Building better biomarkers: brain models in translational neuroimaging

Choong-Wan Woo, Luke J Chang, Martin A Lindquist, Tor D Wager, Choong-Wan Woo, Luke J Chang, Martin A Lindquist, Tor D Wager

Abstract

Despite its great promise, neuroimaging has yet to substantially impact clinical practice and public health. However, a developing synergy between emerging analysis techniques and data-sharing initiatives has the potential to transform the role of neuroimaging in clinical applications. We review the state of translational neuroimaging and outline an approach to developing brain signatures that can be shared, tested in multiple contexts and applied in clinical settings. The approach rests on three pillars: (i) the use of multivariate pattern-recognition techniques to develop brain signatures for clinical outcomes and relevant mental processes; (ii) assessment and optimization of their diagnostic value; and (iii) a program of broad exploration followed by increasingly rigorous assessment of generalizability across samples, research contexts and populations. Increasingly sophisticated models based on these principles will help to overcome some of the obstacles on the road from basic neuroscience to better health and will ultimately serve both basic and applied goals.

Figures

Figure 5
Figure 5
Varieties of predictive models. Developing a predictive model entails making choices about its input data, structural properties, and level of analysis. Five of the most important choices are discussed in Box 1.
Figure 1
Figure 1
Standard mapping versus predictive modeling. (a) Traditional brain mapping, often called mass-univariate analysis or voxelwise encoding model. Brain maps are constructed by conducting massive number of tests on brain voxels one at a time. (b) An example showing small effect sizes (here, explained variance) when one brain region is considered in isolation and larger effect sizes for a multivariate model. Chang et al. showed that local regions, including amygdala, anterior cingulate cortex (ACC), insula or searchlights, explained much less variance in experienced negative emotion than a whole-brain predictive model. (c) Predictive modeling explicitly aims to develop brain models that are tightly coupled with target outcomes. w1, w2, … wn represent predictive weights across voxels. (d) Predictive model development and prospective testing. Here, a predictive map (w⃗) comprised of predictive weights across voxels is developed based on a training sample (i.e., a group of individuals) and tested on independent test samples (i.e., new individuals). The weights specify how to integrate brain data to produce a single prediction about the outcome, which could be continuous or categorical. In this example, calculating the dot product between the predictive map and the test images—a weighted sum of activity across the test image (β⃗), with the predictive map specifying the weights (w⃗)—generates a predicted outcome for each participant. The sensitivity, specificity and other properties of the predictive map are estimated from test samples. Data in b from Chang et al..
Figure 2
Figure 2
A snapshot of translational neuroimaging using multivariate predictive models. We searched PubMed for original neuroimaging research articles (including EEG, positron-emission topography (PET), MRI, diffusion tensor imaging (DTI) and arterial spin labeling (ASL)) published between 1983 and January 2016. The search terms can be found via this link: http://goo.gl/N7oh0i. Nonhuman and nonclinical studies were excluded, as well as those that did not employ multivariate pattern recognition. The initial search yielded 2,767 studies, of which 536 studies were selected based on review of their abstracts. Full-text review was used to select 475 studies that included 615 classification or predictive maps. (a) Top: growth of pattern recognition studies in translational neuroimaging since 2004. Bottom: growth of sample sizes in translational neuroimaging studies. The y-axis shows the largest sample size among studies published each year. (b) Breakdown of studies by diagnostic category. PTSD, post-traumatic stress disorder. ADHD, attention deficit hyperactivity disorder. (c) Uses of pattern recognition models. ‘Diagnosis’ refers to patient vs. control classification and ‘risk group’ to classification of groups at high risk (for example, relatives of people with disorders) vs. controls. ‘Symptom’ refers to prediction of continuous symptom scores. ‘Subtype’ refers to identification of subgroups of patients based on brain patterns. ‘Prognosis’ and ‘treatment response’ refer to predictions of individual differences in disease progression and response to an intervention, respectively. ‘Component process’ studies identify predictive models for basic cognitive or affective processes and apply those to classifying patient groups or to predicting symptoms in patients. (d) Precision-weighted accuracy, based on the square root of the sample size, for patient vs. control classification in model-development samples. Here we show classification accuracy only for patient vs. control classification, which was the most common use across disorders (75% of predictive models). The size of the circles shows the precision estimates, with larger circles indicating larger samples and more precise estimates. Accuracy was nearly always estimated using cross-validation. (e) Classification results from prospective testing on independent data sets. Only a small minority of studies report prospective tests. Lower accuracy in independent tests is indicative of bias in cross-validated accuracy estimates from training samples. Accuracy is lower in most cases reviewed here, with AD classification showing least evidence for bias. (f) Diagnostic classification accuracy as a function of sample size for six types of disorders. As the estimates from the largest studies are the most precise, they are most representative of the true accuracy. Across disorders, very high classification accuracy is reported in some small studies, but these have not been replicated in prospective tests. With a few exceptions, accuracy values for large-sample studies are much more modest. These observations point to the need for improvements in statistical model development, data aggregation and prospective testing of promising models across multiple, diverse samples.
Figure 3
Figure 3
Brain signature development and validation. (a) In this process, broad exploration is the first step. Just as drug development involves screening many candidate drugs, exploring multiple approaches and models is important for identifying promising biomarkers. The most promising models must be tested in independent samples to demonstrate their diagnostic accuracy. During characterization, promising candidate biomarkers should show robust replications of findings (for example, high sensitivity and specificity) across multiple independent samples, laboratories, scanners and research settings. This requires tests in larger, more definitive studies, which can eventually promote identification of these biomarkers as surrogate measures and as endpoints in their own right. (b) Starting with many candidate models, the most promising ones garner support and are carried forward with increasing levels of evidence. In the development phase, models can be developed based on one study sample and model performance can be estimated using cross-validation. In the prospective validation phase, findings and model performance (i.e., sensitivity, specificity and predictive value) are replicated by applying models to new, independent samples of participants. In the generalization phase, findings and model performance are tested across multiple laboratories, scanners and variants of testing procedures to assess the models’ robustness and boundary conditions. In the population-level phase, large-scale tests assess the model’s performance when it is applied to diverse populations and test conditions, and additional moderators (i.e., age, race, culture, gender) and boundary conditions are identified in this phase. In this illustration, colored blocks denote the different phases, and black lines indicate hypothetical candidate models. Within each phase, a model can be more or less thoroughly evaluated and more or less successful at establishing utility. This variability is denoted graphically by the variable locations of dots for each model within each phase. A survey of empirical literature to date reveals that only 9% of neuroimaging-based models go beyond the initial development phase. Some notable exceptions include the named models shown here, which have been tested prospectively on new samples (see Table 2 for details and abbreviations).
Figure 4
Figure 4
Future directions. (a) The direct prediction approach, in which brain features are mapped to clinical diagnostic categories or symptoms directly. (b) The component process approach, in which brain features are mapped to basic component processes, such as sustained attention load, memory load, positive or negative affect, pain and others. By introducing this additional layer, brain signatures can have implications for behavior and function beyond patient status or current diagnostic categories. This new level of analysis provides a way of understanding the nature of dysregulated brain processes, assessing risk factors for brain disorders, and understanding and predicting treatment responses. Rather than constructing one brain marker per disorder, component models provide a set of basis processes that are combined in different ways in different disorders. By analogy with color, three components (red, green and blue) can be combined in different ways to form a virtually infinite number of colors. (c) The NPS is a signature for one such component process, evoked somatic pain, that is potentially dysregulated in multiple disorders. The NPS is defined by brain-wide, mesoscale patterns of fMRI activity across multiple pain-related regions and can be prospectively tested on new individuals and datasets. This allows its properties to be characterized across studies, improving understanding of the types of mental processes and experiences it represents. Top: the NPS pattern map thresholded at q < 0.05, false discovery rate (FDR)-corrected for display purposes; the unthresholded patterns in selected regions (dACC, dorsal ACC; dorsal posterior insula, dpINS; secondary somatosensory cortex, S2) are visualized in the insets. Bottom: the NPS’s ‘psychological receptive field’, which visualizes conditions that activate (sensitivity, in orange and red) or do not activate (specificity, in gray and black) the NPS. Dark colored conditions (in red and black) are from published results,,–, and light colored conditions (in orange and gray) are from unpublished, preliminary results (data on cognitive demand were tested by C.-W.W.; visceral and vaginal pain data were tested by T.D.W.). Characterizing the NPS’s sensitivity and specificity across these conditions and others aides in understanding what NPS alterations in clinical disorders mean from a psychological and functional perspective. In the future, we envision clinical biomarkers composed of sets of interpretable, well characterized models of basic cognitive and affective processes.

References

    1. Mather M, Cacioppo JT, Kanwisher N. Introduction to the special section: 20 years of fMRI-what has it done for understanding cognition? Perspect Psychol Sci. 2013;8:41–43.
    1. Kapur S, Phillips AG, Insel TR. Why has it taken so long for biological psychiatry to develop clinical tests and what to do about it? Mol Psychiatry. 2012;17:1174–1179.
    1. Mayberg HS, et al. Reciprocal limbic-cortical function and negative mood: converging PET findings in depression and normal sadness. Am J Psychiatry. 1999;156:675–682.
    1. Keedwell PA, Andrew C, Williams SC, Brammer MJ, Phillips ML. The neural correlates of anhedonia in major depressive disorder. Biol Psychiatry. 2005;58:843–853.
    1. Tom SM, Fox CR, Trepel C, Poldrack RA. The neural basis of loss aversion in decision-making under risk. Science. 2007;315:515–518.
    1. Rosenberg MD, et al. A neuromarker of sustained attention from whole-brain functional connectivity. Nat Neurosci. 2016;19:165–171.
    1. Sanislow CA, et al. Developing constructs for psychopathology research: research domain criteria. J Abnorm Psychol. 2010;119:631–639.
    1. Scoville WB, Milner B. Loss of recent memory after bilateral hippocampal lesions. J Neurol Neurosurg Psychiatry. 1957;20:11–21.
    1. Fodor JA. The Modularity of Mind. MIT Press; 1983.
    1. Hamani C, et al. Deep brain stimulation for chronic neuropathic pain: long-term outcome and the incidence of insertional effect. Pain. 2006;125:188–196.
    1. Welter ML, et al. Basal ganglia dysfunction in OCD: subthalamic neuronal activity correlates with symptoms severity and predicts high-frequency stimulation efficacy. Transl Psychiatry. 2011;1:e5.
    1. Krack P, et al. Five-year follow-up of bilateral stimulation of the subthalamic nucleus in advanced Parkinson’s disease. N Engl J Med. 2003;349:1925–1934.
    1. Swartz JR, Knodt AR, Radtke SR, Hariri AR. A neural biomarker of psychological vulnerability to future life stress. Neuron. 2015;85:505–511.
    1. Dougherty DD, et al. A randomized sham-controlled trial of deep brain stimulation of the ventral capsule/ventral striatum for chronic treatment-resistant depression. Biol Psychiatry. 2015;78:240–248.
    1. Morishita T, Fayad SM, Higuchi MA, Nestor KA, Foote KD. Deep brain stimulation for treatment-resistant depression: systematic review of clinical outcomes. Neurotherapeutics. 2014;11:475–484.
    1. Reddan MC, Lindquist MA, Wager TD. Effect size estimation in neuroimaging. JAMA Psychiatry. 2017:3356. .
    1. Logothetis NK. What we can do and what we cannot do with fMRI. Nature. 2008;453:869–878.
    1. Kvitsiani D, et al. Distinct behavioural and network correlates of two interneuron types in prefrontal cortex. Nature. 2013;498:363–366.
    1. Price JL, Drevets WC. Neural circuits underlying the pathophysiology of mood disorders. Trends Cogn Sci. 2012;16:61–71.
    1. Roy M, Shohamy D, Wager TD. Ventromedial prefrontal-subcortical systems and the generation of affective meaning. Trends Cogn Sci. 2012;16:147–156.
    1. Wager TD, et al. Pain in the ACC? Proc Natl Acad Sci USA. 2016;113:E2474–E2475.
    1. Poldrack RA. Can cognitive processes be inferred from neuroimaging data? Trends Cogn Sci. 2006;10:59–63.
    1. Wager TD, et al. An fMRI-based neurologic signature of physical pain. N Engl J Med. 2013;368:1388–1397.
    1. Chang LJ, Gianaros PJ, Manuck SB, Krishnan A, Wager TD. A sensitive and specific neural signature for picture-induced negative affect. PLoS Biol. 2015;13:e1002180.
    1. Doyle OM, Mehta MA, Brammer MJ. The role of machine learning in neuroimaging for drug discovery and development. Psychopharmacology (Berl ) 2015;232:4179–4189.
    1. Haynes JD. A primer on pattern-based approaches to fMRI: principles, pitfalls, and perspectives. Neuron. 2015;87:257–270.
    1. Orrù G, Pettersson-Yeo W, Marquand AF, Sartori G, Mechelli A. Using Support Vector Machine to identify imaging biomarkers of neurological and psychiatric disease: a critical review. Neurosci Biobehav Rev. 2012;36:1140–1152.
    1. Hackmack K, Paul F, Weygandt M, Allefeld C, Haynes JD. Multi-scale classification of disease using structural MRI and wavelet transform. Neuroimage. 2012;62:48–58.
    1. Miyawaki Y, et al. Visual image reconstruction from human brain activity using a combination of multiscale local image decoders. Neuron. 2008;60:915–929.
    1. Kamitani Y, Tong F. Decoding the visual and subjective contents of the human brain. Nat Neurosci. 2005;8:679–685.
    1. Kriegeskorte N, Cusack R, Bandettini P. How does an fMRI voxel sample the neuronal activity pattern: compact-kernel or complex spatiotemporal filter? Neuroimage. 2010;49:1965–1976.
    1. Poldrack RA, Gorgolewski KJ. Making big data open: data sharing in neuroimaging. Nat Neurosci. 2014;17:1510–1517.
    1. Abi-Dargham A, Horga G. The search for imaging biomarkers in psychiatric disorders. Nat Med. 2016;22:1248–1255.
    1. Hastie T, Tibshirani R, Friedman JH. The Elements of Statistical Learning: Data Mining, Inference, and Prediction. 2. Springer; 2009.
    1. Mohri M, Rostamizadeh A, Talwalkar A. Foundations of Machine Learning. MIT Press; 2012.
    1. de Leon MJ, et al. Positron emission tomographic studies of aging and Alzheimer disease. AJNR Am J Neuroradiol. 1983;4:568–571.
    1. Kippenhan JS, Barker WW, Pascal S, Nagel J, Duara R. Evaluation of a neural-network classifier for PET scans of normal and Alzheimer’s disease subjects. J Nucl Med. 1992;33:1459–1467.
    1. Doyle OM, et al. Predicting progression of Alzheimer’s disease using ordinal regression. PLoS One. 2014;9:e105542.
    1. Singh G, Samavedham L. Unsupervised learning based feature extraction for differential diagnosis of neurodegenerative diseases: A case study on early-stage diagnosis of Parkinson disease. J Neurosci Methods. 2015;256:30–40.
    1. Koutsouleris N, et al. Use of neuroanatomical pattern classification to identify subjects in at-risk mental states of psychosis and predict disease transition. Arch Gen Psychiatry. 2009;66:700–712.
    1. Sørensen L, et al. Early detection of Alzheimer’s disease using MRI hippocampal texture. Hum Brain Mapp. 2016;37:1148–1161.
    1. Moradi E, Pepe A, Gaser C, Huttunen H, Tohka J. Machine learning framework for early MRI-based Alzheimer’s conversion prediction in MCI subjects. Neuroimage. 2015;104:398–412.
    1. Beardslee WR, et al. Prevention of depression in at-risk adolescents: longer-term effects. JAMA Psychiatry. 2013;70:1161–1170.
    1. Addington J, Heinssen R. Prediction and prevention of psychosis in youth at clinical high risk. Annu Rev Clin Psychol. 2012;8:269–289.
    1. Davatzikos C, Xu F, An Y, Fan Y, Resnick SM. Longitudinal progression of Alzheimer’s-like patterns of atrophy in normal older adults: the SPARE-AD index. Brain. 2009;132:2026–2035.
    1. Fan Y, Batmanghelich N, Clark CM, Davatzikos C Alzheimer’s Disease Neuroimaging Initiative\par. Spatial patterns of brain atrophy in MCI patients, identified via high-dimensional pattern classification, predict subsequent cognitive decline. Neuroimage. 2008;39:1731–1743.
    1. Misra C, Fan Y, Davatzikos C. Baseline and longitudinal patterns of brain atrophy in MCI patients, and their use in prediction of short-term conversion to AD: results from ADNI. Neuroimage. 2009;44:1415–1422.
    1. Tang CC, et al. Differential diagnosis of Parkinsonism: a metabolic imaging study using pattern analysis. Lancet Neurol. 2010;9:149–158.
    1. Pantazatos SP, Talati A, Schneier FR, Hirsch J. Reduced anterior temporal and hippocampal functional connectivity during face processing discriminates individuals with social anxiety disorder from healthy controls and panic disorder, and increases following treatment. Neuropsychopharmacology. 2014;39:425–434.
    1. Anticevic A, et al. Characterizing thalamo-cortical disturbances in schizophrenia and bipolar illness. Cereb Cortex. 2014;24:3116–3130.
    1. Calhoun VD, Maciejewski PK, Pearlson GD, Kiehl KA. Temporal lobe and “default” hemodynamic brain modes discriminate between schizophrenia and bipolar disorder. Hum Brain Mapp. 2008;29:1265–1275.
    1. Insel TR, Cuthbert BN. Medicine. Brain disorders? Precisely. Science. 2015;348:499–500.
    1. Clementz BA, et al. Identification of distinct psychosis biotypes using brain-based biomarkers. Am J Psychiatry. 2016;173:373–384.
    1. Price RB, et al. Parsing heterogeneity in the brain connectivity of depressed and healthy adults during positive mood. Biol Psychiatry. 2016 S0006-3223(16)32540-9.
    1. Drysdale AT, et al. Resting-state connectivity biomarkers define neurophysiological subtypes of depression. Nat Med. 2016
    1. Weinstein JN, et al. The Cancer Genome Atlas Pan-Cancer analysis project. Nat Genet. 2013;45:1113–1120.
    1. Roychowdhury S, Chinnaiyan AM. Translating genomics for precision cancer medicine. Annu Rev Genomics Hum Genet. 2014;15:395–415.
    1. Hahn T, et al. Predicting treatment response to cognitive behavioral therapy in panic disorder with agoraphobia by integrating local neural information. JAMA Psychiatry. 2015;72:68–74.
    1. Doehrmann O, et al. Predicting treatment response in social anxiety disorder from functional magnetic resonance imaging. JAMA Psychiatry. 2013;70:87–97.
    1. Whitfield-Gabrieli S, et al. Brain connectomics predict response to treatment in social anxiety disorder. Mol Psychiatry. 2016;21:680–685.
    1. van Waarde JA, et al. A functional MRI marker may predict the outcome of electroconvulsive therapy in severe and treatment-resistant depression. Mol Psychiatry. 2015;20:609–614.
    1. Widge AS, Avery DH, Zarkowski P. Baseline and treatment-emergent EEG biomarkers of antidepressant medication response do not predict response to repetitive transcranial magnetic stimulation. Brain Stimul. 2013;6:929–931.
    1. Sarpal DK, et al. Baseline striatal functional connectivity as a predictor of response to antipsychotic drug treatment. Am J Psychiatry. 2016;173:69–77.
    1. Ye Z, et al. Predicting beneficial effects of atomoxetine and citalopram on response inhibition in Parkinson’s disease with clinical and neuroimaging measures. Hum Brain Mapp. 2016;37:1026–1037.
    1. Woo CW, Wager TD. Neuroimaging-based biomarker discovery and validation. Pain. 2015;156:1379–1381.
    1. Robinson M, Boissoneault J, Sevel L, Letzen J, Staud R. The effect of base rate on the predictive value of brain biomarkers. J Pain. 2016;17:637–641.
    1. Cronbach LJ, Meehl PE. Construct validity in psychological tests. Psychol Bull. 1955;52:281–302.
    1. Freedman R, et al. The initial field trials of DSM-5: new blooms and old thorns. Am J Psychiatry. 2013;170:1–5.
    1. Iidaka T. Resting state functional magnetic resonance imaging and neural network classified autism and control. Cortex. 2015;63:55–67.
    1. Duffy FH, Als H. A stable pattern of EEG spectral coherence distinguishes children with autism from neuro-typical controls - a large case control study. BMC Med. 2012;10:64.
    1. Deshpande G, Wang P, Rangaprakash D, Wilamowski B. Fully connected cascade artificial neural network architecture for attention deficit hyperactivity disorder classification from functional magnetic resonance imaging data. IEEE Trans Cybern. 2015;45:2668–2679.
    1. Whelan R, Garavan H. When optimism hurts: inflated predictions in psychiatric neuroimaging. Biol Psychiatry. 2014;75:746–748.
    1. Zaki J, Wager TD, Singer T, Keysers C, Gazzola V. The anatomy of suffering: understanding the relationship between nociceptive and empathic pain. Trends Cogn Sci. 2016;20:249–259.
    1. Olivetti E, Sona D, Veeramachaneni S. Gaussian process regression and recurrent neural networks for fmri image classification. Proc. 12th Meeting Org. for Human Brain Mapping; Florence, Italy. 2006.
    1. Ribeiro MT, Singh S, Guestrin C. “Why should I trust you?”: Explaining the predictions of any classifier. 2016 Preprint at arXiv .
    1. HD-200 Consortium. The ADHD-200 Consortium: a model to advance the translational potential of neuroimaging in clinical neuroscience. Front Syst Neurosci. 2012;6:62.
    1. Eloyan A, et al. Automated diagnoses of attention deficit hyperactive disorder using magnetic resonance imaging. Front Syst Neurosci. 2012;6:61.
    1. Eldridge J, Lane AE, Belkin M, Dennis S. Robust features for the automatic identification of autism spectrum disorder in children. J Neurodev Disord. 2014;6:12.
    1. Geurts JJ, Calabrese M, Fisher E, Rudick RA. Measurement and clinical effect of grey matter pathology in multiple sclerosis. Lancet Neurol. 2012;11:1082–1092.
    1. van den Heuvel MP, et al. Abnormal rich club organization and functional brain dynamics in schizophrenia. JAMA Psychiatry. 2013;70:783–792.
    1. Yahata N, et al. A small number of abnormal brain connections predicts adult autism spectrum disorder. Nat Commun. 2016;7:11254.
    1. Huth AG, de Heer WA, Griffiths TL, Theunissen FE, Gallant JL. Natural speech reveals the semantic maps that tile human cerebral cortex. Nature. 2016;532:453–458.
    1. Vemuri P, et al. Antemortem MRI based STructural Abnormality iNDex (STAND)-scores correlate with postmortem Braak neurofibrillary tangle stage. Neuroimage. 2008;42:559–567.
    1. Yarkoni T, Poldrack RA, Nichols TE, Van Essen DC, Wager TD. Large-scale automated synthesis of human functional neuroimaging data. Nat Methods. 2011;8:665–670.
    1. Yeo BT, et al. The organization of the human cerebral cortex estimated by intrinsic functional connectivity. J Neurophysiol. 2011;106:1125–1165.
    1. Glasser MF, et al. A multi-modal parcellation of human cerebral cortex. Nature. 2016;536:171–178.
    1. Bota M, Dong HW, Swanson LW. Brain architecture management system. Neuroinformatics. 2005;3:15–48.
    1. Stephan KE. The history of CoCoMac. Neuroimage. 2013;80:46–52.
    1. Power JD, Schlaggar BL, Petersen SE. Recent progress and outstanding issues in motion correction in resting state fMRI. Neuroimage. 2015;105:536–551.
    1. Gorgolewski KJ, Poldrack RA. A practical guide for improving transparency and reproducibility in neuroimaging research. PLoS Biol. 2016;14:e1002506.
    1. Davatzikos C, Bhatt P, Shaw LM, Batmanghelich KN, Trojanowski JQ. Prediction of MCI to AD conversion, via MRI, CSF biomarkers, and pattern classification. Neurobiol Aging. 2011;32:2322.e19–2322.e27.
    1. Weintraub D, et al. Alzheimer’s disease pattern of brain atrophy predicts cognitive decline in Parkinson’s disease. Brain. 2012;135:170–180.
    1. Toledo JB, et al. Memory, executive, and multidomain subtle cognitive impairment: clinical and biomarker findings. Neurology. 2015;85:144–153.
    1. Habes M, et al. White matter hyperintensities and imaging patterns of brain ageing in the general population. Brain. 2016;139:1164–1179.
    1. Asanuma K, et al. Network modulation in the treatment of Parkinson’s disease. Brain. 2006;129:2667–2678.
    1. Eidelberg D. Metabolic brain networks in neurodegenerative disorders: a functional imaging approach. Trends Neurosci. 2009;32:548–557.
    1. Wu P, et al. Metabolic brain network in the Chinese patients with Parkinson’s disease based on 18F-FDG PET imaging. Parkinsonism Relat Disord. 2013;19:622–627.
    1. Teune LK, et al. Validation of parkinsonian disease-related metabolic brain patterns. Mov Disord. 2013;28:547–551.
    1. Westfall J, Judd CM, Kenny DA. Replicating studies in which samples of participants respond to samples of stimuli. Perspect Psychol Sci. 2015;10:390–399.
    1. Hashmi JA, et al. Shape shifting pain: chronification of back pain shifts brain representation from nociceptive to emotional circuits. Brain. 2013;136:2751–2768.
    1. Petersen SE, et al. Imaging in population science: cardiovascular magnetic resonance in 100,000 participants of UK Biobank - rationale, challenges and approaches. J Cardiovasc Magn Reson. 2013;15:46.
    1. Weiner MW, et al. Impact of the Alzheimer’s Disease Neuroimaging Initiative, 2004 to 2014. Alzheimers Dement. 2015;11:865–884.
    1. Tagliazucchi E, Laufs H. Decoding wakefulness levels from typical fMRI resting-state data reveals reliable drifts between wakefulness and sleep. Neuron. 2014;82:695–708.
    1. Buckner RL, Krienen FM, Yeo BTT. Opportunities and limitations of intrinsic functional connectivity MRI. Nat Neurosci. 2013;16:832–837.
    1. Glover GH, et al. Function biomedical informatics research network recommendations for prospective multicenter functional MRI studies. J Magn Reson Imaging. 2012;36:39–54.
    1. Landis JR, et al. The MAPP research network: design, patient characterization and operations. BMC Urol. 2014;14:58.
    1. Thompson PM, et al. The ENIGMA Consortium: large-scale collaborative analyses of neuroimaging and genetic data. Brain Imaging Behav. 2014;8:153–182.
    1. Borsook D, Becerra L, Hargreaves R. Biomarkers for chronic pain and analgesia. Part 1: the need, reality, challenges, and solutions. Discov Med. 2011;11:197–207.
    1. Hargreaves RJ, et al. Optimizing central nervous system drug development using molecular imaging. Clin Pharmacol Ther. 2015;98:47–60.
    1. López-Solà M, et al. Towards a neurophysiological signature for fibromyalgia. Pain. 2016
    1. Lombardo MV, et al. Different functional neural substrates for good and poor language outcome in autism. Neuron. 2015;86:567–577.
    1. Woolf CJ, Salter MW. Neuronal plasticity: increasing the gain in pain. Science. 2000;288:1765–1769.
    1. Diatchenko L, Nackley AG, Slade GD, Fillingim RB, Maixner W. Idiopathic pain disorders--pathways of vulnerability. Pain. 2006;123:226–230.
    1. Adler G, Gattaz WF. Pain perception threshold in major depression. Biol Psychiatry. 1993;34:687–689.
    1. Krishnan A, et al. Somatic and vicarious pain are represented by dissociable multivariate brain patterns. eLife. 2016;5:e15166.
    1. Woo CW, Roy M, Buhle JT, Wager TD. Distinct brain systems mediate the effects of nociceptive input and self-regulation on pain. PLoS Biol. 2015;13:e1002036.
    1. Ma Y, et al. Serotonin transporter polymorphism alters citalopram effects on human pain responses to physical pain. Neuroimage. 2016;135:186–196.
    1. Bräscher AK, Becker S, Hoeppli ME, Schweinhardt P. Different brain circuitries mediating controllable and uncontrollable pain. J Neurosci. 2016;36:5013–5025.
    1. Wiecki TV, Poland J, Frank MJ. Model-based cognitive neuroscience approaches to computational psychiatry: clustering and classification. Clin Psychol Sci. 2015;3:378–399.
    1. Huys QJ, Maia TV, Frank MJ. Computational psychiatry as a bridge from neuroscience to clinical applications. Nat Neurosci. 2016;19:404–413.
    1. Brodersen KH, et al. Generative embedding for model-based classification of fMRI data. PLOS Comput Biol. 2011;7:e1002079.
    1. Friston KJ, Harrison L, Penny W. Dynamic causal modelling. Neuroimage. 2003;19:1273–1302.
    1. Hein G, Morishima Y, Leiberg S, Sul S, Fehr E. The brain’s functional network architecture reveals human motives. Science. 2016;351:1074–1078.
    1. Fan Y, Resnick SM, Wu X, Davatzikos C. Structural and functional biomarkers of prodromal Alzheimer’s disease: a high-dimensional pattern classification study. Neuroimage. 2008;41:277–285.
    1. Casanova R, et al. Alzheimer’s disease risk assessment using large-scale machine learning methods. PLoS One. 2013;8:e77949.
    1. Tosun D, Joshi S, Weiner MW. Neuroimaging predictors of brain amyloidosis in mild cognitive impairment. Ann Neurol. 2013;74:188–198.
    1. Vemuri P, et al. Alzheimer’s disease diagnosis in individual subjects using structural MR images: validation studies. Neuroimage. 2008;39:1186–1197.
    1. Huang C, et al. Metabolic brain networks associated with cognitive function in Parkinson’s disease. Neuroimage. 2007;34:714–723.
    1. Mure H, et al. Parkinson’s disease tremor-related metabolic network: characterization, progression, and treatment effects. Neuroimage. 2011;54:1244–1253.
    1. Eckert T, et al. Abnormal metabolic networks in atypical parkinsonism. Mov Disord. 2008;23:727–733.
    1. Niethammer M, et al. A disease-specific metabolic brain network associated with corticobasal degeneration. Brain. 2014;137:3036–3046.
    1. Geman S, Bienenstock E, Doursat R. Neural networks and the bias/variance dilemma. Neural Comput. 1992;4:1–58.
    1. Haxby JV, et al. Distributed and overlapping representations of faces and objects in ventral temporal cortex. Science. 2001;293:2425–2430.
    1. Kriegeskorte N, Goebel R, Bandettini P. Information-based functional brain mapping. Proc Natl Acad Sci USA. 2006;103:3863–3868.
    1. Sato JR, et al. Machine learning algorithm accurately detects fMRI signature of vulnerability to major depression. Psychiatry Res. 2015;233:289–291.
    1. Wager TD, Atlas LY, Leotti LA, Rilling JK. Predicting individual differences in placebo analgesia: contributions of brain activity during anticipation and pain experience. J Neurosci. 2011;31:439–452.
    1. Dukart J, Schroeter ML, Mueller K. Age correction in dementia--matching to a healthy brain. PLoS One. 2011;6:e22193.
    1. Naselaris T, Kay KN, Nishimoto S, Gallant JL. Encoding and decoding in fMRI. Neuroimage. 2011;56:400–410.
    1. Mitchell TM, et al. Predicting human brain activity associated with the meanings of nouns. Science. 2008;320:1191–1195.
    1. Krishnan A, Williams LJ, McIntosh AR, Abdi H. Partial Least Squares (PLS) methods for neuroimaging: a tutorial and review. Neuroimage. 2011;56:455–475.
    1. Nishimoto S, et al. Reconstructing visual experiences from brain activity evoked by natural movies. Curr Biol. 2011;21:1641–1646.
    1. Kay KN, Naselaris T, Prenger RJ, Gallant JL. Identifying natural images from human brain activity. Nature. 2008;452:352–355.
    1. Ketz N, O’Reilly RC, Curran T. Classification aided analysis of oscillatory signatures in controlled retrieval. Neuroimage. 2014;85:749–760.
    1. Kim J, Calhoun VD, Shim E, Lee JH. Deep neural network with weight sparsity control and pre-training extracts hierarchical features and enhances classification performance: Evidence from whole-brain resting-state functional connectivity patterns of schizophrenia. Neuroimage. 2016;124(Pt A):127–146.
    1. Kriegeskorte N. Deep neural networks: a new framework for modeling biological vision and brain information processing. Annual Review of Vision Science. 2015;1:417–446.
    1. O’Reilly RC. Biologically based computational models of high-level cognition. Science. 2006;314:91–94.
    1. Poldrack RA, Halchenko YO, Hanson SJ. Decoding the large-scale structure of brain function by classifying mental States across individuals. Psychol Sci. 2009;20:1364–1372.
    1. Todd MT, Nystrom LE, Cohen JD. Confounds in multivariate pattern analysis: Theory and rule representation case study. Neuroimage. 2013;77:157–165.
    1. Etzel JA, Zacks JM, Braver TS. Searchlight analysis: promise, pitfalls, and potential. Neuroimage. 2013;78:261–269.
    1. Haxby JV, et al. A common, high-dimensional model of the representational space in human ventral temporal cortex. Neuron. 2011;72:404–416.

Source: PubMed

3
Subskrybuj