Ann. N.Y. Acad. Sci. ISSN 0077-8923

A N N A L S O F T H E N E W Y O R K A C A D E M Y O F SC I E N C E S Issue: Translational Neuroscience in Psychiatry

In pursuit of neuroimaging biomarkers to guide treatment selection in major depressive disorder: a review of the literature Marc S. Lener and Dan V. Iosifescu Department of Psychiatry, Icahn School of Medicine at Mount Sinai, New York, New York Address for correspondence: Marc S. Lener, M.D., Department of Psychiatry, Icahn School of Medicine at Mount Sinai, One Gustave L. Levy Place, New York, NY 10029. [email protected]

Over the last few decades, neuroimaging techniques have advanced the identification of structural, functional, and neurochemical brain abnormalities that are associated with the increased risk, clinical course, and treatment outcomes of major depressive disorder (MDD). This paper reviews specific neuroimaging abnormalities that, on the basis of early studies, may discriminate between MDD patients who do or do not respond to current therapeutic modalities, such as antidepressants, cognitive behavioral therapy, or novel therapies. Differences in gray matter volume, white matter coherence, brain activity via structural and functional magnetic resonance imaging techniques, and concentrations of specific brain metabolites (as measured with magnetic resonance spectroscopy), are potential biomarkers discussed in this review. Given the heterogeneity of MDD, larger, multisite studies with increased statistical power will be needed to identify more precise imaging biomarkers of treatment response in MDD. Keywords: depression; major depressive disorder; treatment-resistant depression; neuroimaging; voxel based; structural; morphometry; white matter; functional; resting state; SPECT

Introduction Among patients who develop major depressive disorder (MDD), 15–30% will develop a chronic, unremitting depression despite trials of multiple antidepressant medications.1 This poses a significant clinical challenge given that most common antidepressant treatments (such as selective serotonin reuptake inhibitors (SSRI)) and psychotherapies (such as cognitive behavioral therapy (CBT)) are slow to take effect, requiring 6–8 weeks or longer for clinical benefit even among patients who ultimately improve.2 As a consequence, patients frequently suffer from ongoing emotional distress, financial and social difficulties resulting from impaired functioning, negative impact on comorbid medical illness, and suicidality.3 Second, approximately 12–30% of individuals with MDD who fail to respond despite two or more antidepressant treatments during the depressive episode (defined as treatment-resistant depression (TRD)) have a disproportionately high rate of morbidity.4–8 Unfortunately, the absence of a reliable clinical measure to

guide the diagnosis or the next-step treatment selection in MDD (or TRD) significantly limits the ability of clinicians to address these severe problems in a more timely manner. Not surprisingly, the discovery of novel biological targets and the development of more effective and rapid treatment paradigms for MDD has become a priority for the field and for the National Institute of Mental Health (NIMH).9,10 Psychiatric research has struggled to identify biological markers (also known as biomarkers) to be used as reliable and valid measures of psychiatric illness and predictors of treatment outcome. In particular, biomarker identification for MDD has become an important focus of the NIMH and pharmaceutical industry9 and advances in neuroimaging techniques and other neurobiological investigations in MDD. In a recent study using fluorodeoxyglucose (FDG) positron emission tomography (PET), a specific metabolic signature in the insula was found to discriminate between responders to CBT and responders to antidepressants.11 If replicated, this result may become a major finding and a potential doi: 10.1111/nyas.12759

50

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

unique neuroimaging biomarker. However, in this study there were too few patients who completed and survived analysis to account for the vast clinical and neurobiological heterogeneity. A more likely scenario is that neuroimaging biomarkers will enhance detection of vulnerable, protective, and treatment-responsive groups in conjunction with clinical measures as well as genetic and epigenetic biomarkers. It is also possible that biomarkers and specific neuroimaging-defined mechanisms will not describe MDD as a whole, but rather subtypes of MDD or even syndromes that have been defined within the NIMH-proposed Research Domain Criteria (RDoC).12 While a variety of biomarkers have been studied in an effort to (1) enhance the precision of the diagnosis of MDD and to (2) increase the ability to select efficacious next-step treatments, this review focuses primarily on the second category (i.e., biomarkers of treatment efficacy). We will present data from selected neuroimaging studies in an effort to summarize biomarker investigations, provide insights into what has been learned from these studies, and discuss their aid in focusing subsequent studies, particularly in the context of the paradigm represented by the NIMH RDoC. We present data highlighting converging results in the same brain regions and/or brain circuits/networks implicated in MDD, emphasizing (where applicable) the data pertaining to TRD. We chose this approach as evidence suggests that a convergence of multiple imaging and nonimaging techniques may be necessary to identify and confirm specific neurocircuits that underlie this disorder. Methods We used PubMed to search MEDLINE for articles published between January 2000 and July 2014, with an emphasis on papers published after January 2011 by using terms related to depression, MDD, or TRD. We then paired these core terms with secondary search terms—neuroimaging, voxel based, morphometry, structural, white matter, functional, resting state, PET, single-photon emission computed tomography (SPECT), spectroscopy, magnetic resonance spectroscopy (MRS)—and excluded studies that did not include an MDD group. We primarily focused on patients between the ages of 18 and

Neuroimaging biomarkers in MDD

65 years, with exception to three studies of adolescents (12–19 years old) with MDD (one longitudinal study and two cross-sectional studies that were compared with findings from adult MDD), one study of adults with MDD (>18 years old), and four crosssectional studies of the elderly (>65 years old) with MDD. The reference lists of selected studies were additionally searched manually for relevant studies. Where applicable, reviews and meta-analyses were used to summarize findings earlier than January 2011, particularly for studies identifying neuroimaging biomarkers of illness (which was a secondary focus of this manuscript). A more in-depth review of the literature was performed for studies identifying neuroimaging biomarkers of treatment response. Two researchers screened each abstract of the retrieved articles. The articles were included if the abstract referred to primary data collection as a part of the design and if end points of the study were pertinent to clinical outcomes. Brain regional structural abnormalities implicated in MDD An accumulating body of literature investigating patients with MDD has demonstrated a diffuse pattern of gray matter volume abnormalities. Metaanalyses of voxel-based morphometry (VBM) studies investigating patients with MDD as compared to healthy controls (HC) most robustly and notably implicate a significant gray matter volume reduction within the prefrontal cortex (PFC) (including the dorsolateral PFC (dlPFC), medial PFC (mPFC), ventrolateral PFC (vlPFC), and orbitofrontal cortex (OFC)) and limbic areas (such as the hippocampus, amygdala, and the anterior cingulate cortex (ACC)).13–18 It is posited that these regional volumetric changes result in abnormal functioning within a larger network of neural circuits.19 Although the exact neuropathology associated with gray matter volumetric reductions found in patients with MDD is not well understood, postmortem neuropathological studies have shown evidence of a decrement in glial cell quantity, neuron cell size, synaptic density, and neuronal cell quantity in late stages of the illness.20–24 Given that the majority of structural imaging investigations in MDD are cross-sectional studies, it is not known whether these abnormalities confer susceptibility

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

51

Neuroimaging biomarkers in MDD

Lener & Iosifescu

to the development of MDD, represent a pathological consequence or compensatory change after depressive episodes, or are a direct result of medication. In a VBM study comparing MDD patients with HC groups (with and without a family history of MDD), the HC group with a positive family history demonstrated a significantly smaller hippocampal (P = 0.042; k = 597; T = 3.20) and dlPFC (P = 0.031; k = 128; T = 3.32) gray matter volume compared with the HC group without a family history.25 Although the at-risk HC group surprisingly demonstrated a significantly greater reduction in right hippocampal volume compared with patients with MDD (P = 0.017; k = 1156; T = 3.54), the MDD group exhibited smaller volumes in the ACC, dorsal mPFC (dmPFC), and the basal ganglia (all P < 0.049) compared with both HC groups, suggesting that illness vulnerability might be associated with volumetric reductions in the dmPFC and hippocampus. Of note, the hippocampus is a particularly sensitive brain region in patients with MDD; hippocampal volumetric decreases have been reported in association with the duration of illness and number of depressive episodes.26,27 Such hippocampal volumetric reductions in MDD patients have been shown to associate with poorer response to antidepressant treatment,28–31 are greater in magnitude in unmedicated than in medicated MDD patients,32 and may be reversed in patients who receive electroconvulsive therapy (ECT).33 In a longitudinal study of adolescent patients at risk for psychiatric illness, volumetric changes in the hippocampus (e.g., attenuated growth of the hippocampus; F = 6.79, P = 0.011), amygdala (F = 7.61; P = 0.007), and putamen (F = 5.69; P = 0.020) were associated with depression onset, suggesting that developmental abnormalities of the hippocampus and related limbic areas may confer risk of developing MDD.34 To date, only a handful of studies have investigated structural brain abnormalities in association with treatment response in patients with MDD (Table S1). One approach consists of cross-sectional comparisons between MDD patients with depressive episodes in remission and MDD patients with highly recurrent depressive episodes (e.g., Refs. 28, 31, and 35) or MDD patients currently depressed.36–38 Such studies can help identify markers of resilience (remitted depression) as well as markers of unfavorable clinical course (as in 52

TRD). In one study, patients with remitted (as opposed to recurrent) depression demonstrated an increased volume in the subgenual PFC (F = 4.32; P = 0.02) as compared to HCs, suggesting that this regional volumetric change may serve as a dual marker for susceptibility and antidepressant treatment response.35 Regarding anterior insular cortical volume in patients with MDD, there have been conflicting findings with respect to volumetric alterations (increased or decreased) and discriminating patients with recurrent and remitted depression.36,38,39 However, in one study that used both a cross-sectional and longitudinal study design, a volume reduction in the insula was associated with melancholic depression and slower treatment recovery (r = –0.39; P = 0.001).40 In a longitudinal structural magnetic resonance imaging (MRI) study conducted over a 1-year follow-up period, MDD patients who had smaller hippocampal volumes were more treatment refractory and had more relapses than MDD patients whose depression had remitted.29 This was replicated by the same group in a 3-year follow-up study of a different MDD patient cohort.30 In other prospective longitudinal studies, antidepressant treatment response has been associated with larger hippocampal volume after improvements of depression scores over a 5-week treatment period,41 larger ACC39 and hippocampal volumes42 over an 8-week treatment period, and a larger OFC volume in association with a 6-month remission during a 12-month treatment period,43 whereas a nonresponse over a 6-week antidepressant treatment period has been associated with reduced gray matter volume in the left dlPFC.44 The timeframe is important, since volumetric changes over short intervals (5 weeks) may have different biological substrates than changes occurring over 6–12 months. Moreover, short-term volumetric reductions are more likely to represent changes in dendritic arborization, while long-term volumetric changes are more likely to represent changes in underlying cellular populations. In a recent meta-analysis of neuroimaging predictors of treatment response, a reduction in hippocampal volume (standard mean difference: –0.81, 95% CI: –1.36–0.26) was a significant predictor of a lower likelihood of improvement and poorer treatment response.45 Taken together, increased frontal (OFC, subgenual PFC) and hippocampal volumes may

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

portend a greater probability of treatment response, whereas decreased frontal (dlPFC), hippocampal, and insular volumes may predict a lower probability of treatment response with antidepressant therapy. In a high-field MRI study of patients with remitted and nonremitted depression compared with HCs, a machine-learning approach was used to make an MDD diagnosis, with an accuracy of between 67.39% (P = 0.01) and 76.09% (P < 0.001), and to distinguish between the two subgroups of MDD, with an accuracy of 69.57% (P = 0.006).46 This further supports that an abnormal network of brain circuits underlies MDD and that a pattern of volumetric abnormalities particular to MDD can be identified using machine-learning algorithms. Structural connectivity in MDD and TRD The investigation of white matter abnormalities in patients with MDD was initially conducted on the basis of global measures of variations in MR signal intensity (“white matter hyperintensities”) and demonstrated an increased incidence of total brain white matter lesions (WMLs) in elderly MDD subjects47,48 and in some,49 but not all, younger MDD cohorts.50,51 Recognition of the increased prevalence of brain WMLs in MDD has led investigators to describe vascular depression, a subtype of MDD characterized by the presence of cerebrovascular disease (demonstrated on neuroimaging scans by brain WMLs) and poor connectivity between limbic and cortical regions involved in emotional regulation.47,52 In some studies (Table S1) the presence of brain WMLs in MDD subjects was associated with lower rates of response to antidepressant treatment, compared with MDD subjects with no WMLs;51,53 higher rates of irritability and anger attacks;54 as well as higher rates of relapse in long-term follow-up.55 The studies using global measures of WMLs did not inform, however, on the specific circuits affected in MDD; this lack of specificity may also explain the negative results. These limitations were, in part, overcome by the advent of diffusion tensor imaging (DTI), an imaging technique that utilizes properties of water diffusion to measure the integrity of brain tissue. A growing body of DTI studies in MDD has been subject to extensive reviews56 and meta-analyses57 suggesting that microstructural changes in cortical– subcortical white matter in patients with MDD may give rise to a “disconnection syndrome” within

Neuroimaging biomarkers in MDD

key frontolimbic neural circuits and contribute to the emergence of depression. Fractional anisotropy (FA), a measure of white matter coherence (and an indirect measure of the integrity of white matter circuits), is reduced as a result of impairments in the myelination of axons or axonal membranes and/or of a decreased density of axons. FA reductions have been found in frontal regions (e.g., superior and middle frontal white matter, and body/genu of the corpus callosum),56–61 white matter tracts interconnecting limbic areas (e.g., anterior cingulum, fornix, and uncinate fasciculus),58,62,63 and thalamocortical white matter relays57,64 primarily in elderly patients, but also in middle-aged and young patients with recurrent MDD. In addition to prior work showing gray matter volumetric changes within regions that are interconnected by these abnormal white matter tracts,13,25,65 this body of DTI studies is consistent with a dysfunctional neural network66 in depressed patients. However, there are multiple other studies63,67–70 showing no significant differences in FA or other measures of white matter coherence between MDD and matched HCs. It is likely that subjects with significant white matter abnormalities may represent a subgroup of MDD subjects (more prevalent in TRD, as shown below) and that studies of large numbers of treatment-naive MDD subjects may result in a dilution effect. Although the presence of heterogeneity within patients diagnosed with MDD was posited nearly a half century ago,71 these results highlight the importance of revisiting clinical heterogeneity to identify subgroups of patients who may have different clinical, genetic, neurobiological, and treatment effectiveness profiles.72–79 There are fewer structural imaging studies that investigate patients with TRD than MDD.80–86 Although similar frontolimbic abnormalities are seen in patients with TRD and in those with treatment-responsive MDD,83,84 lesions of specific white matter tracts appear to be more severe in TRD patients.85 WMLs in circuits connecting frontostriatal80 and limbic areas appear to be more severe in TRD, particularly in the hippocampus,81,82 and may aid in distinguishing patients who are treatment resistant from those who are treatment responsive. In a whole-brain DTI study of patients with TRD, remitted recurrent MDD, first-episode MDD, and HCs, it was found that patients with TRD had significant FA reductions in white matter within the ventromedial PFC (vmPFC) as compared to the patients

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

53

Neuroimaging biomarkers in MDD

Lener & Iosifescu

with remitted recurrent MDD (P = 0.03).86 In the analysis of prefrontal cortical white matter, Hamilton Depression Rating Scale (HDRS or HAM-D) scores and the number of previous episodes were significant predictors of decreased FA values, accounting for 12% of the variance. Therefore, a frontolimbic disconnection, as demonstrated by reductions in FA within select frontal, striatal, and limbic regions, may be most robustly associated with patients who have unremitting or treatment-resistant forms of MDD. Several studies have investigated white matter abnormalities as a predictive biomarker of the efficacy of diverse treatments, including antidepressants,87 deep-brain stimulation (DBS),88,89 and transcranial magnetic stimulation (TMS).90 In a DTI study of a large sample (n = 74) of elderly MDD patients over 60 years of age, higher FA in the superior frontal gyri and bilateral ACC correlated with nonremission over a 12-week course of sertraline, suggesting that these regions may be involved in the mechanism of antidepressant response and that their abnormal FA can predict both the risk of depression and SSRI treatment response.87 In contrast to this finding, in another large but younger sample of MDD patients, altered connectivity for the cingulum part of the cingulate and stria terminalis tracts significantly predicted remission with 62% accuracy using a cross-validated logistic regression model.91 This suggests that dysconnectivity in MDD may exist within a specific combination of tracts (as opposed to just one), allowing for more accurate treatment prediction. Two overlapping white matter tracts that were previously shown to be effective targets of DBS (subcallosal cingulum and the anterior limb of the internal capsule) were examined to determine shared white matter connections. In a probabilistic connectivity map, Gutman et al. found shared regions within the frontal pole, medial temporal lobe, nucleus accumbens, dorsal thalamus, and hypothalamus,88 suggesting that the efficacy of DBS may rely on correcting central dysfunctional circuits. In an exploratory DTI study of 30 younger patients with TRD enrolled in a double-blind, randomized repetitive TMS (rTMS) study, baseline FA reductions in the left middle frontal gyrus predicted an improvement of depressive symptoms in the active (rTMS) group but not in the sham group.90 FA values were later increased after rTMS 54

treatment. This suggests that rTMS may give rise to a corrective process in specific white matter tracts that can be clinically tracked through a pre- and postprocedure MRI. Future DTI studies will be necessary to replicate and extend these initial results. Functional dysconnectivity in MDD Functional brain imaging techniques such as perfusion SPECT and functional MRI (fMRI) allow for interrogation of functional brain activity in patients with psychiatric disorders. In this review, we focus on studies using fMRI in MDD patients compared with HCs, and within MDD patients as a discriminator of treatment response. Functional MRI is an MRI-based method that measures the blood oxygen–level dependent response, a proxy for regional neural activity. In studies of depression, fMRI can be used in research paradigms designed to test abnormal brain activation patterns in association with a specific cognitive and/or emotionally salient task or can be used to test abnormalities of functional connectivity unrelated to a specific task, known as “resting-state” brain activity. Previously published reviews and meta-analyses of task-related and resting-state fMRI studies have found abnormalities in frontal (dlPFC, vlPFC, and mPFC), limbic (ACC, amygdala), thalamostriatal (thalamus, globus pallidus, putamen, and caudate), and insular cortical brain activation across domains of emotion processing, cognitive control, affective cognition, reward processing, and resting-state functional connectivity.92–95 Owing to the centrality of emotional dysregulation found in depressed patients, the majority of task-related fMRI studies in depression have targeted emotion-processing brain networks through the induction of both implicit (e.g., automatic) or explicit (e.g., effortful) behavioral responses to affect-laden stimuli. The consensus among studies has shown that, relative to HCs, patients with MDD show hyperactivation within limbic regions and hypoactivation of cortical systems in response to emotionally salient stimuli,96–101 a pattern that is, in part, reversed with antidepressant treatment.102 This altered activation in top-down (or high-level cognitive interpretation of emotion) and bottom-up (or encoding of affect-laden stimuli to form a perception) prefrontal–subcortical circuitry has been suggested to underlie the failure to regulate mood in patients with a range of psychiatric disorders,

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

such as MDD,103–105 bipolar disorder,106 anxiety disorders,107,108 posttraumatic stress disorder,109 and borderline personality disorder.110 This relative lack of specificity is unsurprising given that each patient group exhibits a varying form of emotion dysregulation. What may underlie the differences between these groups beyond the environmental and contextual aspects are the differences in the degree of recruitment of cognitive regulatory networks,108,110,111 differential involvement of reward networks,109 or a biologically driven fluctuant pattern as seen with bipolar disorder.106 Two other important task-related fMRI paradigms that have been used to investigate functional brain networks that underlie motivational anhedonia and ruminative brooding in patients with MDD include those that examine reward processing and self-referential processing, respectively. Reward-processing fMRI paradigms detect underlying reward network dysfunction to explain motivational and consummatory anhedonia, hallmark features of the symptom cluster in depression. In response to positively valent images (i.e., images that evoke pleasant emotion, such as images of food or sexually arousing images of people), MDD patients showed decreased activation in the mPFC and increased activation in the inferior frontal cortex, ACC, thalamus, putamen, and insula.112 In a study investigating neural responses to positive- and negative-valence stimuli in a group of depression patients compared with HCs, the vmPFC appeared to demonstrate a double-dissociation pattern of activity successfully delineating the HC from the MDD group.113 In a study using a monetary incentive delay task, patients with MDD showed significantly weaker responses to gains in the left nucleus accumbens and the caudate bilaterally. Furthermore, volumetric reductions in the caudate were associated with anhedonic symptoms and depression, suggesting a major role for basal ganglia and nucleus accumbens activity in mediating consummatory anhedonia.114 In patients with MDD (as well as those with anxiety disorders), rumination is persistent, repetitive, and self-critical in nature, characteristic of a depressive episode,115 and may represent a failure to regulate affect or emotional distance when recalling negative autobiographical memories.116 Studies that have used various strategies to induce rumination

Neuroimaging biomarkers in MDD

demonstrate hyperactivity in the subgenual ACC (sgACC) and dmPFC116–118 and hypoactivity in the dorsomedial thalamus and ventral striatum.119 In a study conducted by Yoshimura et al., activity in the rostral ACC was functionally connected to the mPFC and the amygdala during rumination and mediated the correlation between mPFC activity and depressive symptoms, suggesting that rostral ACC activity may lie at the centerpiece of limbic system dysfunction in depression.118 Finally, in a study that required subjects to passively examine negative pictures and actively reappraise them, MDD patients, but not HCs, exhibited a failure to reduce activity in the vmPFC, ACC, and other regions while both looking at negative pictures and reappraising them, demonstrating a failure to normally downregulate activity broadly within the default-mode network.120 Studies using fMRI to predict treatment response in patients with MDD have used task-related, resting-state, and combined approaches in search of a signature of functional brain activation unique to responders. Although not supported by all studies,121–124 the most consistently replicated finding in task-related fMRI studies of MDD patients is that of increased ACC activation in response to negative stimuli as a marker of improvement in depressive symptoms and a predictor of an antidepressant response to venlafaxine,125 fluoxetine,39 and a variety of other antidepressant medications126,127 (Table S1). This finding reinforces previous and influential work demonstrating a link between ACC hypermetabolism and antidepressant treatment response in patients with MDD.128 In a recent metaanalysis of 23 fMRI studies investigating neural activity before and after antidepressant treatment, 19 of the studies demonstrated that increased neural activity in the rostral ACC robustly (Cohen’s d value: 0.918) predicted better antidepressant treatment response across studies and across different interventions, such as SSRIs, atypical antidepressants, ketamine, sleep deprivation, and rTMS.129 Moreover, ACC hyperactivation has been shown to be a potential biomarker of an antidepressant response to ketamine using other techniques measuring neural activity, such as magnetoencephalography (MEG). In a recent MEG study (n = 11 MDD and 11 HCs), presentation of fearful faces resulted in robust increases in pretreatment ACC

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

55

Neuroimaging biomarkers in MDD

Lener & Iosifescu

activity that positively correlated with a subsequent rapid antidepressant response to ketamine.130 In a subsequent MEG study by the same group using a working memory task to engage the pregenual ACC, patients who showed the least engagement of the pregenual ACC in response to increased working memory load showed the greatest symptomatic improvement within 4 h of ketamine administration (r = 0.82; P = 0.0002; false discovery rate 75% correct classification of response, >70% correct classification of remission).133 In 18 F-FDG PET studies, responses to antidepressants were associated with hypoactivity in the midbrain,134 whereas nonresponse to CBT has been associated with pretreatment hypermetabolism in a region located between the pregenual and subgenual cingulate cortices in a small group of MDD patients receiving either CBT or venlafaxine.135 Although studies have remained inconsistent, it has been proposed that, while ACC hyperactivity predicts positive response to antidepressant medication, ACC hypoactivity may predict response to CBT.45,136,137 In contrast to these findings, a recent FDG PET study found that hypometabolism in the right anterior insula was associated with remission after CBT and poor response to escitalopram, while insula hypermetabolism was associated with remission after escitalopram and poor response to CBT.11 This has been the most compelling argument to date in support of a predictive neuroimaging biomarker with the ability to discriminate by treatment type. Given the small sample size of this study and the discordant results with the previous literature, further replication studies with larger sample sizes would be necessary to confirm the predictive abilities of this biomarker. 56

Neuroimaging studies of biochemical abnormalities in MDD In contrast to PET, which requires a radiotracer tagged specifically to a known molecular target, MRS is an imaging technique that measures the abundance of specific neurochemicals using their specific magnetic resonance signatures. MRS studies in psychiatric disorders most often involve proton (1 H) and phosphorus (31 P) spectroscopy. The former (1 H-MRS) is an important method to identify biochemical regional brain abnormalities by measuring concentrations of certain metabolites associated with neuronal death (e.g., creatinine (Cr), N-acetyl aspartate (NAA), myo-inositol (MI), and choline (CHO)); 1 H-MRS is also used to identify abnormalities in neurotransmitter systems implicated in the pathophysiology of MDD (e.g., gamma-aminobutyric acid (GABA), glutamate (Glu), glutamine (Gln), or the combined Glu–Gln levels (also known as Glx)).138 Studies using 1 H-MRS in patients with MDD have demonstrated higher CHO and CHO/Cr ratios in the basal ganglia, lower Glx levels in the frontal lobes,139–142 and reductions of Glx/Cr and Gln/Cr ratios in the hippocampus.143,144 In a study of MDD patients, PFC concentrations of NAA/Cr were lower in moderately depressed patients when compared to mildly depressed patients and HCs,145 suggesting that reductions of NAA/Cr in the PFC may be larger in more severe depressive episodes, potentially explaining studies that have shown the contrary.139 Finally, reductions in NAA/Cr have been shown to renormalize in association with antidepressant treatment within the basal ganglia146 and mPFC,147 while an increase in CHO compounds were found in association with ECT in the hippocampus after five ECT treatments;148 hippocampal CHO appears to eventually decrease to pretreatment levels after long-term (20 months) follow-up.143 These studies suggest that current treatments for depression may be associated with increased neuronal trophicity and lipid membrane turnover (neuroplasticity). Clearly, further studies will be necessary to determine whether these metabolite changes can predict treatment response. Studies using 1 H-MRS can also detect levels of neurotransmitters associated with glutamate metabolism (such as GABA, Glu, and Gln). GABA concentrations in depressed patients were

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

reportedly lower in the occipital cortex149 and more pronounced in a melancholic subtype of depression.150 Occipital GABA levels appeared to increase after successful antidepressant treatment with SSRIs151 or ECT.152 The methodological limitations in these earlier studies (i.e., surface coil used), which restricted GABA measurements to the occipital cortex, were surpassed in subsequent studies of unmedicated MDD patients showing reductions of GABA in the ACC and associations with anhedonia in an adolescent group,153 and reduced GABA in the dorsomedial and dorsal anterolateral PFC154 in adults. Hasler et al. also found reduced Glx levels in MDD patients in the same brain regions as well as within the vmPFC, suggesting that regional alterations of inhibitory/excitatory neurotransmission may exist in MDD patients. Several studies have demonstrated reductions in Glu in the PFC and ACC,155,156 coinciding with a growing number of clinical studies that show a rapid antidepressant response to ketamine, an NMDA glutamate receptor antagonist.157–159 It is yet to be determined whether specific levels of Glu or an excitatory/inhibitory neurotransmitter system imbalance in the PFC and/or ACC can help predict the subset of patients who may be particularly responsive to ketamine or more established antidepressant treatments.160 An emerging and developing hypothesis161,162 that a relationship exists between abnormal inflammatory states and the development of MDD and other stress-related disorders has prompted studies of MDD patients to examine peripheral inflammatory markers and in vivo neuroimaging markers of oxidative stress, specifically of glutathione (GSH), the most prevalent antioxidant in the brain. Dysregulation of the GSH system has been shown in animal models of depression163 and may be related to reductions in glutamatergic activity at the glutamate N-methyl-d-aspartate (NMDA) receptor and attenuation in the production of neurotrophic factors as demonstrated in various neurodegenerative disorders. Although clinical studies have shown limited efficacy, this pattern of abnormality has been reversed through the administration of the GSH precursor N-acetylcysteine.164,165 Owing to greater signal separation techniques,166,167 1 H-MRS detection of GSH has led to a growing number of studies measuring GSH as a marker of oxidative stress in patients with MDD. Reductions in GSH have been associated with MDD168 and, more

Neuroimaging biomarkers in MDD

specifically, anhedonia in patients with MDD.169 However, the use of 1 H-MRS GSH has not been promising as a treatment biomarker. Despite a 6-week escitalopram treatment that resulted in an improvement of depression, concentrations of GSH did not change (P = 0.08) following treatment course.168 Further studies using other inflammatory markers may become more fruitful in determining an adequate treatment biomarker. 31 P-MRS is used to determine cerebral levels of high-energy phosphates, including phosphocreatine (PCr) and nucleoside triphosphates (NTPs) such as adenosine triphosphate (ATP), a key intracellular energy carrier and main component of the NTP peaks. MDD patients exhibit abnormalities in brain energy metabolism, reflected by decreases in NTP levels in the frontal lobe and basal ganglia, particularly the ␤-NTP fraction most closely reflecting ATP levels.170–172 While ATP levels are decreased, levels of PCr, a reservoir for high-energy phosphates that can be used to generate ATP, are increased in MDD patients.173 Taken together, these results suggest that the NTP decreases reflect reductions in cellular bioenergetic metabolism, consistent with the previously discussed alterations in brain phospholipid metabolism reflected by increased Cho levels. Both sets of data suggest that mitochondrial dysfunction may play a critical role in the pathophysiology of MDD. Importantly, bioenergetic metabolism has been correlated with the response to antidepressant treatment. MDD subjects who responded to antidepressant treatment had lower NTP and higher PCr levels at baseline than treatment nonresponders.172,173 Baseline PCr was found to be a potentially useful predictor of antidepressant response with 83% sensitivity and 75% specificity.173 Also, during antidepressant treatment, total NTP and ␤-NTP increased while PCr decreased in treatment responders, but these changes were not seen in nonresponders. Conclusion and future directions Major brain regions implicated in structural and functional studies of patients with MDD are summarized in Table 1 and depicted in Figure 1. Given the convergence of structural and functional brain abnormalities within frontolimbic brain regions, symptomatologic manifestation of MDD is likely owing to an interplay of molecular and cellular processes causing sequential changes to specific

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

57

Neuroimaging biomarkers in MDD

Lener & Iosifescu

Table 1. Summary of select potential neuroimaging biomarkers of MDD and MDD treatment prediction Gray matter volume Brain regions Prefrontal regions

Limbic regions

Hippocampus Striatal regions

Insular cortex

MDD

White matter (fractional anisotropy)

Treatment biomarker

Treatment biomarker

MDD

⇓dlPFC ⇓vlPFC ⇓mPFC ⇓OFC ⇓ACC ⇓Amyg

⇑OFC

⇓Sup. frontal ⇓Mid. frontal ⇓Body/genu of CC

⇑Sup. frontal ⇓L mid. frontal

⇑ACC

⇑Ant. cingulum

⇓Hippo

⇑Hippo

⇓Ant. cingulum ⇓Fornix ⇓UF ⇓Fornix ⇓Thalamo-cortical WM relays

⇓Ant. insula

⇓UF

Brain activation MDD

Treatment biomarker

⇓dlPFCa ⇓vlPFCa ⇓mPFCa,b ⇑dmPFCc ⇑ACCa,b,c ⇑Amyga

⇑ACC,a,b AD ⇓ACC,a,b CBT

⇑Thalamusb ⇓Thalamusc ⇑Putamenb ⇓Caudateb ⇓NAccb,c ⇑Insulab

a Associated

with emotion processing. with reward processing. c Associated with self-referential processing. b Associated

neural systems involved in emotion regulation, reward, and cognition that can be observed in response to a task or at rest. Peripheral systems, such as adrenal and inflammatory systems, may also be involved in mediating microinjuries to vulnerable brain regions, causing metabolic changes and subsequent imbalances to excitatory and inhibitory

neurotransmitter systems that lead to functional brain abnormalities. Neuroimaging modalities have shown promise in detecting abnormalities associated with MDD and with response to specific treatments. In studies examining the influence of genes on brain volume in MDD, growing evidence supports a role of

Figure 1. Brain regions implicated in structural and functional studies as potential illness and treatment biomarkers in major depressive disorder. (A) Lateral view of the cortical surface of the brain, showing the dorsolateral prefrontal cortex (dlPFC), ventrolateral PFC (vlPFC), orbitofrontal cortex (OFC), and insular cortex (ins). (B) Mid-sagittal view of the brain, showing the medial PFC (mPFC), anterior cingulate cortex (subgenual (sgACC), rostral (rACC), dorsal (dACC)), striatum (stria), amygdala (Am), and hippocampus (Hp). Regions are implicated as a potential (a) illness or (b) treatment biomarker. Of note, the striatum, amygdala, and insula are deep to the region indicated and are not seen on the specimen. Photographs of anatomical specimens courtesy of T.P. Naidich, M.E. Fowkes, and C.Y. Tang.

58

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

the BDNF Val66Met polymorphism on gray matter volumetric reductions (e.g., Refs. 174 and 175) and COMT Val158Met on white matter volumetric reductions.176,177 Therefore, the development of treatment-predictive neuroimaging biomarkers will likely require an integrated approach among different imaging modalities and linking them to peripheral biomarkers (e.g., inflammatory), genetic and epigenetic biomarkers, and clinical symptoms (e.g., anhedonia) or domains (e.g., reward processing). Future directions of research determining biomarkers of MDD have already begun as several groups have formed collaborations in assembling large, multicenter data sets of patients with MDD, such as the initiatives Establishing Moderators and Biosignatures of Antidepressant Response for Clinical Care for Depression (EMBARC; http://embarc.utsouthwestern.edu) and International Study to Predict Optimized Treatment in Depression (iSPOT-D) (http://www.brainresource. com/research/ispot/ispotd). These large studies using multimodal neuroimaging may help circumvent the limitation of low-powered studies to examine the association of biological and psychological domains of depressive illness with specific neuroimaging biomarkers. Conflicts of interest Dr. Lener declares no conflicts of interest. In the past three years, Dr. Iosifescu has consulted for Avanir, Axsome, CNS Response, INSYS Therapeutics, Lundbeck, Otsuka, Servier, and Sunovion, and has received research support through the Icahn School of Medicine at Mount Sinai from Alkermes, Astra Zeneca, Brainsway, Euthymics, Neosync, Roche, and Shire. Supporting Information Additional supporting information may be found in the online version of this article. Table S1. Reviewed studies of neuroimaging biomarkers of treatment response. References 1. Mueller, T.I., A.C. Leon, M.B. Keller, et al. 1999. Recurrence after recovery from major depressive disorder during 15 years of observational follow-up. Am. J. Psychiatry 156: 1000–1006.

Neuroimaging biomarkers in MDD

2. Rush, A.J., M.H. Trivedi, S.R. Wisniewski, et al. 2006. Acute and longer-term outcomes in depressed outpatients requiring one or several treatment steps: a STAR*D report. Am. J. Psychiatry 163: 1905–1917. 3. Kessler, R.C., P. Berglund, O. Demler, et al. 2003. The epidemiology of major depressive disorder: results from the National Comorbidity Survey Replication (NCS-R). JAMA 289: 3095–3105. 4. Fava, M. & K.G. Davidson. Deinition and epidemiology of treatment-resistant depression. Psychiatr. Clin. North Am. 19: 179–200. 5. Fava, M. 2003. Diagnosis and definition of treatmentresistant depression. Biol. Psychiatry 53: 649–659. 6. Fava, M. & A.J. Rush. 2006. Current status of augmentation and combination treatments for major depressive disorder: a literature review and a proposal for a novel approach to improve practice. Psychother. Psychosom. 75: 139–153. 7. World Health Organization. 2006. World Health Report 2006. Statistical annex. http://www.who.int/whr/2006/en/ 8. Mrazek, D.A., J.C. Hornberger, C.A. Altar & I. Degtiar. 2014. A review of the clinical, economic, and societal burden of treatment-resistant depression: 1996–2013. Psychiatr. Serv 65: 977–987. 9. Cuthbert, B. 2010. New, rapidly-acting treatments for treatment-resistant depression (RAPID). National Advisory Mental Health Council (NAMHC) Concept Clearance. National Institute of Mental Health. May 14, 2010. 10. Niciu, M.J., D.C. Mathews, A.C. Nugent, et al. 2014. Developing biomarkers in mood disorders research through the use of rapid-acting antidepressants. Depress Anxiety 31: 297–307. 11. McGrath, C.L., M.E. Kelley, P.E. Holtzheimer, et al. 2013. Toward a neuroimaging treatment selection biomarker for major depressive disorder. JAMA Psychiatr. 70: 821– 829. 12. Insel, T., B. Cuthbert, M. Garvey, et al. 2010. Resarch Domain Criteria (RDoC): toward a new classification framework for research on mental disorders. Am. J. Psychiatry 167: 748–751. 13. Lorenzetti, V., N.B. Allen, A. Fornito & M. Y¨ucel. 2009. Structural brain abnormalities in major depressive disorder: a selective review of recent MRI studies. J. Affect Disord. 117: 1–17. ` et al. 2011. 14. Kempton, M.J., Z. Salvador, M.R. Munafo, Structural neuroimaging studies in major depressive disorder. Meta-analysis and comparison with bipolar disorder. Arch. Gen. Psychiatry 68: 675–690. 15. Bora, E., B.J. Harrison, C.G. Davey, et al. 2012. Metaanalysis of volumetric abnormalities in cortico-striatalpallidal-thalamic circuits in major depressive disorder. Psychol. Med. 42: 671–681. 16. Du, M.Y., Q.Z. Wu, Q. Yue, et al. 2012. Voxelwise metaanalysis of gray matter reduction in major depressive disorder. Prog. Neuropsychopharmacol. Biol. Psychiatry 36: 11–16. 17. Sacher, J., J. Neumann, T. F¨unfst¨uck, et al. 2012. Mapping the depressed brain: a meta-analysis of structural and functional alterations in major depressive disorder. J. Affect Disord. 140: 142–148.

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

59

Neuroimaging biomarkers in MDD

Lener & Iosifescu

18. Lai, C.H. 2013. Gray matter volume in major depressive disorder: a meta-analysis of voxel-based morphometry studies. Psychiatry Res. 211: 37–46. 19. Drevets, W.C., J.L. Price & M.L. Furey. 2008. Brain structural and functional abnormalities in mood disorders: implications for neurocircuitry models of depression. Brain Struct. Funct. 213: 93–118. 20. Rajkowska, G. et al. 1999. Morphometric evidence for neuronal and glial prefrontal cell pathology in major depression. Biol. Psychiatry 45: 1085–1098. 21. Rajkowska, G. 2000. Postmortem studies in mood disorders indicate altered numbers of neurons and glial cells. Biol. Psychiatry 48: 766–777. 22. Ongür, D., W.C. Drevets & J.L. Price. 1998. Glial reduction in the subgenual prefrontal cortex in mood disorders. Proc. Nat. Acad. Sci. U.S.A. 95: 13290–13295. 23. Duman, R.S. & G.K. Aghajanian. 2012. Synaptic dysfunction in depression: potential therapeutic targets. Science 338: 68–72. 24. Willard, S.L., D.R. Riddle, M.E. Forbes & C.A. Shively. 2013. Cell number and neuropil alterations in subregions of the anterior hippocampus in a female monkey model of depression. Biol. Psychiatry 74: 890–897. 25. Amico, F., E. Meisenzahl, N. Koutsouleris, et al. 2011. Structural MRI correlates for vulnerability and resilience to major depressive disorder. J. Psychiatry Neurosci. 36: 15– 22. 26. Bremner, J.D., M. Narayan, E.R. Anderson, et al. 2000. Hippocampal volume reduction in major depression. Am. J. Psychiatry 157: 115–118. 27. McKinnon, M.C., K. Yucel, A. Nazarov & G.M. MacQueen. 2009. A meta-analysis examining clinical predictors of hippocampal volume in patients with major depressive disorder. J. Psychiatry Neurosci. 34: 41–54. 28. Sheline, Y.I., M.H. Gado & H.C. Kraemer. 2003. Untreated depression and hippocampal volume loss. Am. J. Psychiatry 160: 1516–1518. 29. Frodl, T., E.M. Meisenzahl, T. Zetzsche, et al. 2004. Hippocampal and amygdala changes in patients with major depressive disorder and healthy controls during a 1-year follow-up. J. Clin. Psychiatry 65: 492–499. 30. Frodl, T., M. J¨ager, I. Smajstrlova, et al. 2008. Effect of hippocampal and amygdala volumes on 28. Clinical outcomes in major depression: a 3-year prospective magnetic resonance imaging study. J. Psychiatry Neurosci. 33: 423–430. 31. Malykhin, N.V., R. Carter, P. Seres & N.J. Coupland. 2010. Structural changes in the hippocampus in major depressive disorder: contributions of disease and treatment. J. Psychiatry Neurosci. 35: 337–343. 32. Huang, Y., N.J. Coupland, R.M. Lebel, et al. 2013. Structural changes in hippocampal subfields in major depressive disorder: a high–field magnetic resonance imaging study. Biol. Psychiatry 74: 62–68. 33. Nordanskog, P., U. Dahlstrand, M.R. Larsson, et al. 2010. Increase in hippocampal volume after electroconvulsive therapy in patients with depression: a volumetric magnetic resonance imaging study. J ECT. 26: 62–67. 34. Whittle, S., R. Lichter, M. Dennison, et al. 2014. Structural brain development and depression onset during adolescence: a prospective longitudinal study. Am. J. Psychiatry 171: 564–571. 60

35. Yucel, K., M. McKinnon, R. Chahal, et al. 2009. Increased subgenual prefrontal cortex size in remitted patients with major depressive disorder. Psychiatry Res. 173: 71–76. 36. Takahashi, T., M. Y¨ucel, V. Lorenzetti, et al. 2010. Volumetric MRI study of the insular cortex in individuals with current and past major depression. J. Affect. Disord. 121: 231–238. 37. Salvadore, G., A.C. Nugent, H. Lemaitre, et al. 2011. Prefrontal cortical abnormalities in currently depressed versus currently remitted patients with major depressive disorder. Neuroimage 54: 2643–2651. 38. Liu, C.H., B. Jing, X. Ma, et al. 2014. Voxel-based morphometry study of the insular cortex in female patients with current and remitted depression. Neuroscience 262: 190–199. 39. Chen, C.H., K. Ridler, J. Suckling, et al. 2007. Brain imaging correlates of depressive symptom severity and predictors of symptom improvement after antidepressant treatment. Biol. Psychiatry 62: 407–414. 40. Soriano-Mas, C., R. Hern´andez-Ribas, J. Pujol, et al. 2011. Cross-sectional and longitudinal assessment of structural brain alterations in melancholic depression. Biol. Psychiatry 69: 318–325. 41. S¨amann, P.G., D. H¨ohn, N. Chechko, et al. 2013. Prediction of antidepressant treatment response from gray matter volume across diagnostic categories. Eur. Neuropsychopharmacol 23: 1503–1515. 42. MacQueen, G.M., K. Yucel, V.H. Taylor, et al. 2008. Posterior hippocampal volumes are associated with remission rates in patients with major depressive disorder. Biol. Psychiatry 64: 880–883. 43. Phillips, J.L., L.A. Batten, F. Aldosary, et al. 2012. Brainvolume increase with sustained remission in patients with treatment-resistant unipolar depression. J. Clin. Psychiatry 73: 625–631. 44. Li, C.T., C.P. Lin, K.H. Chou, et al. 2010. Structural and cognitive deficits in remitting and non-remitting recurrent depression: a voxel-based morphometric study. Neuroimage 50: 347–356. 45. Fu, C.H., H. Steiner & S.G. Costafreda. 2013. Predictive neural biomarkers of clinical response in depression: a meta-analysis of functional and structural neuroimaging studies of pharmacological and psychological therapies. Neurobiol. Dis. 52: 75–83. 46. Gong, Q., Q. Wu, C. Scarpazza, et al. 2011. Prognostic prediction of therapeutic response in depression using highfield MR imaging. Neuroimage 55: 1497–1503. 47. Krishnan, K.R., J.C. Hays & D.G. Blazer. 1997. MRI-defined vascular depression. Am. J. Psychiatry 154: 497–501. 48. deGroot, J.C., F.E. deLeeuw, M. Oudkerk, et al. 2000. Cerebral white matter lesions and depressive symptoms in elderly adults. Arch. Gen. Psychiatry 57: 1071– 1076. 49. Lyoo, I.K., H.K. Lee, J.H. Jung, et al. 2002. White matter hyperintensities on magnetic resonance imaging of the brain in children with psychiatric disorders. Compr. Psychiatry 43: 361–368. 50. Lenze, E., D. Cross, D. McKeel, et al. 1999. White matter hyperintensities and gray matter lesions in physically healthy depressed subjects. Am. J. Psychiatry 156: 1602– 1607.

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

51. Iosifescu, D.V., P.F. Renshaw, I.K. Lyoo, et al. 2006. Brain white-matter hyperintensities and treatment outcome in major depressive disorder. Br. J. Psychiatry 188: 180–185. 52. Alexopoulos, G.S., B.S. Meyers, R.C. Young, et al. 1997. ‘Vascular depression’ hypothesis. Arch. Gen. Psychiatry 54: 915–922. 53. Simpson, S., R.C. Baldwin, A. Jackson & A.S. Burns. 1998. Is subcortical disease associated with a poor response to antidepressants? Neurological, neuropsychological and neuroradiological findings in late-life depression. Psychol. Med. 28: 1015–1026. 54. Iosifescu, D.V., P.F. Renshaw, D.D. Dougherty, et al. 2007. Major depressive disorder with anger attacks and subcortical MRI white matter hyperintensities. J. Nerv. Ment. Dis. 195: 175–178. 55. O’Brien, J., D. Ames, E. Chiu, I. Schweitzer, et al. 1998. Severe deep white matter lesions and outcome in elderly patients with major depressive disorder: follow up study. BMJ 317: 982–984. 56. Sexton, C.E., C.E. Mackay, K.P. Ebmeier. 2009. A systematic review of diffusion tensor imaging studies in affective disorders. Biol. Psychiatry 66: 814–823. 57. Liao, Y., X. Huang, Q. Wu, et al. 2013. Is depression a disconnection syndrome? Meta–analysis of diffusion tensor imaging studies in patients with MDD. J. Psychiatry Neurosci. 38: 49–56. 58. Alexopoulos, G.S., D.N. Kiosses, S.J. Choi, et al. 2002. Frontal white matter microstructure and treatment response of late-life depression: a preliminary study. Am. J. Psychiatry 159: 1929–1932. 59. Taylor, W.D., J.R. MacFall, M.E. Payne, et al. 2004. Late-life depression and microstructural abnormalities in dorsolateral prefrontal cortex white matter. Am. J. Psychiatry 161: 1293–1296. 60. Bae, J.N., J.R. MacFall, K.R. Krishnan, et al. 2006. Dorsolateral prefrontal cortex and anterior cingulate cortex white matter alterations in late-life depression. Biol. Psychiatry 60: 1356–1363. 61. Cole, J., C.A. Chaddock, A.E. Farmer, et al. 2012. White matter abnormalities and illness severity in major depressive disorder. Br. J. Psychiatry 201: 33–39. 62. Carballedo, A., F. Amico, I. Ugwu, et al. 2012. Reduced FA in the uncinate fasciculus in patients with MD carrying the met-allele of the Val66Met BDNF genotype. Am. J. Med. Genet. B Neuropsychiatr. Genet. 159B: 537–548. 63. Cullen, K.R., B. Klimes-Dougan, R. Muetzel, et al. 2010. Altered white matter microstructure in adolescents with major depression: a preliminary study. J. Am. Acad. Child Adolesc. Psychiatry 49: 173–183. 64. Osoba, A., J. H¨anggi, M. Li, et al. 2013. Disease severity is correlated to tract specific changes of fractional anisotropy in MD and CM thalamus—a DTI study in major depressive disorder. J. Affect. Disord. 149: 116–128. 65. Zou, K., W. Deng, T. Li, et al. 2010. Changes of brain morphometry in first-episode, drug-na¨ıve, non-late-life adult patients with major depression: an optimized voxel-based morphometry study. Biol. Psychiatry 67: 186–188. 66. Price, J.L. & W.C. Drevets. 2010. Neurocircuitry of mood disorders. Neuropsychopharmacology 35: 192–216.

Neuroimaging biomarkers in MDD

67. Abe, O., H. Takao, W. Gonoi, et al. 2010. Voxel-based analysis of the diffusion tensor. Neuroradiology 52: 699–710. 68. Kieseppa, T., M. Eerola, R. Mantyla, et al. 2010. Major depressive disorder and white matter abnormalities: a diffusion tensor imaging study with tract-based spatial statistics. J. Affect. Disord. 120: 240–244. 69. Korgaonkar, M.S., S.M. Grieve, S.H. Koslow, et al. 2010. Loss of white matter integrity in major depressive disorder: evidence using tract-based spatial statistical analysis of diffusion tensor imaging. Hum. Brain Mapp. 32: 2161– 2171. 70. Choi, K.S., P.E. Holtzheimer, A.R. Franco, et al. 2014. Reconciling variable findings of white matter integrity in major depressive disorder. Neuropsychopharmacology 39: 1332– 1339. 71. Kear-Colwell, J.J. 1972. A taxonomy of depressive phenomena and its relationships to the reactive-endogenous dichotomy. Br. J. Psychiatry 121: 665–671. 72. Quinn, C., A. Harris, & A.H. Kemp. 2012. The interdependence of subtype and severity: contributions of clinical and neuropsychological features to melancholia and nonmelancholia in an outpatient sample. J. Int. Neuropsychol. Soc. 18: 361–369. 73. Teicher, M.H. & J.A. Samson. 2013. Childhood maltreatment and psychopathology: a case for ecophenotypic variants as clinically and neurobiologically distinct subtypes. Am. J. Psychiatry 170: 1114–1133. 74. Malki, K., R. Keers, M.G. Tosto, et al. 2014. The endogenous and reactive depression subtypes revisited: integrative animal and human studies implicate multiple distinct molecular mechanisms underlying major depressive disorder. BMC Med. 12: 73. 75. Ionescu, D.F., M.J. Niciu, D.C. Mathews, et al. 2013. Neurobiology of anxious depression: a review. Depress Anxiety 30: 374–385. 76. Spashett, R., G. Fernie, I.C. Reid & I.M. Cameron. 2014. MADRS symptom subtypes in ECT-treated depressed patients: relationship to response and subsequent ECT. J. ECT 30: 227–231. 77. Flint, J. & K.S. Kendler. 2014. The genetics of major depression. Neuron 81: 484–503. 78. van Loo, H.M., T. Cai, M.J. Gruber, et al. 2014. Major depressive disorder subtypes to predict long-term course. Depress Anxiety 31: 765–777. 79. B¨uhler, J., 2014. Seem¨uller F2, L¨age D3. The predictive power of subgroups: an empirical approach to identify depressive symptom patterns that predict response to treatment. J. Affect. Disord. 163: 81–87. 80. Shah, P.J., M.F. Glabus, G.M. Goodwin & K.P. Ebmeier. 2002. Chronic, treatment-resistant depression and right fronto-striatal atrophy. Br. J. Psychiatry 180: 434–440. 81. Li, C.T., C.P. Lin, K.H. Chou, et al. 2010. Structural and cognitive deficits in remitting and non-remitting recurrent depression: a voxel-based morphometric study. Neuroimage 50: 347–356. 82. Zhou, Y., L.D. Qin, J. Chen, et al. 2011. Brain microstructural abnormalities revealed by diffusion tensor images in patients with treatment-resistant depression compared with major depressive disorder before treatment. Eur. J. Radiol. 80: 450–454.

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

61

Neuroimaging biomarkers in MDD

Lener & Iosifescu

83. Guo, W.B., F. Liu, J.D. Chen, et al. 2012. Altered white matter integrity of forebrain in treatment-resistant depression: a diffusion tensor imaging study with tract-based spatial statistics. Prog. Neuropsychopharmacol. Biol. Psychiatry 38: 201–206. 84. Peng, H.J., H.R. Zheng, Y.P. Ning, et al. 2013. Abnormalities of cortical-limbic-cerebellar white matter networks may contribute to treatment-resistant depression: a diffusion tensor imaging study. BMC Psychiatry 13: 72. 85. Hoogenboom, W.S., R.H. Perlis, J.W. Smoller, et al. 2014. Limbic system white matter microstructure and long-term treatment outcome in major depressive disorder: a diffusion tensor imaging study using legacy data. World J. Biol. Psychiatry 15: 122–134. ´ ´ et al. 86. de Diego-Adeli˜no, J., P. Pires, B. Gomez-Ans on, 2014. Microstructural white-matter abnormalities associated with treatment resistance, severity and duration of illness in major depression. Psychol. Med 44: 1171–1182. 87. Taylor, W.D., M. Kuchibhatla, M.E. Payne, et al. 2008. Frontal white matter anisotropy and antidepressant remission in late-life depression. PLoS One 3: e3267. 88. Gutman, D.A., P.E. Holtzheimer, T.E. Behrens, et al. 2009. A tractography analysis of two deep brain stimulation white matter targets for depression. Biol. Psychiatry 65: 276–282. 89. H¨oflich, A., M. Savli, E. Comasco, et al. 2013. Neuropsychiatric deep brain stimulation for translational neuroimaging. Neuroimage 9: 30–41. 90. Peng, H., H. Zheng, L. Li, et al. 2012. High-frequency rTMS treatment increases white matter FA in the left middle frontal gyrus in young patients with treatment-resistant depression. J. Affect. Disord. 136: 249–257. 91. Korgaonkar, M.S., L.M. Williams, Y.J. Song, et al. 2014. Diffusion tensor imaging predictors of treatment outcomes in major depressive disorder. Br. J. Psychiatry 205: 321–328. 92. Diener, C., C. Kuehner, W. Brusniak, et al. 2012. A metaanalysis of neurofunctional imaging studies of emotion and cognition in major depression. Neuroimage 61: 677–685. 93. Hamilton, J.P., A. Etkin, D.J. Furman, et al. 2012. Functional neuroimaging of major depressive disorder: a metaanalysis and new integration of base line activation and neural response data. Am. J. Psychiatry 169: 693–703. 94. Kerestes, R., C.G. Davey, K. Stephanou, et al. 2013. Functional brain imaging studies of youth depression: a systematic review. Neuroimage Clin. 4: 209–231. 95. Treadway, M.T. & D.A. Pizzagalli. 2014. Imaging the pathophysiology of major depressive disorder—from localist models to circuit-based analysis. Biol. Mood Anxiety Disord. 4: 5. 96. Drevets, W.C., T.O. Videen, J.L. Price, et al. 1992. A functional anatomical study of unipolar depression. J. Neurosci. 12: 3628–3641. 97. Sheline, Y.I., D.M. Barch, J.M. Donnelly, et al. 2001. Increased amygdala response to masked emotional faces in depressed subjects resolves with antidepressant treatment: an fMRI study. Biol. Psychiatry 50: 651–658. 98. Siegle, G.J., S.R. Steinhauer, M.E. Thase, et al. 2002. Can’t shake that feeling: event-related fMRI assessment of sustained amygdala activity in response to emotional information in depressed individuals. Biol. Psychiatry 51: 693–707.

62

99. Phillips, M.L., W.C. Drevets, S.L. Rauch & R. Lane. 2003. Neurobiology of emotion perception II: implications for major psychiatric disorders. Biol. Psychiatry 54: 515–528. 100. Victor, T.A., M.L. Furey, S.J. Fromm, et al. 2010. Relationship between amygdala responses to masked faces and mood state and treatment in major depressive disorder. Arch. Gen. Psychiatry 67: 1128–1138. 101. Fossati, P., S.J. Hevenor, S.J. Graham, et al. 2003. In search of the emotional self: an fMRI study using positive and negative emotional words. Am. J. Psychiatry 160: 1938– 1945. 102. Fu, C.H., S.C. Williams, A.J. Cleare, et al. 2004. Attenuation of the neural response to sad faces in major depression by antidepressant treatment: a prospective, event-related functional magnetic resonance imaging study. Arch. Gen. Psychiatry 61: 877–889. 103. Ochsner, K.N., R.D. Ray, J.C. Cooper, et al. 2004. For better or for worse: neural systems supporting the cognitive down- and up-regulation of negative emotion. Neuroimage 23: 483–499. 104. Ochsner, K.N., R.R. Ray, B. Hughes, et al. 2009. Bottom-up and top-down processes in emotion generation: common and distinct neural mechanisms. Psychol. Sci. 20: 1322– 1331. 105. Johnstone, T., C.M. vanReekum, H.L. Urry, et al. 2007. Failure to regulate: counterproductive recruitment of topdown prefrontal-subcortical circuitry in major depression. J. Neurosci. 27: 8877–8884. 106. Townsend, J. & L.L. Altshuler. 2012. Emotion processing and regulation in bipolar disorder: a review. Bipolar Disord. 14: 326–339. 107. Mochcovitch, M.D., R.C. da Rocha Freire, R.F. Garcia & A.E. Nardi. 2014. A systematic review of fMRI studies in generalized anxiety disorder: evaluating its neural and cognitive basis. J. Affect. Disord. 167C: 336–342. 108. Etkin, A. & A.F. Schatzberg. 2011. Common abnormalities and disorder-specific compensation during implicit regulation of emotional processing in generalized anxiety and major depressive disorders. Am. J. Psychiatry 168: 968–978. 109. Etkin, A. & T.D. Wager. 2007. Functional neuroimaging of anxiety: a meta-analysis of emotional processing in PTSD, social anxiety disorder, and specific phobia. Am. J. Psychiatry 164: 1476–1488. 110. Hazlett, E.A., J. Zhang, A.S. New, et al. 2012. Potentiated amygdala response to repeated emotional pictures in borderline personality disorder. Biol. Psychiatry 72: 448–456. 111. Malhi, G.S., M. Tanious, K. Fritz, et al. 2013. Differential engagement of the fronto-limbic network during emotion processing distinguishes bipolar and borderline personality disorder. Mol. Psychiatry 18: 1247–1248. 112. Mitterschiffthaler, M.T., V. Kumari, G.S. Malhi, et al. 2003. Neural response to pleasant stimuli in anhedonia: an fMRI study. Neuroreport 14: 177–182. 113. Keedwell, P.A., C. Andrew, S.C. Williams, et al. 2005. A double dissociation of ventromedial prefrontal cortical responses to sad and happy stimuli in depressed and healthy individuals. Biol. Psychiatry 58: 495–503. 114. Pizzagalli, D.A., A.J. Holmes, D.G. Dillon, et al. 2009. Reduced caudate and nucleus accumbens response to

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

115.

116.

117.

118.

119.

120.

121.

122.

123.

124.

125.

126.

127.

128.

rewards in unmedicated individuals with major depressive disorder. Am. J. Psychiatry 166: 702–710. Nejad, A.B., P. Fossati & C. Lemogne. 2013. Self-referential processing, rumination, and cortical midline structures in major depression. Front. Hum. Neurosci. 7: 666. Kross, E., M. Davidson, J. Weber & K. Ochsner. 2009. Coping with emotions past: the neural bases of regulating affect associated with negative autobiographical memories. Biol. Psychiatry 65: 361–366. Lemogne, C., G. le Bastard, H. Mayberg, et al. 2009. In search of the depressive self: extended medial prefrontal network during self-referential processing in major depression. Soc. Cogn. Affect. Neurosci. 4: 305–312. Yoshimura, S., Y. Okamoto, K. Onoda, et al. 2010. Rostral anterior cingulate cortex activity mediates the relationship between the depressive symptoms and the medial prefrontal cortex activity. J. Affect. Disord. 122: 76–85. Grimm, S., J. Ernst, P. Boesiger, et al. 2009. Increased self-focus in major depressive disorder is related to neural abnormalities in subcortical-cortical midline structures. Hum. Brain Mapp. 30: 2617–2627. Sheline, Y.I., D.M. Barch, J.L. Price, et al. 2009. The default mode network and self-referential processes in depression. Proc. Natl. Acad. Sci. U.S.A. 106: 1942–1947. Brody, A.L., S. Saxena, D.H. Silverman, et al. 1999. Brain metabolic changes in major depressive disorder from preto posttreatment with paroxetine. Psychiatry Res. 91: 127– 139. Little, J.T., T.A. Ketter, T.A. Kimbrell, et al. 2005. Bupropion and venlafaxine responders differ in pretreatment regional cerebral metabolism in unipolar depression. Biol. Psychiatry 57: 220–228. Miller, J.M., N. Schneck, G.J. Siegle, et al. 2013. fMRI response to negative words and SSRI treatment outcome in major depressive disorder: a preliminary study. Psychiatry Res. 214: 296–305. Miller, J.M., N. Schneck, G.J. Siegle, et al. 2014. Corrigendum to “fMRI response to negative words and SSRI treatment outcome in major depressive disorder: a preliminary study” [Psychiatry Res.: Neuroimaging 214 (2013) 296–305] Psychiatry Res. 222: 173–175. Davidson, R.J., W. Irwin, M.J. Anderle & N.H. Kalin. 2003. The neural substrates of affective processing in depressed patients treated with venlafaxine. Am. J. Psychiatry 160: 64–75. Keedwell, P.A., D. Drapier, S. Surguladze, et al. 2010. Subgenual cingulate and visual cortex responses to sad faces predict clinical outcome during antidepressant treatment for depression. J. Affect. Disord. 120: 120–125. Keedwell, P., D. Drapier, S. Surguladze, et al. 2009. Neural markers of symptomatic improvement during antidepressant therapy in severe depression: subgenual cingulate and visual cortical responses to sad, but not happy, facial stimuli are correlated with changes in symptom score. J. Psychopharmacol. 23: 775–788. Mayberg, H.S., S.K. Brannan, R.K. Mahurin, et al. 1997. Cingulate function in depression: a potential predictor of treatment response. Neuroreport 8: 1057–1061.

Neuroimaging biomarkers in MDD

129. Pizzagalli, D.A. 2011. Frontocingulate dysfunction in depression: toward biomarkers of treatment response. Neuropsychopharmacology 36: 183–206. 130. Salvadore, G., B.R. Cornwell, V. Colon-Rosario, et al. 2009. Increased anterior cingulate cortical activity in response to fearful faces: a neurophysiological biomarker that predicts rapid antidepressant response to ketamine. Biol. Psychiatry 65: 289–295. 131. Salvadore, G., B.R. Cornwell, F. Sambataro, et al. 2010. Anterior cingulate desynchronization and functional connectivity with the amygdala during a working memory task predict rapid antidepressant response to ketamine. Neuropsychopharmacology 35: 1415–1422. 132. Siegle, G.J., C.S. Carter & M.E. Thase. 2006. Use of fMRI to predict recovery from unipolar depression with cognitive behavior therapy. Am. J. Psychiatry 163: 735–738. 133. Siegle, G.J., W.K. Thompson, A. Collier, et al. 2012. Toward clinically useful neuroimaging in depression treatment: prognostic utility of subgenual cingulate activity for determining depression outcome in cognitive therapy across studies, scanners, and patient characteristics. Arch. Gen. Psychiatry 69: 913–924. 134. Milak, M.S., R.V. Parsey, L. Lee, et al. 2009. Pretreatment regional brain glucose uptake in the midbrain on PET may predict remission from a major depressive episode after three months of treatment. Psychiatry Res. 173: 63–70. 135. Konarski, J.Z., S.H. Kennedy, Z.V. Segal, et al. 2009. Predictors of nonresponse to cognitive behavioural therapy or venlafaxine using glucose metabolism in major depressive disorder. J. Psychiatry Neurosci. 34: 175–180. 136. DeRubeis, R.J., G.J. Siegle & S.D. Hollon. 2008. Cognitive therapy versus medication for depression: treatment outcomes and neural mechanisms. Nat. Rev. Neurosci. 9: 788–796. 137. Roiser, J.P., R. Elliott & B.J. Sahakian. 2012. Cognitive mechanisms of treatment in depression. Neuropsychopharmacology 37: 117–136. 138. Sanacora, G., G. Treccani & M. Popoli. 2012. Towards a glutamate hypothesis of depression: an emerging frontier of neuropsychopharmacology for mood disorders. Neuropharmacology 62: 63–77. 139. Yildiz–Yesiloglu, A. & D.P. Ankerst. 2006. Review of 1H magnetic resonance spectroscopy findings in major depressive disorder: a meta-analysis. Psychiatry Res. 147: 1–25. 140. Gabbay, V., D.A. Hess, S. Liu, et al. 2007. Lateralized caudate metabolic abnormalities in adolescent major depressive disorder: a proton MR spectroscopy study. Am. J. Psychiatry 164: 1881–1889. 141. Kondo, D.G., T.L. Hellem, Y.H. Sung, et al. 2011. Review: magnetic resonance spectroscopy studies of pediatric major depressive disorder. Depress Res. Treat. 2011: 650450. 142. Caverzasi, E., A. Pichiecchio, G.U. Poloni, et al. 2012. Magnetic resonance spectroscopy in the evaluation of treatment efficacy in unipolar major depressive disorder: a review of the literature. Funct. Neurol. 27: 13–22. 143. Obergriesser, T., G. Ende, D.F. Braus & F.A. Henn. 2003. Long-term follow-up of magnetic resonance-detectable choline signal changes in the hippocampus of patients

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

63

Neuroimaging biomarkers in MDD

144.

145.

146.

147.

148.

149.

150.

151.

152.

153.

154.

155.

156.

157.

158.

64

Lener & Iosifescu

treated with electroconvulsive therapy. J. Clin. Psychiatry 64: 775–780. Block, W., F. Tr¨aber, O. vonWiddern, et al. 2009. Proton MR spectroscopy of the hippocampus at 3 T in patients with unipolar major depressive disorder: correlates and predictors of treatment response. Int. J. Neuropsychopharmacol. 12: 415–422. S¨ozeri-Varma, G., N. Kalkan-O˘guzhanoglu, M. Efe, et al. 2013. Neurochemical metabolites in prefrontal cortex in patients with mild/moderate levels in first-episode depression. Neuropsychiatr. Dis. Treat. 9: 1053–1059. Sonawalla, S.B., P.F. Renshaw, C.M. Moore, et al. 1999. Compounds containing cytosolic choline in the basal ganglia: a potential biological marker of true drug response to fluoxetine. Am. J. Psychiatry 156: 1638–1640. Gonul, A.S., O. Kitis, E. Ozan, et al. 2006. The effect of antidepressant treatment on N-acetyl aspartate levels of medial frontal cortex in drug-free depressed patients. Prog. Neuropsychopharmacol. Biol. Psychiatry 30: 120–125. Ende, G., D.F. Braus, S. Walter, et al. 2000. The hippocampus in patients treated with electroconvulsive therapy: a proton magnetic resonance spectroscopic imaging study. Arch. Gen. Psychiatry 57: 937–943. Sanacora, G., G.F. Mason, D.L. Rothman, et al. 1999. Reduced cortical ␥ -aminobutyric acid levels in depressed patients determined by proton magnetic resonance spectroscopy. Arch. Gen. Psychiatry 56: 1043–1047. Sanacora, G., R. Gueorguieva, C.N. Epperson, et al. 2004. Subtype-specific alterations of gamma-aminobutyric acid and glutamate in patients with major depression. Arch. Gen. Psychiatry 61: 705–713. Sanacora, G., G.F. Mason, D.L. Rothman & J.H. Krystal. 2002. Increased occipital cortex GABA concentrations in depressed patients after therapy with selective serotonin reuptake inhibitors. Am. J. Psychiatry 159: 663–665. Sanacora, G., G.F. Mason, D.L. Rothman, et al. 2003. Increased cortical GABA concentrations in depressed patients receiving ECT. Am. J. Psychiatry 160: 577–579. Gabbay, V., X. Mao, R.G. Klein, et al. 2012. Anterior cingulate cortex ␥ -aminobutyric acid in depressed adolescents: relationship to anhedonia. Arch. Gen. Psychiatry 69: 139– 149. Hasler, G., A. Neumeister, J.W. vander Veen, et al. 2005. Normal prefrontal gamma-aminobutyric acid levels in remitted depressed subjects determined by proton magnetic resonance spectroscopy. Biol. Psychiatry 58: 969– 973. ¨ ur. 2010. Magnetic resonance specY¨uksel, C. & D. Ong¨ troscopy studies of glutamate-related abnormalities in mood disorders. Biol. Psychiatry 68: 785–794. Luykx, J.J., K.G. Laban, M.P. vanden Heuvel, et al. 2012. Region and state specific glutamate downregulation in major depressive disorder: a meta-analysis of (1)H-MRS findings. Neurosci. Biobehav. Rev. 36: 198–205. Berman, R.M., A. Cappiello, A. Anand, et al. 2000. Antidepressant effects of ketamine in depressed patients. Biol. Psychiatry 47: 351–354. Murrough, J.W., D.V. Iosifescu, L.C. Chang, et al. 2013. Antidepressant efficacy of ketamine in treatment-resistant

159.

160.

161.

162.

163.

164.

165.

166.

167.

168.

169.

170.

171.

172.

173.

174.

major depression: a two-site randomized controlled trial. Am. J. Psychiatry 170: 1134–1142. McGirr, A., M.T. Berlim, D.J. Bond, et al. 2015. A systematic review and meta-analysis of randomized, double-blind, placebo-controlled trials of ketamine in the rapid treatment of major depressive episodes. Psychol Med. 45: 693–704. Stone, J.M., C. Dietrich, R. Edden, et al. 2012. Ketamine effects on brain GABA and glutamate levels with 1H-MRS: relationship to ketamine-induced psychopathology. Mol. Psychiatry 17: 664–665. Dantzer, R., J.C. O’Connor, G.G. Freund, et al. 2008. From inflammation to sickness and depression: when the immune system subjugates the brain. Nat. Rev. Neurosci. 9: 46–56. Dantzer, R., J.C. O’Connor, M.A. Lawson & K.W. Kelley. 2011. Inflammation-associated depression: from serotonin to kynurenine. Psychoneuroendocrinology 36: 426–436. Lopresti, A.L., G.L. Maker, S.D. Hood & P.D. Drummond. 2014. A review of peripheral biomarkers in major depression: the potential of inflammatory and oxidative stress biomarkers. Prog. Neuropsychopharmacol. Biol. Psychiatry 48: 102–111. Berk, M., F. Ng, O. Dean, et al. 2008. Glutathione: a novel treatment target in psychiatry. Trends Pharmacol. Sci. 29: 346–351. Berk, M., O.M. Dean, S.M. Cotton, et al. 2014. The efficacy of adjunctive N-acetylcysteine in major depressive disorder: a double-blind, randomized, placebo-controlled trial. J. Clin. Psychiatry 75: 628–636. ´ Kaiser, L.G., M. Marjanska, G.B. Matson, et al. 2010. (1)H MRS detection of glycine residue of reduced glutathione in vivo. J. Magn. Reson. 202: 259–266. Shungu, D.C., N. Weiduschat, J.W. Murrough, et al. 2012. Increased ventricular lactate in chronic fatigue syndrome. III. Relationships to cortical glutathione and clinical symptoms implicate oxidative stress in disorder pathophysiology. NMR Biomed. 25: 1073–1087. Godlewska, B.R., J. Near & P.J. Cowen. 2015. Neurochemistry of major depression: a study using magnetic resonance spectroscopy. Psychopharmacology (Berl). 232: 501–507. Lapidus, K.A., V. Gabbay, X. Mao, et al. 2014. In vivo (1)H MRS study of potential associations between glutathione, oxidative stress and anhedonia in major depressive disorder. Neurosci. Lett. 569: 74–79. Volz, H.P., et al. 1998. 31P magnetic resonance spectroscopy in the frontal lobe of major depressed patients. Eur. Arch. Psychiatry Clin. Neurosci. 248: 289–295. Moore, C.M., et al., 1997. Lower levels of nucleoside triphosphate in the basal ganglia of depressed subjects: a phosphorous-31 magnetic resonance spectroscopy study. Am. J. Psychiatry 154: 116–118. Renshaw, P.F., et al. 2001. Multinuclear magnetic resonance spectroscopy studies of brain purines in major depression. Am. J. Psychiatry 158: 2048–2055. Iosifescu, D.V., et al. 2008. Brain bioenergetics and response to triiodothyronine augmentation in major depressive disorder. Biol. Psychiatry 63: 1127–1134. Frodl, T., C. Sch¨ule, G. Schmitt, et al. 2007. Association of the brain-derived neurotrophic factor Val66Met poly-

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

Lener & Iosifescu

morphism with reduced hippocampal volumes in major depression. Arch. Gen. Psychiatry 64: 410–416. 175. Forde, N.J., L. Ronan, J. Suckling, et al. 2014. Structural neuroimaging correlates of allelic variation of the BDNF val66met polymorphism. Neuroimage 90: 280– 289. 176. Hayashi, K., R. Yoshimura, S. Kakeda, et al. 2014. COMT Val158Met, but not BDNF Val66Met, is associated with

Neuroimaging biomarkers in MDD

white matter abnormalities of the temporal lobe in patients with first-episode, treatment-na¨ıve major depressive disorder: a diffusion tensor imaging study. Neuropsychiatr. Dis. Treat. 10: 1183–1190. 177. Seok, J.H., S. Choi, H.K. Lim, et al. 2013. Effect of the COMT val158met polymorphism on white matter connectivity in patients with major depressive disorder. Neurosci. Lett. 545: 35–39.

C 2015 New York Academy of Sciences. Ann. N.Y. Acad. Sci. 1344 (2015) 50–65 

65

In pursuit of neuroimaging biomarkers to guide treatment selection in major depressive disorder: a review of the literature.

Over the last few decades, neuroimaging techniques have advanced the identification of structural, functional, and neurochemical brain abnormalities t...
299KB Sizes 0 Downloads 8 Views