators. Neuroimage, 57(4), 1524–1533. doi:10.1016/j. Yang, H., Lu, L.  H., Wu, M., Stevens, M., Wegbreit, E., Fitzgerald,
neuroimage.2011.06.001 J., . . . Pavuluri, M. N. (2013). Time course of recovery showing ini-
Tesli, M., Skatun, K.  C., Ousdal, O.  T., Brown, A.  A., Thoresen, C., tial prefrontal cortex changes at 16 weeks, extending to subcortical
Agartz, I., . . . Andreassen, O.  A. (2013). CACNA1C risk vari- changes by 3 years in pediatric bipolar disorder. Journal of Affective
ant and amygdala activity in bipolar disorder, schizophrenia and Disorders, 150(2), 571–577. doi:10.1016/j.jad.2013.02.007
healthy controls. PLoS One, 8(2), e56970. doi:10.1371/journal. Yurgelun-Todd, D.  A., Silveri, M.  M., Gruber, S.  A., Rohan, M.  L.,
pone.0056970 & Pimentel, P. J. (2007). White matter abnormalities observed in
Townsend, J., & Altshuler, L.  L. (2012). Emotion processing and bipolar disorder: A diffusion tensor imaging study. Bipolar Disor-
regulation in bipolar disorder:  A  review. Bipolar Disorders, 14(4), ders, 9(5), 504–512. doi:10.1111/j.1399-5618.2007.00395.x
326–339. doi:10.1111/j.1399-5618.2012.01021.x
van den Heuvel, M.  P., Stam, C.  J., Kahn, R.  S., & Hulshoff Pol,
H.  E. (2009). Efficiency of functional brain networks and

2 2 4   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
18.
M AGNETIC R ESONA NCE SPECTROSCOPY FINDINGS
A ND THEIR TR EATMENT IMPACT
Marsal Sanches, Frank K. Hu, and Jair C. Soares

INTRODUCTION This chapter comprises a critical analysis of the current


evidence on the potential role of magnetic resonance spec-
Bipolar disorder (BD) is a chronic and potentially severe troscopy (MRS), one of the currently available brain imag-
psychiatric condition that affects up to 4% of the popu- ing approaches, in the treatment of BD. It begins with basic
lation (Merikangas & Lamers, 2012). In addition to its concepts of MRS techniques, followed by a review of the
considerable impact on the psychological well-being and most consistent MRS findings in BD. Next, MRS results
functional impact on patients, BD is associated with high that seem to reflect the effects of medications on the brain
rates of suicide and is currently listed as one of the leading are described, as well as findings associated with clinical
causes of disability worldwide (Zimmerman, 2010). response to certain treatments. Finally, some perspectives
Since the description of the therapeutic effects of lithium regarding the potential applications of MRS for the treat-
in the late 1940s, treatment of BD has undergone count- ment of BD are discussed.
less progress. In addition to mood stabilizers, the advent of
atypical antipsychotics revolutionized the pharmacologi-
cal management of this condition (Sanches, Newberg, & B A S IC C ONC E P T S OF 
Soares, 2010). Yet available evidence demonstrates that lack M R S I M AG I NG
of adequate treatment response is common among bipo-
lar patients. Prospective studies have shown that residual MRS is a neurochemical imaging technique that allows
depressive symptoms, frequent cycling, and chronic cogni- measurement of different molecules in the living brain,
tive impairment seem to be the rule rather than the excep- including metabolites and some medications (Figure 18.1).
tion in BD (Miklowitz, 2011; Vieta, Reinares, & Rosa, It is based on the same principle as magnetic resonance
2011). Longitudinal results from the STEP-BD study imaging (MRI) and utilizes a strong magnetic field and a
indicate that, over a two-year period of follow-up, 48.5% radio frequency current, allowing detection of the mag-
of bipolar patients experienced recurrence of mood symp- netic signal of certain isotopes present in different neuro-
toms. Moreover, among patients who were symptomatic chemicals. A chemical spectra, containing different peaks,
at the time of inclusion in the study, just 58.5% achieved is then obtained (Figure 18.2). The area under each peak
recovery (Perlis et al., 2006). reflects the concentration of the measured chemical in the
Over the past two decades, neuroimaging studies brain (Sassi & Soares, 2003).
have provided valuable advances in the understanding of For example, proton spectroscopy (Table 18.1) enables
the brain circuits involved in the pathophysiology of BD the measurement of neurotransmitters such as gamma
(Brambilla, Glahn, Balestrieri, & Soares, 2005; Brambilla, amino butyric acid (GABA), glutamate, and glutamine,
Hatch, & Soares, 2008; Soares & Mann, 1997). More as well as myo-inositol, choline, and N-acetylaspartate, an
recently, emphasis has been placed on the clinical implica- indicator of neuronal viability (Stanley, 2002). On the other
tion of these findings, with the ultimate goal of advancing hand, 31-phosphorus spectroscopy allows the measurement
the diagnosis and management of this illness (Phillips, of neurochemicals involved in the neuronal membrane
Travis, Fagiolini, & Kupfer, 2008; Soares, 2002). phospholipid turnover, including adenosine triphosphate,

2 25
(a) (b)

(c)

Figure 18.1 
Sagittal (a), axial (b), and coronal (c) views of 1H spectroscopy images showing the location of the 8 cm3 voxel in the left dorsolateral
prefrontal cortex (Reproduced from Nery et al., 2009, with permission)

phosphocreatine, and phosphodiesters. Different imag- and usually requires the participation of a magnetic reso-
ing protocols allow the measurement of these metabolites nance physicist (Drost, Riddle, Clarke, & AAPM MR
in distinct brain areas. Furthermore, spectroscopy can be Task Group #9, 2002). Among other issues, the strength
valuable in the assessment of the pharmacokinetic prop- of the magnetic field is a particularly critical feature, which
erties of psychotropic drugs that contain isotopes such as impacts the spatial resolution and the signal-to-noise
7Li and 19F, whose brain concentration can be detected ratio (Dager, Corrigan, Richards, & Posse, 2008). Lower
through MRS (Soares, Boada, & Keshavan, 2000). strengths also implicate in prolonged acquisition time,
This technique soon became widely used for the assess- which can be troublesome when assessing certain groups
ment of brain neurochemicals across the lifespan in dif- of subjects, such as children, patients in acute mania,
ferent psychiatric conditions, including BD. Since MRS and claustrophobic subjects, who are supposed to remain
utilizes no ionizing radiation, it allows the performance motionless during the entire duration of the scan.
of repeated imaging assessments, which makes MRS well
suited for the identification of neurochemical differences
among bipolar patients across the different phases of the OV E RV I E W OF   M R S F I N DI NG S
illness (Sassi & Soares, 2003). It also facilitates the search I N   B I P OL A R DI S OR DE R
for treatment effects through the performance of imaging
studies pre- and posttreatment, allowing the identification
PROTON M R S
of possible correlations between the levels of certain neu-
rochemicals and the clinical response to the treatment in Available evidence suggests that N-acetylaspartate (NAA)
question. Nevertheless, the appropriate implementation of levels are decreased in the dorsolateral prefrontal cortex
a MRS acquisition protocol may be technically complex (DLPC), anterior cingulate gyrus, hippocampus, and

2 2 6   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
possibly the basal ganglia of patients with BD (Yildiz &
Ankerst, 2006). NAA is synthesized in the mitochondria
from the amino-acid aspartic acid. It is considered a marker
of neuronal viability, reflecting neuronal metabolism and
also mitochondrial function (Clark, 1998).
Although these findings regarding NAA levels in the
bipolar brain seem to be mood state-independent, it is
acquired and unclear, at this point, whether the changes in NAA levels
modeled among bipolar patients are mediated by genetic or environ-
spectrum
mental factors. Decreased NAA in the DLPC and cerebel-
lar vermis seems to be a consistent finding among children
residual
with BD (Yildiz & Ankerst, 2006), suggesting that these
NAA changes occur early in the course of illness. On the other
hand, a more recent study (Gallelli et al., 2005) addressed
specifically the potential role of decreased NAA as an endo-
PCr+Cr PCr+Cr phenotype for BD. In that study, 60 offspring (children and
GPC+PC
adolescents) of bipolar patients underwent MRS, aiming at
myo-Ins the measurement of NAA levels in the DLPC. No statisti-
Glu Glu Individual cally significant differences between offspring and controls
NAA
Gln
modeled were found in regard to NAA levels. The authors concluded
(main)
metabolites that NAA levels in the DLPC might not be used as a marker
of vulnerability for BD and that decreases in NAA among
4.0 3.6 3.2 2.8 2.4 2.0 1.6 1.2 0.80 0.40
bipolar patients possibly take place later in the course of
Chemical shift (ppm)
illness. However, we should note that how many of these
Figure 18.2 
An example of quantifying a 1H spectrum acquired from children will develop BD is yet unknown; a longitudinal
the dorsolateral prefrontal cortex. The modeled spectrum (thick line) MRS investigation in thee subgroups with manifestation of
is superimposed on the acquired spectrum, and the residuals and the
individual modeled curves of the main metabolites are shown below. BD would be much informative. In addition, despite some
NAA = N-acetyl-aspartate, Glu = glutamate, Gln = glutamine, PCr + negative studies, it is well established that lithium use is
Cr = phosphocreatine plus creatine, GPC + PC = choline-containing
compounds, myo-Ins = myo-inositol. (Reproduced from Nery et al.,
associated with increases in the NAA levels in several brain
2009, with permission). regions (see “Medication Effects in MRS Studies” section).

Table 18 .1.  PATHOPHYSIOLOGICAL R ELEVANCE OF MAIN BR AIN METABOLITES QUANTIFIED THROUGH


PROTON MAGNETIC R ESONANCE SPECTROSCOPYa

METABOLITE R ELEVANCE

N-acetylaspartate Synthesized in the mitochondria from the amino-acid aspartic acid. Considered a marker of neuronal
viability, reflecting neuronal metabolism and also mitochondrial function
Myo-inositol Results from the dephosphorylation of inositol monophosphates by the enzyme inositol monophospha-
tase. Involved in maintenance of calcium homeostasis and activity of protein kinase C, with eventual
effects on neuronal cytoskeletal restructuring

Glutamate The main excitatory neurotransmitter in the central nervous system. Increased glutamatergic activity
seems to produce neurotoxicity. Due to technical issues and overlapping between the glutamate, GABA,
and glutamine peaks, glutamine is often adopted as a common description of these three combined
metabolites.

Creatine/phosphocreatine The creatine/phosphocreatine system works as a storage and transportation mechanism for high-energy
phosphate bonds. Usually adopted as a standard for comparison, due to its relatively stability

Choline Reflects density of cell membranes, as well as membrane and myelin turnover

Gamma aminobutyric acid The most important inhibitory neurotransmitter in the brain. Results from glutamate decarboxylation by
the enzyme glutamic acid decarboxylase
a
Based on Stanley (2002); Sassi and Soares (2003); Maddock and Buonocore (in press).

M agnetic R esonance S pectroscopy F indings and T heir T reatment   I mpact   •   2 2 7


On the other hand, in a more recent meta-analysis, 43 in bipolar patients compared to healthy subjects when all
studies on MRS findings among BD patients were pooled brain areas were combined and also when results on the
(Kraguljac et  al., 2012). Analyses of NAA, creatine, and frontal brain areas were analyzed separately. The authors
choline in the frontal lobe (anterior cingulate cortex and argue that the contribution of glutamine to the glutamate
DLPFC), hippocampus, and basal ganglia were performed. plus glutamine signal is relatively small and that their find-
Results indicated that decreased NAA levels in the basal ings are consistent with increased glutamine in the bipolar
ganglia were the most consistent MRS finding among BD brain. However, when whole-brain glutamate was specifi-
patients. Interestingly, the analysis showed increased NAA cally analyzed, just a statistical trend toward higher levels of
levels in the DLPC of bipolar patients compared to healthy glutamate among patients was found.
controls. The authors suggest that their findings are consis- The creatine/phosphocreatine ratio has also been
tent with evidence describing direct correlations between assessed among patients with BD. The creatine/phos-
glutamate and NAA levels among healthy volunteers phocreatine system works as a storage and transportation
(Waddell et al., 2011) and point to the possibility that NAA, mechanism for high-energy phosphate bonds (Maddock
rather than just an indicator of neuronal viability, might also & Buonocore, 2012). It is usually adopted as a standard
be involved in other pathophysiological processes still poorly for comparison due to its relatively stability. Some studies
understood. However, the authors of the meta-analysis in have found significant differences in the levels of creatine
question were not able to carry out subgroup analysis tak- in different brain areas among bipolar patients compared
ing into account possible medication effects, which could to controls, but the results are inconsistent in regard to
explain the discrepant findings mentioned earlier. the relationship between these findings and the differ-
Moreover, myo-inositol levels seem to be abnormal ent phases of illness, the brain areas where these findings
in the frontal lobes, temporal lobes, cingulate gyrus, and seem to be more prominent (hippocampus, thalamus,
basal ganglia of bipolar patients during the depressive and frontal lobe), and their actual meaning under a patho-
manic phases. While manic patients seem to have increased physiological point of view (Yildiz & Ankerst, 2006).
myo-inositol levels when compared to controls, results Finally, there is evidence suggesting increased brain
among patients with bipolar depression are less consis- choline levels among BD patients. Patients with bipolar
tent, with reports of increased and decreased myo-inositol depression seem to have higher choline levels in the basal
concentrations (Silverstone, McGrath, & Kim, 2005). ganglia and anterior cingulate cortex when compared
Myo-inositol is a downstream product of the phospha- to healthy controls. The choline levels are regarded as a
tidylinositol cycle and results from the dephosphoryla- measure of membrane phospholipids metabolism, since
tion of inositol monophosphates by the enzyme inositol it seems to reflect not only the density of cell membranes
monophosphatase. Dysfunctions in this cycle may affect but also cell membrane and myelin turnover (Maddock
calcium homeostasis and produce abnormal activity of pro- & Buonocore, 2012). Of note, lithium treatment seems to
tein kinase C, with eventual disruptions in neuronal cyto- have a decreasing effect on choline (see “Medication Effects
skeletal restructuring (Machado-Vieira, Manji, & Zarate, in MRS Studies” section).
2009). Lithium is an inhibitor of inositol monophospha-
tase (Yildiz, Demopulos, Moore, Renshaw, & Sachs, 2001),
PHO S PHORUS M R S
and its therapeutic effect in the treatment of BD led to the
so-called inositol depletion hypothesis of BD. A  critical Through phosphorus magnetic resonance spectros-
appraisal of this hypothesis in light of available evidence is copy, it is possible to separate metabolites containing
discussed in the next section. the 31-phosphorous nucleus into different peaks, which
Studies also suggest that glutamate levels are increased reflect phospholipid anabolism (phosphomonoesters, such
in several brain regions (including the parietal area, as phosphocholine and phosphoethanolamine), catabo-
DLPFC, and basal ganglia) among BD patients indepen- lism (phosphodiesters, like glycerophosphocholine and
dently of mood state or treatment status (Yildiz & Ankerst, glycerophosphoethanolamine), neuronal pH (inorganic
2006). Glutamate is the main excitatory neurotransmit- phosphate), as well as phosphocreatine and adenosine tri-
ter in the central nervous system, and increased gluta- phosphate. Phospholipids are major constituents of cell
matergic activity seems to be the result in neurotoxicity. membranes, and alterations in membrane phospholipid
A  recent meta-analysis (Gigante et  al., 2012)  reviewed 17 composition may be related to the cognitive impairment
studies describing glutamate findings among BD patients found in different conditions, including BD (Agarwal,
and found increased levels of glutamate plus glutamine Port, Bazzocchi, & Renshaw, 2010).

2 2 8   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
Moreover, state-dependent alterations in membrane evaluation of medication effects on the brain through the
phospholipids have been reported among bipolar patients performance of repeated imaging assessments, allowing the
(Yildiz, Sachs, Dorer, & Renshaw, 2001). Euthymic patients comparison of pre- versus posttreatment imaging findings.
with BD seem to have lower phosphomonoester (PME) lev- In this sense, the comparison of MRS findings in treated
els in the frontal lobe and possibly the temporal lobe when versus untreated patients seems to be particularly attractive,
compared to healthy controls, as well as lower whole-brain given the relatively safety of this technique and its high sen-
pH. On the other hand, patients acutely depressed or manic sitivity in the detection of overtime changes.
have higher pH and PME levels compared to euthymic
patients. No differences seem to be present between bipo-
LITHIUM
lar patients and controls in regard to phosphodiesters levels
(for a review, see Yildiz, Sachs, et al., 2001). These findings One of the most effective medications for the treatment
suggest that trait- and state-dependent disruptions in mem- of BD, lithium seems to have markedly neuroprotec-
brane phospholipid metabolism seem to be involved in the tive effects. MRS findings among lithium-treated bipo-
pathophysiology of BD. lar patients might be able to fill this gap, contributing to
a better understanding of the mechanisms involved in the
mood-stabilizing effect of this medication.
M E DIC AT ION E F F E C T S I N   M R S For instance, some studies found no differences
S T U DI E S:  T H E B IOL O G IC A L between lithium-treated BD patients and healthy controls
B A S I S OF   R E C OV E RY in terms of NAA levels in the anterior cingulate cortex,
I N   B I P OL A R DI S OR DE R prefrontal cortex, and DLPC. Since evidence from studies
with untreated patients overall point to decreased levels of
The putative effect of treatment status on neuroimaging NAA in BD, these findings may reflect a possible normaliz-
findings in psychiatric disorders, including BD, has tradi- ing effect of lithium treatment on NAA (Yildiz & Ankerst,
tionally been regarded as a potential confounding factor. 2006). This hypothesis is supported by other cross-sectional
Consequently, the inclusion of patients currently treated studies that describe increased NAA levels in the temporal
with psychotropics in neuroimaging studies is usually listed lobe and basal ganglia of lithium-treated bipolar subjects.
as a methodological limitation, often evoked as a possible Similarly, Hajek et  al. (2012) showed significant lower
explanation for negative findings in cross-sectional studies. DLPFC NAA levels among untreated BD patients com-
Nevertheless, in recent years this position has been revis- pared to healthy controls, whose NAA levels were similar
ited, not only due to difficulties in studying unmedicated to those of lithium-treated bipolar patients. On the other
patients but also to the fact that including medicated sub- hand, several different cross-sectional studies failed to dem-
jects can provide valuable information about the mecha- onstrate significant differences between lithium-treated
nism of action of psychotropics, medications currently bipolar patients and controls in regard to NAA levels in dif-
available for the treatment of mental disorders (Table ferent brain regions, including the anterior cingulate cortex
18.2). Furthermore, prospective designs allow the direct and the basal ganglia (DelBello & Strakowski, 2004).

Table 18 .2 .  OVERVIEW OF EVIDENCE ON MEDICATION EFFECTS IN MAGNETIC R ESONANCE


SPECTROSCOPY STUDIES AMONG BIPOLAR PATIENTS

POSSIBLE EFFECT ON METABOLITES (DIFFER ENT R EGIONS OF INTER EST)


MEDICATIONS ↑ IN NA A ↓ IN MYO-INOSITOL ↓ IN GLUTA M ATE ↑ IN GABA ↓ IN CHOLINE

Lithium +++ a
+ b
+ – +
VPA – + + + –

Lamotrigine – – + – –

Atypical antipsychotics + – + – +

NOTE: NAA = N-acetylaspartate; GABA = gamma-aminobutyric acid; VPA = valproic acid.


a
Positive results regarding mainly the dorsolateral prefrontal cortex, anterior cingulate, and prefrontal cortex; may represent a disease-specific effect (not found
among controls medicated with lithium).
b
Results are overall inconsistent; one study shows increases in myo-inositol levels among lithium-treated patients (Sharma et al., 1992).

M agnetic R esonance S pectroscopy F indings and T heir T reatment   I mpact   •   2 2 9


Other studies analyzed the longitudinal effect of lith- available mood stabilizers, including lithium, apparently
ium treatment on MRS findings. Again, results seem to do not play a significant role in the modulation of the glu-
be inconsistent. At least two studies described positive tamatergic system. Alternatively, technical limitations on
results changes in NAA levels associated with lithium obtaining individual peaks for glutamate, glutamine, and
treatment. In one of these studies (Moore et  al., 2000), GABA, enabling their individual concentration estimates
12 medication-free BD patients and 9 healthy controls in the commonly used lower field magnets might have
underwent MRS at baseline and after four weeks of lith- masked disease- or treatment-induced alterations in gluta-
ium administration. Results indicated statistically signifi- mate. In contrast, another study demonstrated decreases
cant increases in NAA levels in all brain areas analyzed, in the glutamine and glutamate plus glutamine levels in
including the frontal, temporal, and occipital lobes. In the basal ganglia of bipolar patients following two weeks
contrary, another study, which analyzed adolescents with of lithium administration (Shibuya-Tayoshi et  al., 2008).
bipolar depression, described decreases in ventral prefron- GABA levels, on the contrary, did not show any significant
tal NAA levels after six weeks of lithium treatment (Patel changes associated with lithium administration.
et al., 2008). Also utilizing a prospective design, our group Another study (Colla et al., 2009) attempted to address
(Brambilla et al., 2004) failed to demonstrate any increases the possible role of lithium in modulating and counterbal-
in the DLPFC NAA levels among healthy controls after ancing the effects of hypothalamic–pituitary–adrenal axis
four weeks of lithium treatment. These results suggest that hyperactivity on the hippocampus among bipolar patients.
increases in NAA may be a specific effect of lithium on the In that study, 21 euthymic patients with BD type I  on
brain of bipolar patients, not found among patients who are chronic lithium maintenance treatment underwent proton
in depressive phase of the illness or do not suffer from that MRS of the left and right hippocampus, as well as measure-
disorder. ment of salivary cortisol. While absolute concentrations of
Further, given the putative effect of lithium on the NAA, choline, and creatine were similar among patients
phosphatidylinositol cycle (see previous section), the pos- and controls, hippocampal glutamate concentrations were
sible impact of lithium treatment on myo-inositol levels is higher among patients. Further, patients and controls dis-
of high interest (Yildiz, Demopulos, et al., 2001). However, played inverse correlations between salivary cortisol levels
most studies have failed to demonstrate significant dif- and not only hippocampal NAA values but also glutamate
ferences between lithium-treated bipolar patients and levels. These results suggest that the neurotropic effects
controls in regard to myo-inositol levels in different brain of lithium, which may counterbalance the toxic effects of
areas. Sharma, Venkatasubramanian, Bárány, and Davis increased hypothalamic–pituitary–adrenal activity on the
(1992) found increased levels of myo-inositol in the basal hippocampus, seem to take place through mechanisms dis-
ganglia among lithium-treated patients. In another study tinct from decreasing the glutamatergic activity.
(Forester et al., 2008), utilizing a different design, nine geri- Overall, results pointing to the effects of lithium treat-
atric patients with BD type I underwent proton MRS and ment on MRS findings are only partially consistent. Several
lithium spectroscopy. Results indicated direct correlations methodological limitations may account for these hetero-
between lithium levels and brain NAA and myo-inositol geneous results, including variable duration of lithium
levels. Yet in a recent review of available evidence, it was treatment, different lithium doses, and distinct history of
proposed that, regardless of the promising evidence regard- previous lithium exposure. One of the studies mentioned
ing the effect of lithium on inositol monophosphatase, it is previously (Hajek et  al., 2012)  suggested that duration of
unlikely that lithium might produce a persistent reduction illness and illness burden might affect the brain response to
in myo-inositol levels and phosphatidylinositol cycle turn- lithium, and these factors should be controlled when assess-
over (Silverstone et al., 2005). ing the neurochemical changes associated with lithium
With respect to other metabolites, a recent study treatment.
described significantly higher glutamine/glutamate ratio in
the anterior cingulate cortex and parieto-occipital cortex
VA L PROIC AC I D
among BD patients in acute mania compared to controls
(Ongür et  al., 2008). No differences were found between Valproic acid (VPA) is widely used for the treatment of
unmedicated patients and those receiving lithium, anticon- BD, particularly manic states. It has strong GABAergic
vulsants, or benzodiazepines. These findings suggest that effects, which seem to account, at least in part, for its mood
while increased glutamatergic activity seems to be involved stabilizer properties. To date, very few studies have focused
in the pathophysiology of acute manic states, currently specifically on MRS findings associated with use of VPA.

2 3 0   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
In several studies, bipolar patients on VPA were included and no differences were found between medicated versus
in the same pool as patients receiving lithium or other anti- unmedicated subjects. Interestingly, the previously men-
convulsants, making the identification of VPA-specific tioned findings mirrored the results obtained in rodents
effects rather difficult. For example, in one study euthy- medicated with mood stabilizers, which displayed decreases
mic bipolar patients on divalproex and gabapentin were in the levels of myo-inositol and creatine (with lithium and
found to have increased occipital and medial prefrontal VPA ministration), increases in glutamine levels with con-
GABA levels when compared to healthy controls (Wang sequent decreases in the glutamate/glutamine ratio (with
et al., 2002). VPA only), and increases in GABA (with lithium only).
It has been suggested that VPA may share the same Furthermore, in rodent studies, no changes in the NAA
putative normalizing effect on some brain metabolites as levels were found in association with the administration
lithium. For instance, in Silverstone et  al.’s (2002) study, of mood stabilizers. The authors suggested that the lack of
patients on VPA were found to have myo-inositol and PME differences between medicated and unmedicated subjects is
levels similar to lithium treated patients and healthy con- most likely secondary to methodological issues, since it is
trols. In another study (O’Donnell et al., 2003), rats showed not possible to ascertain the patients’ compliance with their
statistically significant increases in the concentration of medications. The fact that some of the medicated patients
inositol monophosphates and decreases in the levels of had committed suicide may indicate that they had been
myo-inositol following lithium treatment. Administration noncompliant with the mood stabilizers or, alternatively,
of VPA produced similar results. The authors concluded that the medications had not been clinically effective. The
that lithium and VPA might share similar effects on the mirroring effect between the postmortem findings and
phosphatidylinositol cycle, which is probably related to the results among medicated rats supports the hypothesis
their mood stabilizer effect. of a normalizing effect of mood stabilizers on the levels of
In contrast, other studies point to differences between the metabolites in question. Finally, the lack of increases
lithium and VPA-treated patients in regard to MRS find- in NAA levels in response to the administration of mood
ings. Silverstone et al. (2003) described elevated concentra- stabilizers in rats is mentioned as an argument in favor of
tions of NAA among euthymic BD patients treated with the hypothesis that increases in NAA in response to mood
lithium but not with VPA. These results are partially in stabilizers may be a disease-specific finding, present only
agreement with another study (Cecil, DelBello, Morey, & among bipolar patients.
Strakowski, 2002), which described negative correlations In summary, these findings are heterogeneous and
between NAA concentration in the orbitofrontal cortex suggest that, while lithium and VPA might share some
and duration of VPA treatment among bipolar patients in common mechanisms in regard to their mood stabilizer
mania. In another study (Friedman et al., 2004), patients on properties, they might have distinct profiles regarding their
lithium were found to have decreased glutamate-glutamine neuroprotective effects.
and increased myo-inositol compared to controls, whereas
subjects on VPA failed to display any significant differences.
Similarly, another group compared patients on lithium
L A MOT R IG I N E
and VPA in regard to their choline levels measured through
MRS (Wu et al., 2004). When euthymic BD patients on An anticonvulsant, lamotrigine is Food and Drug
either lithium or VPA were pooled together and compared Administration approved for the maintenance treatment
to healthy controls, significantly lower temporal lobe cho- of bipolar I disorder, although there is evidence suggesting
line/creatine ratios were found among patients. On the it might also be effective for the treatment of acute bipolar
other hand, when only VPA-treated patients were com- depression. Its mechanism of action is complex and involves
pared to healthy controls, no differences on choline levels blockage of voltage-sensitive sodium channels, as well as
were identified. some possible neuroprotective effects through glutamater-
Finally, one study utilized proton MRS for the post- gic modulation (Sanches & Soares, 2011). However, scant
mortem assessment of the DLPFC among patients with evidence is currently available regarding brain metabolites
a history of BD (Lan et  al., 2009). Findings were then and lamotrigine. In rats, the administration of anticonvul-
compared to MRS findings in the brain of rats following sants (including lamotrigine) did not seem to be associ-
administration of mood stabilizers (lithium or VPA). The ated with increases in myo-inositol levels, contrary to the
brains of BD subjects displayed higher levels of glutamate, administration of lithium (McGrath, Greenshaw, McKay,
myo-inositol, and creatine when compared to controls, Slupsky, & Silverstone, 2007).

M agnetic R esonance S pectroscopy F indings and T heir T reatment   I mpact   •   2 31


On the other hand, in the only longitudinal study to treatment. However, some patients included in the analysis
assess the possible impact of lamotrigine use on MRS find- were receiving mood stabilizers, limiting the generalization
ings (Frye et  al., 2007), 23 unmedicated depressed BD of the findings.
patients were compared to 12 controls in regard to MRS Moore et  al. (2007) compared MRS findings among
findings. After 12 weeks of treatment with lamotrigine, unmedicated children with BD and bipolar children medi-
patients underwent a new MRS study. At baseline, patients cated with risperidone. Unmedicated patients showed lower
displayed lower levels of glutamate, glutamate plus gluta- glutamate plus glutamine/creatine ratios in the anterior
mine, and creatine plus phosphocreatine when compared cingulate cortex compared to risperidone-medicated chil-
to controls. Results after 12 weeks of treatment showed a dren. Of note, all unmedicated patients were manic at the
strong interaction with treatment response, with signifi- time of the scan, while medicated patients were euthymic.
cantly lower levels of glutamine among responders com- The authors argue that these findings suggest decreased
pared to nonresponders. metabolism in the anterior cingulate cortex of bipolar chil-
dren in mania, which could be reversed by the treatment
with risperidone. However, confounding by combined con-
AT Y PIC A L A N T I P S YCHOT IC S
centration estimates of three different metabolites of gluta-
Atypical antipsychotics are becoming increasingly popular mate, glutamine, and GABA is a concern.
for the treatment of BD, although marked differences seem Finally, one longitudinal study analyzed the impact of
to be present in regard to their efficacy across the different treatment with olanzapine on MRS findings among first
phases of the illness (Sanches & Soares, 2011). Over the episode manic adolescents (DelBello, Cecil, Adler, Daniels,
past few years, several MRS studies have addressed possible & Strakowski, 2006). Overall, olanzapine seemed to pro-
brain neurochemical changes associated with their use. In duce significant increases in ventral prefrontal choline lev-
some of these studies, atypical antipsychotics were com- els, but not in NAA. However, remitters and nonremitters
bined with mood stabilizers. seemed to display distinct patterns of changes in NAA con-
For example, Atmaca, Yildirim, Ozdemir, Ogur, and centrations. These and other results involving prediction of
Tezcan (2007) hippocampal MRS findings were compared response will be discussed at the next section of the present
across four different groups:  medication-free BD patient, chapter.
VPA-treated BD patients, VPA plus quetiapine treated–BD
patients, and healthy controls. Untreated patients showed
statistically significant lower NAA/creatine and NAA/ C A N M AG N E T IC R E S ON A NC E
choline ratios compared to both groups of treated patients, S PE C T RO S C OP Y F I N DI NG S
as well as to healthy controls. On the other hand, patients PR E DIC T R E S P ON S E
on VPA plus quetiapine showed higher NAA/choline ratios TO   T R E AT M E N T I N 
compared to bipolar patients treated only with VPA. These B I P OL A R PAT I E N T S?
findings suggest that augmentation of VPA with quetiap-
ine may produce additional neuroprotection compared to The identification of specific features able to predict the
VPA monotherapy. clinical response to certain therapeutic interventions for
In another cross-sectional study (Shahana et al., 2011), BD is possibly one of the most fascinating potential prac-
MRS findings on the caudate among unmedicated bipo- tical applications of neuroimaging techniques. However,
lar patients, medicated patients (most of whom receiving despite some promising results, evidence in that regard is
atypical antipsychotics), and healthy controls were com- still preliminary.
pared. Bipolar patients receiving atypical antipsychotics
displayed statistically significantly higher creatine/choline
PROTON M R S F I N DI NG S
ratios when compared to controls and unmedicated bipo-
lar patients, as well as significantly lower choline/NAA Some studies have attempted to identify predictors of
ratios when compared to unmedicated BD patients. These future response to lithium through the performance of
findings suggest that atypical antipsychotics may cause repeated MRS scans over the course of the treatment. In
decreases in the choline levels in the caudate of bipolar one (Moore et al., 1999), depressed bipolar patients under-
patients, a finding that may reflect a curing effect on the went MRS at baseline (when off medications for at least
putative oxidative injury (as well as some possible neuro- two weeks), after five to seven days on lithium and after
protective effect) associated with atypical antipsychotic three to four weeks of lithium treatment. Results indicated

2 3 2   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
rapid decreases in the myo-inositol levels in the right frontal on the glutamatergic system (Strawn et al., 2012). No differ-
lobe at the time of the second MRS assessment, which per- ences were found in the glutamate and glutamate plus gluta-
sisted until the time of the third imaging study. The authors mine levels at the different brain regions analyzed (anterior
concluded that although lithium seems to induce decreases cingulate cortex and left and right ventrolateral prefrontal
in the myo-inositol levels, these changes do not seem to cortex) when patients and controls were compared at base-
be correlated with clinical improvement in the depressive line. Among patients, those who achieved remission after
symptoms, since they seem to occur early in the course of four weeks of treatment displayed lower glutamate plus glu-
the treatment with lithium, before the onset of its thera- tamine concentrations at the left ventrolateral prefrontal
peutic effect. Alternatively, decreases in myo-inositol may cortex compared to nonremitters. Moreover, decreases in
represent the first step of a sequence of lithium-induced glutamate concentrations at the left ventrolateral prefron-
neurochemical changes, which are eventually responsible tal cortex showed direct correlations with changes in the
for its therapeutic effect. Young Mania Rating Scale scores among remitters. The
These results are in contrast those from Davanzo et al. authors concluded that response to treatment with dival-
(2001), in which 11 children with BD underwent MRS proex is most likely related to modulation of the glutama-
scans at baseline and after seven days of lithium treat- tergic system induced by that mood stabilizer.
ment. Differently from nonresponders, lithium respond- Finally, Change et  al. (2012) performed serial MRS
ers displayed a statistically significant decrease in the scans to analyze the impact of monotherapy treatment
myo-inositol/creatine ratio in the anterior cingulate cortex with quetiapine on the neurochemical aspects of bipolar
from baseline to follow-up. adolescents in the depressive phase. Patients were random-
Adopting a similar design, DelBello et  al. (2006) uti- ized to receive either quetiapine or placebo, and scans were
lized repeated proton MRS studies to search for brain performed at baseline and after eight weeks of treatment.
neurochemical changes associated with olanzapine mono- No differences were observed between the medication and
therapy treatment among bipolar adolescents with a first the placebo groups in regard to changes in the NAA levels
manic or mixed episode. Patients who achieved remission at the DLPF and anterior cingulate cortex. As previously,
after four weeks of treatment displayed different patterns of no baseline differences were found when responders were
change in their NAA levels in the medial ventral prefrontal compared with nonresponders regarding the levels of any
cortex compared to nonremitters. While among remitters metabolites. However, responders showed statistically sig-
the NAA levels increased from baseline to endpoint, the nificantly lower concentrations of myo-inositol in the ante-
opposite happened among nonremitters. Further, NAA rior cingulate cortex when compared to nonresponders.
levels among remitters showed a direct correlation with These findings point to the possibility that, similarly to
decreases in the Young Mania Rating Scale scores. Finally, lithium, quetiapine might also display effects on the phos-
although choline levels increased in response to treatment phatidylinositol cycle, which may be correlated with its
among remitters and nonremitters, baseline (pretreatment) clinical efficacy.
levels of choline at the medial ventral prefrontal cortex
were higher among remitters than nonremitters. These
findings suggest that response to olanzapine may be asso- PHO S PHORUS M R S
ciated with possible neurotropic effects of this medication
or with normalization of abnormal mitochondrial metabo- In one of the few studies that utilized phosphorus MRS
lism, reflected by the increases in NAA levels. The authors (P MRS) in an attempt to identify predictors of response
also suggested that the increase in choline levels induced by to lithium, 32 patients with BD type I underwent serial P
olanzapine may contribute to its therapeutic effect through MRS scans (Kato, Inubushi, & Kato, 2000). Responders
putative inhibition of hyperactive second-messenger sys- were found to have lower brain intracellular pH when com-
tems by choline and downstream inhibition of cell mem- pared to nonresponders. The authors hypothesized that
brane breakdown. This hypothesis would explain why the decreased pH might reflect differences in the patho-
patients with higher baseline concentrations of choline physiology of BD between responders and nonresponders,
would be more likely to respond to the treatment with although the actual meaning of these findings remains
olanzapine. unclear.
Similarly, 25 adolescent bipolar patients in manic/ Murashita, Kato, Shioiri, Inubushi, and Kato (2000)
mixed states on treatment with divalproex underwent serial also compared lithium-responsive, lithium-resistant, and
MRS scans to assess the possible impact of that medication healthy controls utilizing serial 31P MRS scans. The subjects

M agnetic R esonance S pectroscopy F indings and T heir T reatment   I mpact   •   2 33


underwent photic stimulation, and scans were performed levels and clinical features. Among older (>50 years) BD
at baseline and at two different endpoints during the photic patients, lithium levels showed a direct association with
stimulation treatment. Lithium-resistant patients, differ- frontal cognitive impairment and higher scores on the
ently from controls and lithium responders, showed lower Hamilton Depression Rating Scale. The authors argue
levels of phosphocreatine levels at the two follow-up scans that, among older patients, lower brain levels of lithium
than at the pretreatment scan. The authors concluded that might be more appropriate, since younger patients had
mitochondrial dysfunction might be involved in lithium similar brain concentrations but did not display the asso-
resistance in BD. ciations in question.
Thus in regard to 7Li MRS studies trying to predict
treatment response, results have mostly been inconclusive,
L I T H I U M M R S F I N DI NG S
largely due to poor anatomical resolution of the technique
Lithium MRS (7Li MRS) enables measurement of lith- (voxels are generally very large), as well as pronounced
ium levels in the living human brain, allowing not only variability in brain lithium concentration across different
the acquisition of pharmacokinetic data and regional dif- anatomical regions. It is unclear whether this might be
ferences in lithium levels at the brain but also the search a way forward in trying to understand lithium’s mecha-
for possible correlations for the brain lithium levels and nisms of action or predicting and monitoring treatment
response to lithium treatment in BD (Lyoo & Renshaw, response.
2003; Sanches, Brambilla, & Soares, 2006; Soares et  al.,
2000). Nevertheless, while several studies over the past two
decades have addressed lithium pharmacokinetics, little PE R S PE C T I V E S ON   M R S A N D I T S
research has analyzed brain lithium concentrations as a I M PAC T ON   T H E T R E AT M E N T
predictor of response. OF   B I P OL A R DI S OR DE R
Overall, studies point to a positive correlation between
brain and serum lithium levels among bipolar patients. Despite the growing number of MRS studies in BD and
However, the strength of this correlation seems to decrease some promising results regarding its potential role in
when a narrower range of serum lithium levels is consid- improving the treatment of this condition, currently avail-
ered (Sachs et al., 1995). These results suggest that, within able evidence is still in a preliminary phase. Relatively few
the usual therapeutic range adopted in clinical practice in studies have addressed predictors of response to treatment,
regard to serum lithium levels, the correlation with brain and overall MRS results on medication effects are not very
lithium levels might not be as strong as usually assumed (for consistent.
a review see DelBello & Strakowski, 2004). Several methodological factors may account for these
Some studies have attempted to identify a minimal brain heterogeneous results. Some of them are methodological in
concentration of lithium necessary for clinical response to nature, including
lithium. In Kushnir et al. (1993), no statistically significant
differences between lithium responders and nonresponders 1. Variability regarding the choice of the areas of inter-
were identified among patients whose serum concentration est. Current evidence suggests that the areas that most
ranged from 0.28 to 0.86 mEq/L. In Kato et al. (1994), poor consistently display positive findings in MRS studies
response was significantly correlated with brain levels of of BD are the DLPFC, the anterior cingulate cor-
lithium below 0.2 mEq/L. In the same study, brain lithium tex, the prefrontal cortex, the hippocampus, and the
concentration was strongly correlated with improvement basal ganglia. Due to technical issues regarding the
in manic symptoms. While improvement was also corre- acquisition of MRS data and the long period of time
lated with the brain/serum lithium ratio, it was not corre- to complete an MRS sequence when using scans with
lated with serum concentration per se. Finally, in regard to lower field strength, most of the currently available
side effects, patients who experience lithium-related hand studies seem to prioritize just some of these areas. The
tremor seem to have significantly higher brain concentra- growing availability of scans with higher field strength
tions of lithium than those who do not display that side should allow the performance of more comprehen-
effect (Kato, Fujii, Shioiri, Inubushi, & Takahashi, 1996). sive spectroscopy studies, allowing the comparison
Moreover, in a more recent study (Forester et al., 2009), of results across different groups regarding the pos-
younger and older bipolar patients were found to have sible impact of treatment on different brain regions.
distinct patterns of association between brain lithium Similarly, technical improvements should allow the

2 3 4   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
acquisition of 31P MRS and 3Li MRS data with better in a timely manner. It is possible that the current MRS,
spatial resolution and greater reliability. although sound, are not sensitive enough to detect more
subtle neurochemical changes in different metabolites that
2. The longitudinal and changeable nature of BD. While
could take place downstream and might be more directly
some of the currently available results come from
associated with clinical response. The constant improve-
samples comprised of euthymic patients, others were
ment in the technical aspects of MRS should be able to
obtained from patients in acute mania or during
address these issues in the near future.
depressive episodes. Given the cyclical course of BD, it
is expected that distinct MRS patterns are associated
with response to treatment across the different phases
of the disease. Cross-sectional studies aiming at treat- C ONC LUS ION
ment effects, although easy to implement, provide very
limited information in a condition like BD due to the After two decades of research utilizing these techniques
potential presence of confounders and large variability in the field of psychiatry, MRS has served as a valuable
in terms of duration of illness, severity of mood symp- research tool, but it has yet to yield much that can be trans-
toms, degree of functional impact, and previous treat- lated clinically. This is not unique to MRS as a technique
ment history. Longitudinal studies examining the same and corroborates the complexity of the diseases we study,
patients across different phases of the disease should illustrating ongoing challenges in the understanding of the
progressively become the rule rather than the exception inner workings of the living brain and the neurobiological
when addressing treatment effects in MRS. basis of major mental illnesses. Nonetheless, primarily as
resolution improves and the field of psychiatry continues
3. The excessive focus on BD as a categorical construct. to make progress in further refining our diagnoses and tar-
Virtually all studies to address treatment effects in get phenotypes, MRS continues to be a promising imaging
MRS among bipolar patients adopt categorical con- approach for the study of in vivo brain mechanisms and the
ceptualizations not only for the diagnosis of the illness identification of treatment effects and predictors of clinical
but also for characterization of mood states. It is well response in BD. Future research should focus on longitu-
established, however, that subsyndromical mood symp- dinal studies, which should include serial MRS scans over
toms may implicate in important psychopathological the course of the treatment, enabling the better character-
and functional burden, yet patients are often character- ization of neurochemical findings among responders and
ized as remitted when experiencing such symptoms. nonresponders, as well as the identification of pretreatment
Moreover, growing evidence suggests that a dimen- neurochemical profiles that can anticipate the response to
sional approach for BD may be more appropriate than specific pharmacological approaches.
the currently adopted categorical view. Future MRS
studies on treatment effect and predictors of response
should include a dimensional measure in the character- AC K NOW L E D G E M E N T S
ization of the sample, as well as take into consideration
the possible presence of subsyndromic symptoms when This work was partly funded by the Pat Rutherford Jr.
full criteria for an acute mood episode are not met. Chair in Psychiatry (UTHealth) and National Institutes
of Health grant 1R01MH085667
In addition, other difficulties in the identification of treat-
ment effects in MRS studies result from the relatively lack of
knowledge, in light of available evidence, in regard to which
R E F E R E NC E S
neurochemical changes should be targeted as possible indi-
cators of medication effect or response to treatment. The Agarwal N., Port, J. D., Bazzocchi, M., & Renshaw, P. F. (2010). Update
current approach is usually based on assumptions regarding on the use of MR for assessment and diagnosis of psychiatric dis-
the mechanism of action of these medications. Often these eases. Radiology, 255(1), 23–41.
Atmaca, M., Yildirim, H., Ozdemir, H., Ogur, E., & Tezcan, E. (2007).
assumptions result from preclinical information on the Hippocampal 1H MRS in patients with bipolar disorder taking
agents in question. However, some of the currently avail- valproate versus valproate plus quetiapine. Psychological Medicine,
able results suggest that many of the medication-related 37(1), 121–129.
Brambilla, P., Hatch, J. P., & Soares, J. C. (2008). Limbic changes iden-
changes detected through MRS seem to be disease-specific tified by imaging in bipolar patients. Current Psychiatry Reports,
and not necessarily associated with clinical improvement 10(6), 505–509.

M agnetic R esonance S pectroscopy F indings and T heir T reatment   I mpact   •   2 35


Brambilla, P., Glahn, D.  C., Balestrieri, M., & Soares, J.  C. (2005). Gigante, A.  D., Bond, D.  J., Lafer, B., Lam, R.  W., Young. L.  T., &
Magnetic resonance findings in bipolar disorder. Psychiatric Clinics Yatham, L.  N. (2012). Brain glutamate levels measured by mag-
of North America, 28(2), 443–467. netic resonance spectroscopy in patients with bipolar disorder:  a
Brambilla, P., Stanley, J.  A., Sassi, R.  B., Nicoletti, M.  A., Mallinger, meta-analysis. Bipolar Disorders, 14(5), 478–487.
A. G., Keshavan M. S., & Soares, J. C. (2004). 1H MRS study of Hajek, T., Bauer, M., Pfennig, A., Cullis, J., Ploch, J., O’Donovan,
dorsolateral prefrontal cortex in healthy individuals before and C., . . . Alda, M. (2012). Large positive effect of lithium on prefrontal
after lithium administration. Neuropsychopharmacology, 29(10), cortex N-acetylaspartate in patients with bipolar disorder: 2-centre
1918–1924. study. Journal of Psychiatry & Neuroscience, 37(3), 185–192.
Cecil, K. M., DelBello, M. P., Morey, R., & Strakowski, S. M. (2002). Kato, T., Fujii, K., Shioiri, T., Inubushi, T., & Takahashi, S. (1996).
Frontal lobe differences in bipolar disorder as determined by proton Lithium side effects in relation to brain lithium concentration mea-
MR spectroscopy. Bipolar Disorders, 4(6), 357–365. sured by lithium-7 magnetic resonance spectroscopy. Progress in
Chang, K., Delbello, M., Chu, W.  J., Garrett, A., Kelley, R., Neuro-Psychopharmacology & Biological Psychiatry, 20(1), 87–97.
Mills, N., . . . Strakowski, S.  M. (2012). Neurometabolite effects Kato, T., Inubushi, T., & Kato, N. (2000). Prediction of lithium
of response to quetiapine and placebo in adolescents with bipolar response by 31P-MRS in bipolar disorder. International Journal of
depression. Journal of Child and Adolescent Psychopharmacology, Neuropsychopharmacology, 3(1), 83–85.
22(4), 261–268. Kato, T., Inubushi, T., & Takahashi, S. (1994). Relationship of lith-
Clark, J. B. (1998). N-acetyl aspartate: A marker for neuronal loss or ium concentrations in the brain measured by lithium-7 magnetic
mitochondrial dysfunction. Developmental Neuroscience, 20(4–5), resonance spectroscopy to treatment response in mania. Journal of
271–276. Clinical Psychopharmacology, 14(5), 330–335.
Colla, M., Schubert, F., Bubner, M., Heidenreich, J.  O., Bajbouj, M., Kraguljac, N. V., Reid, M., White, D., Jones, R., den Hollander, J., Low-
Seifert, F., . . . Kronenberg G. (2009). Glutamate as a spectroscopic man, D., & Lahti, A. C. (2012). Neurometabolites in schizophrenia
marker of hippocampal structural plasticity is elevated in long-term and bipolar disorder—A systematic review and meta-analysis. Psy-
euthymic bipolar patients on chronic lithium therapy and corre- chiatry Research, 203(2–3), 111–125.
lates inversely with diurnal cortisol. Molecular Psychiatry, 14(7), Kushnir, T., Itzchak, Y., Valevski, A., Lask, M., Modai, I., & Navon,
696–704, 647. G. (1993). Relaxation times and concentrations of 7Li in the
Dager, S.  R., Corrigan, N.  M., Richards, T.  L., & Posse, S. (2008). brain of patients receiving lithium therapy. NMR in Biomedicine,
Research applications of magnetic resonance spectroscopy to inves- 6(1), 39–42.
tigate psychiatric disorders. Topics in Magnetic Resonance Imaging, Lan, M.  J., McLoughlin, G.  A., Griffin, J.  L., Tsang, T.  M., Huang,
19(2), 81–96. J. T., Yuan, P., . . . Bahn, S. (2009). Metabonomic analysis identifies
Davanzo, P., Thomas, M. A., Yue, K., Oshiro, T., Belin, T., Strober, M., molecular changes associated with the pathophysiology and drug
McCracken, J. (2001). Decreased anterior cingulate myo-inositol/ treatment of bipolar disorder. Molecular Psychiatry, 14(3), 269–279.
creatine spectroscopy resonance with lithium treatment in children Lyoo, I. K., & Renshaw, P. F. (2003). Magnetic spectroscopy as a tool for
with bipolar disorder. Neuropsychopharmacology, 24(4), 359–369. psychopharmacological studies. In J. C. Soares (Ed.), Brain imaging
DelBello M.  P., Cecil, K.  M., Adler, C.  M., Daniels, J.  P., & Stra- in affective disorders (pp. 245–282). New York: Marcel Dekker.
kowski, S.  M. (2006). Neurochemical effects of olanzapine in Machado-Vieira, R., Manji, H. K., & Zarate, C. A. Jr. (2009). The role
first-hospitalization manic adolescents:  A  proton magnetic reso- of lithium in the treatment of bipolar disorder: Convergent evidence
nance spectroscopy study. Neuropsychopharmacology, 31(6), for neurotrophic effects as a unifying hypothesis. Bipolar Disorders,
1264–1273. 11(Suppl. 2), 92–109.
Delbello, M. P., & Strakowski, S. M. (2004). Neurochemical predictors Maddock, R. J., & Buonocore, M. H. (2012). MR spectroscopic studies
of response to pharmacologic treatments for bipolar disorder. Cur- of the brain in psychiatric disorders. Current Topics in Behavioral
rent Psychiatry Reports, 6(6), 466–472. Neurosciences, 11, 199–251.
Drost, D. J., Riddle, W. R., Clarke, G. D., & AAPM MR Task Group McGrath, B. M., Greenshaw, A. J., McKay, R., Slupsky, C. M., & Sil-
#9. (2002). Proton magnetic resonance spectroscopy in the verstone, P. H. (2007). Unlike lithium, anticonvulsants and antide-
brain: Report of AAPM MR Task Group #9. Medical Physics, 29(9), pressants do not alter rat brain myo-inositol. NeuroReport, 18(15),
2177–2197. 1595–1598.
Forester, B. P., Finn, C. T., Berlow, Y. A., Wardrop, M., Renshaw, P. F., Merikangas, K. R., & Lamers, F. (2012). The “true” prevalence of bipo-
& Moore, C.  M. (2008). Brain lithium, N-acetyl aspartate and lar II disorder. Current Opinions in Psychiatry, 25(1), 19–23.
myo-inositol levels in older adults with bipolar disorder treated with Miklowitz, D.  J. (2011). Functional impairment, stress, and psycho-
lithium: A lithium-7 and proton magnetic resonance spectroscopy social intervention in bipolar disorder. Current Psychiatry Reports,
study. Bipolar Disorders, 10(6), 691–700. 13(6), 504–512.
Forester, B. P., Streeter, C. C., Berlow, Y. A., Tian, H., Wardrop, M., Moore, C. M., Biederman, J., Wozniak, J., Mick, E., Aleardi, M., Ward-
Finn, C T., . . Moore, C. M. (2009). Brain lithium levels and effects rop, M., . . . Renshaw, P. F. (2007). Mania, glutamate/glutamine and
on cognition and mood in geriatric bipolar disorder:  A  lithium-7 risperidone in pediatric bipolar disorder: A proton magnetic reso-
magnetic resonance spectroscopy study. American Journal of Geri- nance spectroscopy study of the anterior cingulate cortex. Journal of
atric Psychiatry, 17(1), 13–23. Affective Disorders, 99(1–3), 19–25.
Friedman, S.  D., Dager, S.  R., Parow, A., Hirashima, F., Demopulos, Moore, G. J., Bebchuk, J. M., Hasanat, K., Chen, G., Seraji-Bozorgzad,
C., . . . Renshaw P.  F. (2004). Lithium and valproic acid treatment N., Wilds, I.  B.,  . 
. 
. 
Manji, H.  K. (2000). Lithium increases
effects on brain chemistry in bipolar disorder. Biological Psychiatry, N-acetyl-aspartate in the human brain: in vivo evidence in support
56(5), 340–348. of bcl-2’s neurotrophic effects? Biological Psychiatry; 48(1), 1–8.
Frye, M.  A., Watzl, J., Banakar, S., O’Neill, J., Mintz, J., Davanzo, Moore, G. J., Bebchuk, J. M., Parrish, J. K., Faulk, M. W., Arfken, C. L.,
P., . . . Thomas, M.  A. (2007). Increased anterior cingulate/medial Strahl-Bevacqua, J., & Manji, H. K. (1999). Temporal dissociation
prefrontal cortical glutamate and creatine in bipolar depression. between lithium-induced changes in frontal lobe myo-inositol and
Neuropsychopharmacology, 32(12), 2490–2499. clinical response in manic-depressive illness. The American Journal
Gallelli, K. A., Wagner, C. M., Karchemskiy, A., Howe, M., Spielman, of Psychiatry, 156(12), 1902–1908.
D., Reiss, A., & Chang, K.  D. (2005). N-acetylaspartate levels in Murashita, J., Kato, T., Shioiri, T., Inubushi, T., & Kato, N. (2000).
bipolar offspring with and at high-risk for bipolar disorder. Bipolar Altered brain energy metabolism in lithium-resistant bipolar
Disorders, 7(6), 589–597.

2 3 6   •   S E C T I O N I I :   I N S I G H T S O N PAT H O PH Y S I O L O G Y A N D F U T U R E T R E AT M E N T S
disorder detected by photic stimulated 31P-MR spectroscopy. Psy- Silverstone, P.  H., Wu, R.  H., O’Donnell, T., Ulrich, M., Asghar,
chological Medicine, 30(1), 107–115. S.  J., & Hanstock, C.  C. (2002). Chronic treatment with both
Nery, F. G., Stanley, J. A., Chen, H. H., Hatch, J. P., Nicoletti, M. A., lithium and sodium valproate may normalize phosphoinositol
Monkul, E. S., . . . Soares, J. C. (2009). Normal metabolite levels in cycle activity in bipolar patients. Human Psychopharmacology,
the left dorsolateral prefrontal cortex of unmedicated major depres- 17(7), 321–327.
sive disorder patients: A single voxel (1)H spectroscopy study. Psy- Silverstone, P. H., Wu, R. H., O’Donnell, T., Ulrich, M., Asghar, S. J.,
chiatry Research, 174(3), 177–183. & Hanstock, C.  C. (2003). Chronic treatment with lithium, but
O’Donnell, T., Rotzinger, S., Nakashima, T.  T., Hanstock, C.  C., not sodium valproate, increases cortical N-acetyl-aspartate concen-
Ulrich, M., & Silverstone, P.  H. (2003). Chronic lithium and trations in euthymic bipolar patients. International Clinical Psycho-
sodium valproate both decrease the concentration of myoinosi- pharmacology, 18(2), 73–79.
tol and increase the concentration of inositol monophosphates in Soares, J. C. (2002). Can brain-imaging studies provide a “mood stabi-
rat brain. European Journal of Neuropsychopharmacology, 13(3), lizer signature?” Molecular Psychiatry, 7(Suppl. 1), S64–S70.
199–207. Soares, J.  C., Boada, F., & Keshavan, M.  S. (2000). Brain lithium
Ongür, D., Jensen, J. E., Prescot, A. P., Stork, C., Lundy, M., Cohen, measurements with (7)Li magnetic resonance spectroscopy
B.  M., & Renshaw, P.  F. (2008). Abnormal glutamatergic neuro- (MRS):  A  literature review. European Neuropsychopharmacology,
transmission and neuronal-glial interactions in acute mania. Bio- 10(3), 151–158.
logical Psychiatry, 64(8), 718–726. Soares, J.  C., & Mann, J.  J. (1997). The anatomy of mood
Patel, N. C., Cecil, K. M., Strakowski, S. M., Adler, C. M., & DelBello, disorders—review of structural neuroimaging studies. Biological
M.  P. (2008). Neurochemical alterations in adolescent bipolar Psychiatry, 41(1), 86–106.
depression: a proton magnetic resonance spectroscopy pilot study of Stanley, J. A. (2002). In vivo magnetic resonance spectroscopy and its
the prefrontal cortex. Journal of Child and Adolescent Psychophar- application to neuropsychiatric disorders. Canadian Journal of Psy-
macology, 18(6), 623–627. chiatry, 47(4), 315–326.
Perlis, R. H., Ostacher, M. J., Patel, J. K., Marangell, L. B., Zhang, H., Strawn, J. R., Patel, N. C., Chu, W. J., Lee, J. H., Adler, C. M., Kim,
Wisniewski, S. R., . . . Thase, M. E. (2006). Predictors of recurrence M. J., . . . DelBello, M. P. (2012). Glutamatergic effects of divalproex
in bipolar disorder: Primary outcomes from the Systematic Treat- in adolescents with mania: A proton magnetic resonance spectros-
ment Enhancement Program for Bipolar Disorder (STEP-BD). The copy study. Journal of the American Academy of Child & Adolescent
American Journal of Psychiatry, 163, 217–224. Psychiatry, 51(6), 642–651.
Phillips, M.  L., Travis, M.  J., Fagiolini, A., & Kupfer, D.  J. (2008). Vieta, E., Reinares, M., & Rosa, A. R. (2011). Staging bipolar disorder.
Medication effects in neuroimaging studies of bipolar disorder. The Neurotoxicity Research, 19(2), 279–285.
American Journal of Psychiatry, 165(3), 313–320. Waddell, K.  W., Zanjanipour, P., Pradhan, S., Xu, L., Welch, E.  B.,
Sachs, G. S., Renshaw, P. F., Lafer, B., Stoll, A. L., Guimarães, A. R., Joers, J. M., . . . Gore, J. C. (2011). Anterior cingulate and cerebellar
Rosenbaum, J. F., Gonzalez, R. G. (1995). Variability of brain lith- GABA and Glu correlations measured by ¹H J-difference spectros-
ium levels during maintenance treatment: a magnetic resonance copy. Magnetic Resonance Imaging, 29(1), 19–24.
spectroscopy study. Biological Psychiatry, 1, 38(7), 422–428. Wang, P. W., Dieckmann, N., Sailasuta, N., Adalsteinsson, E., Spiel-
Sanches, M., & Soares, J.  C. (2011). New drugs for bipolar disorder. man, D. & Ketter, T. A. (2002). 3 Tesla 1H-magnetic resonance spec-
Current Psychiatry Reports, 13(6), 513–521 troscopic measurements of prefrontal cortical gamma-aminobutyric
Sanches, M., Brambilla, P., & Soares, J. C. (2006). The role of imaging acid (GABA) levels in bipolar disorder patients and healthy volun-
techniques in the development and clinical use of psychotropic drugs. teers. Paper presented at the 57th Annual Convention and Scientific
In F. López-Muñoz & C. Alamo (Eds.), Historia de la psicofarmacolo- Program of the Society of Biological Psychiatry. Philadelphia.
gia (pp. 479–494). Barcelona: Editorial Medica Panamericana. Wu, R. H., O’Donnell, T., Ulrich, M., Asghar, S. J., Hanstock, C. C.,
Sanches, M., Newberg, A. R., & Soares, J. C. (2010). Emerging drugs & Silverstone, P. H. (2004). Brain choline concentrations may not
for bipolar disorder. Expert Opinion on Emerging Drugs, 15(3), be altered in euthymic bipolar disorder patients chronically treated
453–466. with either lithium or sodium valproate. Annals of General Psychia-
Sassi, R., & Soares, J. C. (2003). Brain imaging methods in neuropsy- try, 3(1), 13.
chiatry. In J. C. Soares (Ed.), Brain imaging in affective disorders (pp. Yildiz, A., Demopulos, C. M., Moore, C. M., Renshaw, P. F., & Sachs,
1–17). New York: Marcel Dekker. G. S. (2001). Effect of lithium on phosphoinositide metabolism in
Shahana, N., Delbello, M., Chu, W.  J., Jarvis, K., Fleck, D., Welge, human brain: A proton-decoupled 31P MRS study. Biological Psy-
J., 
. 
. 
. 
Adler, C. (2011). Neurochemical alteration in the cau- chiatry, 50(1), 3–7.
date: Implications for the pathophysiology of bipolar disorder. Psy- Yildiz, A., Sachs, G.  S., Dorer, D.  J., & Renshaw, P.  F. (2001). 31P
chiatry Research, 193(2), 107–112. Nuclear magnetic resonance spectroscopy findings in bipolar ill-
Sharma, R., Venkatasubramanian, P.  N., Bárány, M., & Davis, J.  M. ness: A meta-analysis. Psychiatry Research, 106(3), 181–191.
(1992). Proton magnetic resonance spectroscopy of the brain Yildiz-Yesiloglu, A., & Ankerst, D.  P. (2006). Neurochemical
in schizophrenic and affective patients. Schizophrenia Research, alterations of the brain in bipolar disorder and their implica-
8(1), 43–49. tions for pathophysiology:  A  systematic review of the in vivo
Shibuya-Tayoshi, S., Tayoshi, S., Sumitani, S., Ueno, S., Harada, proton magnetic resonance spectroscopy findings. Progress
M., & Ohmori, T. (2008). Lithium effects on brain glutama- in Neuro-Psychopharmacology & Biological Psychiatry, 30(6),
tergic and GABAergic systems of healthy volunteers as mea- 969–995.
sured by proton magnetic resonance spectroscopy. Progress Zimmerman, M. (2010). Problems diagnosing bipolar disorder in
in Neuro-Psychopharmacology & Biological Psychiatry, 32(1), clinical practice. Expert Review of Neurotherapeutics, 10(7),
249–256. 1019–1021.
Silverstone, P. H., McGrath, B. M., & Kim, H. (2005). Bipolar disorder
and myo-inositol: A review of the magnetic resonance spectroscopy
findings. Bipolar Disorders, 7(1), 1–10.

M agnetic R esonance S pectroscopy F indings and T heir T reatment   I mpact   •   2 3 7

You might also like