Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-23T12:29:10.246Z Has data issue: false hasContentIssue false

The potential of precision psychiatry: what is in reach?

Published online by Cambridge University Press:  31 March 2022

Lana Kambeitz-Ilankovic*
Affiliation:
Department of Psychiatry and Psychotherapy, Faculty of Medicine and University Hospital of Cologne, Germany; and Department of Psychiatry and Psychotherapy, Ludwig-Maximilian University, Munich, Germany
Nikolaos Koutsouleris
Affiliation:
Department of Psychiatry and Psychotherapy, Ludwig-Maximilian University, Munich, Germany; Max-Planck Institute of Psychiatry, Munich, Germany; and Institute of Psychiatry, Psychology and Neurosciences, King's College London, London, UK
Rachel Upthegrove
Affiliation:
Early Intervention Service, Birmingham Women's and Children's NHS Foundation Trust, UK; and Institute for Mental Health, University of Birmingham, UK
*
Correspondence: Lana Kambeitz-Ilankovic. Email: lana.kambeitz-ilankovic@uk-koeln.de
Rights & Permissions [Opens in a new window]

Summary

Progress in developing personalised care for mental disorders is supported by numerous proof-of-concept machine learning studies in the area of risk assessment, diagnostics and precision prescribing. Most of these studies primarily use clinical data, but models might benefit from additional neuroimaging, blood and genetic data to improve accuracy. Combined, multimodal models might offer potential for stratification of patients for treatment. Clinical implementation of machine learning is impeded by a lack of wider generalisability, with efforts primarily focused on psychosis and dementia. Studies across all diagnostic groups should work to test the robustness of machine learning models, which is an essential first step to clinical implementation, and then move to prospective clinical validation. Models need to exceed clinicians’ heuristics to be useful, and safe, in routine decision-making. Engagement of clinicians, researchers and patients in digitalisation and ‘big data’ approaches are vital to allow the generation and accessibility of large, longitudinal, prospective data needed for precision psychiatry to be applied into real-world psychiatric care.

Type
Analysis
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press on behalf of the Royal College of Psychiatrists

The past decade has seen substantial investment in the field of data science to develop precision healthcare for the treatment and prevention of mental illness. Precision healthcare promises to move away from a ‘one-size-fits-all’ approach to treatment decisions by using objective and replicable psychosocial and/or neurobiological measures. Interventions would ideally be tailored to these individual profiles and address unique characteristics of individual patients by maximising clinical response.Reference Arns, van Dijk H, Luykx, van Wingen and Olbrich1 The drive to embrace data science results from core challenges that psychiatry has not yet been able to resolve: how we define illnesses, develop accurate diagnostic categories, identify biomarkers and predict outcomes, together with how we understand and manage the heterogeneity that is the norm in mental health populations. Finding accurate prediction models and defining meaningful phenotypes or biologically informed groups could be transformative. However, although considerable investment has resulted in progress and several key achievements, caution is clearly still needed. Although a number of machine learning studies to date have potential for real-world application, some are closer to bedside testing than others and many steps are still needed before data-driven precision healthcare is in place to aid everyday clinical care. We present here an analysis of the current position of machine learning research that is closest to real clinical practice, covering prognostic risk, diagnostic stratification and treatment response, with critical insight into current gaps and challenges (Fig. 1).

Fig. 1 Individual prognosis along the disease trajectory.

Black lines indicate fields with stronger translational potential due to a larger number of validation studies; grey lines indicate fields of research with currently less translational perspective, owing to a sparse number of studies and validation attempts. Diff. diagnostics, differential diagnostics.

Prognostic risk

Mental disorders with onset in early adulthood frequently lead to enduring disability.Reference Koutsouleris, Kambeitz-Ilankovic, Ruhrmann, Rosen, Ruef and Dwyer2 Disorders occurring later in life, such as dementia, result in significant burden on family members and institutional care in the last decades of life. Early detection could reduce this burden by enabling increased support and preventive interventions. Recently published machine learning studies bring some hope for this goal, as they show partially generalisable multimodal prognostic models able to predict individual functional outcomes with some accuracy.Reference Koutsouleris, Kambeitz-Ilankovic, Ruhrmann, Rosen, Ruef and Dwyer2 Using algorithmic pattern recognition, this work showed better accuracy than human prognosis. The North American Prodrome Longitudinal Study (NAPLS) individual risk calculator for development of psychosis from the clinical high-risk (CHR) stateReference Cannon, Yu, Addington, Bearden, Cadenhead and Cornblatt3 has been recently validated in more broadly defined CHR groups from multiple countries,Reference Koutsouleris, Worthington, Dwyer, Kambeitz-Ilankovic, Sanfelici and Fusar-Poli4 including patients with recent-onset depression from the PRONIA consortia. This valuable generalisable model points to younger age at onset and reduced cognitive processing speed as increasingly relevant risk markers in broader risk cohorts. Harmonised models from PRONIA and NAPLS are based on a concise pattern of demographic, clinical and neuropsychological variables, in addition to attenuated psychotic symptoms, that can be more easily applied in clinical practice.

Currently, evidence suggests that neurobiological data may add some predictive accuracy to clinical models for risk prediction, yet at present this may not be at the level of significance to warrant everyday use.

Higher specificity sometimes seen in neurobiologically based models, when compared with solely clinical data, may remain important in identifying underlying aetiological processes or new staging, given the heterogeneity in clinical phenomenology.Reference Koutsouleris, Kambeitz-Ilankovic, Ruhrmann, Rosen, Ruef and Dwyer2 The added value of neurobiological data becomes more evident in older patients, for example when predicting fast progression from mild cognitive impairment to Alzheimer's dementia. Cerebrospinal fluid, cerebral amyloid or tau and, in particular, neurodegenerative markers so far prove to be key neurobiological predictorsReference Davatzikos, Bhatt, Shaw, Batmanghelich and Trojanowski5,Reference Varatharajah, Ramanan, Iyer and Vemuri6 that have been validated in a number of muticentre dementia studies.Reference van Maurik7

Genetic data may be similar to neuroimaging data, in that they could improve overall accuracy of models, but are not able to deliver self-standing findings, i.e. at present they only complement clinical data. For over a decade expectations were directed at the level of single candidate genes, for example COMT for schizophrenia or ApoE4 for Alzheimer's dementia, but contemporary research of prediction relies on polygenic risk scores.Reference Murray, Lin, Austin, McGrath, Hickie and Wray8 Recent advances from the Psychiatric Genomics Consortium–UK Biobank–23 and Me genome-wide association study report that polygenic risk scores may be useful for prediction of vulnerability to depression and resilience under stress.Reference Fang, Scott, Song, Burmeister and Sen9 Similarly, studies on prognostic flowsReference Koutsouleris, Dwyer, Degenhardt, Maj, Urquijo-Castro and Sanfelici10 applied in psychotic disorders agree that polygenic risk scores slightly augment the performance of models based on clinical cognitive data, yet remain insufficient for risk screening in the general population.

These multimodal prognostic flows may be extended to cohorts of young adolescents and adults, although a similar caution should be taken. Most recent multisite longitudinal adolescent studies, for example IMAGEN,Reference Toenders, Kottaram, Dinga, Davey, Banaschewski and Bokde11 emphasise the relevance of a risk pattern for depression in adolescence driven by baseline depressive symptoms, female gender, neuroticism and stressful events, accompanied by surface reduction in the supramarginal gyrus. A broad population-based study, the Philadelphia Neurodevelopmental Cohort (PNC),Reference Satterthwaite, Connolly, Ruparel, Calkins, Jackson and Elliott12 pursued clinical and neurobehavioural characterisation of genotyped youths for the prospective emergence of psychiatric illness. The PNC has so far delivered a solid normative ground for cognitive milestones and neural development in children and adolescents from age 8 to 21 years. However, the full potential of sufficiently validated developmental biomarkers identified through epidemiological cohorts is yet to be confirmed, and challenges include the infrequent nature of target outcomes, which mean that very large prospective samples are needed.

Diagnostic classification and stratification

Early supervised machine learning studies were driven by the idea that different diagnostic categories have distinct neurobiological underpinnings that can be used to identify biomarkers for psychiatric diseases, similar to those in physical health conditions.Reference Koutsouleris, Meisenzahl, Davatzikos, Bottlender, Frodl and Scheuerecker13 Long clinical interviews would become obsolete and techniques such as structural magnetic resonance imaging (MRI) would be used to deliver a robust psychiatric diagnosis and facilitate an accurate prognosis and treatment choice. However, less algorithmic precision than initially expected has been achieved, with predictive accuracies in ranges that would fail validation tests.Reference Mechelli and Vieira14 This has led to further scepticism regarding discrete diagnostic categories and also the potential of machine learning methods.Reference Kambeitz, Kambeitz-Ilankovic, Leucht, Wood, Davatzikos and Malchow15 Distinct mental disorders often have many individual symptoms in common, and similarly the majority of neurobiological substrates are present across diagnostic categories. Although diagnostic categories help to conceptualise the high variability of symptoms, we need to be able to accurately stratify patient subgroups on the basis of both reliable clinical and relevant biomarker data to foresee clinical outcome and facilitate the development of selective and indicated treatments. Depression is arguably the one of the most heterogeneous conditions, with differing disease trajectories and treatment responses, and there has been some success with machine learning models defining subgroups based on large-scale population and clinical data.Reference Dwyer, Kalman, Budde, Kambeitz, Ruef and Antonucci16

Recent research has demonstrated the utility of data science in stratification of psychosis and discovery of new targets.Reference Koutsouleris, Kambeitz-Ilankovic, Ruhrmann, Rosen, Ruef and Dwyer2 It has been estimated that 60% of young people who experience a first episode of psychosis never fully recoverReference Lally and Gaughran17 and over 20% will develop severe treatment-resistant schizophrenia (TRS). In schizophrenia, unsupervised clustering has found and replicated subgroups with greater structural brain changes (cortical and subcortical volume reduction) associated with chronicity and cognitive dysfunction.Reference Chand, Dwyer, Erus, Sotiras, Varol and Srinivasan18 In early-onset disorders, supervised machine learning models aimed at interrogating diagnostic weight and boundaries suggest a transdiagnostic signature of poorer outcome across depression and psychosis.Reference Lalousis, Wood, Schmaal, Chisholm, Griffiths and Reniers19 Subgroup identification using blood-based biomarkers builds on univariate and group-level approaches that have identified that 35–50% of people with schizophrenia show some evidence of immune dysfunction, as assessed by circulating pro-inflammatory cytokines.Reference Laskaris, Mancuso, Shannon Weickert, Zalesky and Chana20 Further, Boerrigter et al used a recursive two-step cluster analysis to define subgroups of people with psychosis based on pro-inflammatory cytokine mRNA levels.Reference Boerrigter, Weickert, Lenroot, O'Donnell, Galletly and Liu21

Similar work combining multimodal data is advancing stratification of mild cognitive impairment and dementia. Young and colleagues report the working pipeline to uncover stage and subtype of dementias with fine-grained patient stratification, enabling advanced prediction of progression patterns.Reference Young, Marinescu, Oxtoby, Bocchetta, Yong and Firth22 However, in all aspects of stratification, complex interactions at individual level and acknowledgement of environmental and illness layers remain a challenge.

Prediction of treatment response, adherence and relapse

There has been a large growth in published models able to predict treatment response and treatment resistance in schizophrenia and depression in recent years. Potentially, tools developed from data science may be embedded into clinical practice, so that a clinician and patient can be guided in treatment choice, rather than trial-and-error prescribing. Prediction of treatment response can be framed either in predicting broadly determined non-response (e.g. in TRS or treatment-resistant depression (TRD)) or in predicting specific response to individual interventions (e.g. response to specific antidepressants).

Prediction of risk for TRD and TRS could potentially provide alerts for increased monitoring and timely use of existing treatments (e.g. clozapine, electroconvulsive therapy). Pigoni recently reviewed prediction of TRD with eight studies, five of which focused solely on clinical and demographic data.Reference Pigoni, Delvecchio, Madonna, Bressi, Soares and Brambilla23 Most studies reported reasonable predictive accuracy, including those with external validation samples. Leighton et al developed and externally validated a supervised model based on clinical data from the UK national EDEN study, able to predict symptom remission with an area under the curve (AUC) of 0.70.Reference Leighton, Upthegrove, Krishnadas, Benros, Broome and Gkoutos24 Legge et al used clinical and genomic data with a conditional inference forest model to predict treatment-resistant psychosis, finding a lower accuracy of prediction (AUC = 0.59): young age at onset, family history, IQ, and poor social and occupational functioning at baseline were significant features, whereas genetic liability was not associated with treatment resistance.Reference Legge, Dennison, Pardiñas, Rees, Lynham and Hopkins25

In this issue, Lee et alReference Lee, Leighton, Thomas, Gkoutos, Wood and Fenton26 review 13 studies presenting models that predict outcome after first-episode psychosis, only one including early treatment resistance,Reference Lee, akeuchi, Fervaha, Sin, Foussias and Agid27 with multimodal models to date largely built to predict broad outcomes after a first episode of psychosis (e.g. functioning, recovery) rather than TRS or transition to psychosis from the CHR state.Reference Koutsouleris, Kambeitz-Ilankovic, Ruhrmann, Rosen, Ruef and Dwyer2

In terms of individual response to a specific treatment, there is more research activity in depression than in disorders such as schizophrenia or dementia, perhaps owing to the lack of diversity in medication options for those conditions (e.g. all current antipsychotics act as dopamine antagonists). Several models have been developed for antidepressant response, including those originating from the STAR*D trial, for example Chekroud et alReference Chekroud, Zotti, Shehzad, Gueorguieva, Johnson and Trivedi28 and Laje et al,Reference Laje, Perlis, Rush and McMahon29 with an externally validated model to predict response to citalopram. Recent advances in pharmacogenetic biomarkers, including gene expression profiles and single nucleotide polymorphisms, hold promise in predicting adverse drug reactions and response to antidepressants.Reference Lin, Lin and Lane30

Recent studies have also aimed to disentangle the response to non-pharmacological treatments (e.g. ELECT-TDCS, involving transcranial direct current stimulation in major depression)Reference Kambeitz, Goerigk, Gattaz, Falkai, Benseñor and Lotufo31 and models predicting response to cognitive–behavioural psychotherapy (CBT)Reference Hilbert, Kunas, Lueken, Kathmann, Fydrich and Fehm32 and cognitive training.Reference Haas, Antonucci, Wenzel, Ruef, Biagianti and Paolini33 Moreover, digital psychotherapeutic interventions are increasingly common, and machine learning approaches have been used to predict symptom change in response to an internet intervention for depression.Reference Pearson, Pisner, Meyer, Shumake and Beevers34 Their predictions outperform linear regression models and use easily accessible clinical data, increasing the potential for clinical implementation.

Prediction of treatment response is tightly connected to the prediction of adherence, hospital readmission and side-effects. In this issue, Bannemann et alReference Bennemann, Schwartz, Giesemann and Lutz35 compare different machine learning algorithms to identify the most clinically useful model that predicts response to CBT in naturalistic settings. The authors report that tree-based and boosted algorithms that include a variable selection process are the most well-suited to predict drop out from CBT. The highest AUC (of 63. 4%) was found for lower education and younger age, as well as strongly pronounced negativistic and antisocial personality traits in contrast to less pronounced compulsiveness traits. Work has also been done on the predicting of hospital readmission within 2 years of follow-up in patients with depressionReference Cearns, Opel, Clark, Kaehler, Thalamuthu and Heindel36 that indicates that again models based on a combination of biomarkers and clinical data outperform models based on clinical variables alone.Reference Koutsouleris, Kambeitz-Ilankovic, Ruhrmann, Rosen, Ruef and Dwyer2,Reference Koutsouleris, Dwyer, Degenhardt, Maj, Urquijo-Castro and Sanfelici10 The most far-reaching in terms of multi-centre generalisability is the prediction of readmission to hospital, with up to 74% balanced accuracy (BAC), using data from the European First Episode Schizophrenia Trial (EUFEST).Reference Koutsouleris, Kahn, Chekroud, Leucht, Falkai and Wobrock37

Predictive models for the early identification of the risk for developing a disorder, relapse, therapy response or adherence may provide prompter identification of individuals requiring close clinical monitoring.Reference Torous and Keshavan38 Digital approaches such as the experience sampling method (ESM) could be used to actively monitor self-rated mental states, and passive digital phenotyping (phone messages, keyboard use, etc.) also have potential to inform machine learning models, adding real-time data.Reference Koppe, Meyer-Lindenberg and Durstewitz39 Most digital measures, either active or passive, are acquired in a longitudinal manner. As such they may be more ecologically valid then symptom rating acquired in traditional cross-sectional studies and they bypass difficulties in bringing patients to the clinical setting. However, no systematically validated prediction models using ESM and digital data are currently available, and ethical considerations, including data protection, are still to be addressed.

Finally, deep learning models may bring the most future promise by outperforming more classic machine learning algorithms. This is possibly due to the suitability of deep neural networks (DNNs) for the high-level representations with minimal domain-specific knowledge and prior feature construction.Reference Durstewitz, Koppe and Meyer-Lindenberg40 DNNs require large data-sets containing thousands of data points to provide enough material for the models to learn. As psychiatry is generally struggling with the scale of data-sets needed, the implementation of deep learning paradigms would require coordinated efforts of clinicians, researchers and healthcare providers to deliver faster progress in this field.

Conclusions

Advanced multimodal data science utilising clinical, neuroimaging, proteomic, genomic and digital biomarker data has the potential to address key challenges in psychiatry. This includes the identification of subgroups for novel targeted treatments, improved individual targeting of existing treatments and identification of those at risk of developing a disorder or relapse of existing conditions. However, the routine use of machine learning to guide clinical judgement has not yet come to fruition, and its independent use has yet to surpass the ability of clinicians’ best guesses. However, this may not be the fundamental flaw of precision psychiatry, but the present challenge of data availability for already developed, highly performing models which now need to be applied in prospective real-world conditions. The gap is in this last, but most profound step in translation. Coordinated research-ready mental health services are needed to support the scale of clinical, sociodemographic, biomarker and intervention data that would allow the advancement of precision psychiatry to equal that achieved in other areas of precision medicine.

Data availability

Data availability is not applicable to this article as no new data were created or analysed in this study.

Author contributions

L.K.-I. and R.U. jointly conceptualised this analysis piece, selected the content and drafted the manuscript. N.K. was involved in conceptualisation of the piece and commented on the manuscript draft.

Funding

R.U. is supported by the UK Research and Innovation (UKRI) Medical Research Council (MRC) Psychosis Immune Mechanism Stratified Medicine Study (PIMS) (grant: UKRI MR/S037675/1). L.K.-I. is supported by Brain and Behavior Research Foundation through Young Investigator Award No 28474. N.K. is supported by EU-FP7 project PRONIA (‘Personalised Prognostic Tools for Early Psychosis Management’) under the Grant Agreement No. 602152. None of the funding programmes were directly used to fund writing of this manuscript.

Declaration of interest

R.U. reports grants from: the Medical Research Council; the National Institute for Health Research Health Technology Assessment; the European Commission – Research: The Seventh Framework Programme; and the National Institute for Mental Health; and personal fees from Sunovion outside the submitted work. N.K. reports grants from: the European Commission; US National Institute of Mental Health; the EraPerMed programme; and the German Federal Ministry of Science and Education; and patent no. US10463313B2 ‘Adaptive pattern recognition for psychosis risk modelling’. R.U. is deputy editor of the BJPsych and L.K.-I. is guest editor for the themed issue ‘Precision medicine and personalised healthcare in psychiatry’. R.U. and L.K.-I. did not take part in the review or decision-making process of this paper.

References

Arns, M, van Dijk H, , Luykx, JJ, van Wingen, G, Olbrich, S. Stratified psychiatry: tomorrow's precision psychiatry? Eur Neuropsychopharmacol 2021; 55: 14–9.10.1016/j.euroneuro.2021.10.863CrossRefGoogle Scholar
Koutsouleris, N, Kambeitz-Ilankovic, L, Ruhrmann, S, Rosen, M, Ruef, A, Dwyer, DB, et al. Prediction models of functional outcomes for individuals in the clinical high-risk state for psychosis or with recent-onset depression: a multimodal, multisite machine learning analysis. JAMA Psychiatry 2018; 75: 1156–72.10.1001/jamapsychiatry.2018.2165CrossRefGoogle ScholarPubMed
Cannon, TD, Yu, C, Addington, J, Bearden, CE, Cadenhead, KS, Cornblatt, BA, et al. An individualized risk calculator for research in prodromal psychosis. Am J Psychiatry 2016; 173: 980–8.10.1176/appi.ajp.2016.15070890CrossRefGoogle ScholarPubMed
Koutsouleris, N, Worthington, M, Dwyer, DB, Kambeitz-Ilankovic, L, Sanfelici, R, Fusar-Poli, P. Toward generalizable and transdiagnostic tools for psychosis prediction: an independent validation and improvement of the NAPLS-2 risk calculator in the multisite PRONIA cohort. Biol Psychiatry 2021; 90: 632–42.10.1016/j.biopsych.2021.06.023CrossRefGoogle ScholarPubMed
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.e19–27.10.1016/j.neurobiolaging.2010.05.023CrossRefGoogle ScholarPubMed
Varatharajah, Y, Ramanan, VK, Iyer, R, Vemuri, P. Predicting short-term MCI-to-AD progression using imaging, CSF, genetic factors, cognitive resilience, and demographics. Sci Rep 2019; 9: 2235.10.1038/s41598-019-38793-3CrossRefGoogle ScholarPubMed
van Maurik, IS, et al. Biomarker-based prognosis for people with mild cognitive impairment (ABIDE): a modelling study. Lancet Neurol 2019; 18: 1034–44.10.1016/S1474-4422(19)30283-2CrossRefGoogle ScholarPubMed
Murray, GK, Lin, T, Austin, J, McGrath, JJ, Hickie, IB, Wray, NR. Could polygenic risk scores be useful in psychiatry? A review. JAMA Psychiatry 2021; 78: 210–19.10.1001/jamapsychiatry.2020.3042CrossRefGoogle ScholarPubMed
Fang, Y, Scott, L, Song, P, Burmeister, M, Sen, S. Genomic prediction of depression risk and resilience under stress. Nat Hum Behav 2020; 4: 111–8.10.1038/s41562-019-0759-3CrossRefGoogle ScholarPubMed
Koutsouleris, N, Dwyer, DB, Degenhardt, F, Maj, C, Urquijo-Castro, MF, Sanfelici, R, et al. Multimodal machine learning workflows for prediction of psychosis in patients with clinical high-risk syndromes and recent-onset depression. JAMA Psychiatry 2021; 78: 195209.10.1001/jamapsychiatry.2020.3604CrossRefGoogle ScholarPubMed
Toenders, YJ, Kottaram, A, Dinga, R, Davey, CG, Banaschewski, T, Bokde, ALW, et al. Predicting depression onset in young people based on clinical, cognitive, environmental, and neurobiological data. Biol Psychiatry Cogn Neurosci Neuroimaging [Epub ahead of print] 19 Mar 2021. Available from: https://doi.org/10.1016/j.bpsc.2021.03.005.Google ScholarPubMed
Satterthwaite, TD, Connolly, JJ, Ruparel, K, Calkins, ME, Jackson, C, Elliott, MA, et al. The Philadelphia neurodevelopmental cohort: a publicly available resource for the study of normal and abnormal brain development in youth. NeuroImage 2016; 124: 1115–9.10.1016/j.neuroimage.2015.03.056CrossRefGoogle Scholar
Koutsouleris, N, Meisenzahl, EM, Davatzikos, C, Bottlender, R, Frodl, T, Scheuerecker, J, 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: 700–12.10.1001/archgenpsychiatry.2009.62CrossRefGoogle ScholarPubMed
Mechelli, A, Vieira, S. From models to tools: clinical translation of machine learning studies in psychosis. NPJ Schizophr 2020; 6(1): 4.10.1038/s41537-020-0094-8CrossRefGoogle ScholarPubMed
Kambeitz, J, Kambeitz-Ilankovic, L, Leucht, S, Wood, S, Davatzikos, C, Malchow, B, et al. Detecting neuroimaging biomarkers for schizophrenia: a meta-analysis of multivariate pattern recognition studies. Neuropsychopharmacol Off Publ Am Coll Neuropsychopharmacol 2015; 40: 1742–51.10.1038/npp.2015.22CrossRefGoogle ScholarPubMed
Dwyer, D, Kalman, JL, Budde, M, Kambeitz, J, Ruef, A, Antonucci, LA, et al. An investigation of psychosis subgroups with prognostic validation and exploration of genetic underpinnings: the PsyCourse study. JAMA Psychiatry 2020; 77: 523–33.10.1001/jamapsychiatry.2019.4910CrossRefGoogle ScholarPubMed
Lally, J, Gaughran, F. Treatment resistant schizophrenia – review and a call to action. Ir J Psychol Med 2019; 36: 279–91.10.1017/ipm.2018.47CrossRefGoogle Scholar
Chand, GB, Dwyer, DB, Erus, G, Sotiras, A, Varol, E, Srinivasan, D, et al. Two distinct neuroanatomical subtypes of schizophrenia revealed using machine learning. Brain 2020; 143: 1027–38.10.1093/brain/awaa025CrossRefGoogle ScholarPubMed
Lalousis, PA, Wood, SJ, Schmaal, L, Chisholm, K, Griffiths, SL, Reniers, RLEP, et al. Heterogeneity and classification of recent onset psychosis and depression: a multimodal machine learning approach. Schizophr Bull 2021; 47: 1130–40.10.1093/schbul/sbaa185CrossRefGoogle ScholarPubMed
Laskaris, L, Mancuso, S, Shannon Weickert, C, Zalesky, A, Chana, G, et al. Brain morphology is differentially impacted by peripheral cytokines in schizophrenia-spectrum disorder. Brain Behav Immun 2021; 95: 299309.10.1016/j.bbi.2021.04.002CrossRefGoogle ScholarPubMed
Boerrigter, D, Weickert, TW, Lenroot, R, O'Donnell, M, Galletly, C, Liu, D, et al. Using blood cytokine measures to define high inflammatory biotype of schizophrenia and schizoaffective disorder. J Neuroinflammation 2017; 14: 188.10.1186/s12974-017-0962-yCrossRefGoogle ScholarPubMed
Young, AL, Marinescu, RV, Oxtoby, NP, Bocchetta, M, Yong, K, Firth, NC, et al. Uncovering the heterogeneity and temporal complexity of neurodegenerative diseases with subtype and stage inference. Nat Commun 2018; 9: 4273.10.1038/s41467-018-05892-0CrossRefGoogle ScholarPubMed
Pigoni, A, Delvecchio, G, Madonna, D, Bressi, C, Soares, J, Brambilla, P. Can Machine Learning help us in dealing with treatment resistant depression? A review. J Affect Disord 2019; 259: 21–6.10.1016/j.jad.2019.08.009CrossRefGoogle ScholarPubMed
Leighton, SP, Upthegrove, R, Krishnadas, R, Benros, ME, Broome, MR, Gkoutos, GV, et al. Development and validation of multivariable prediction models of remission, recovery, and quality of life outcomes in people with first episode psychosis: a machine learning approach. Lancet Digit Health 2019; 1: e261–e70.10.1016/S2589-7500(19)30121-9CrossRefGoogle ScholarPubMed
Legge, SE, Dennison, CA, Pardiñas, AF, Rees, E, Lynham, AJ, Hopkins, L, et al. Clinical indicators of treatment-resistant psychosis. Br J Psychiatry 2020; 216: 259–66.10.1192/bjp.2019.120CrossRefGoogle ScholarPubMed
Lee, R, Leighton, SP, Thomas, L, Gkoutos, GV, Wood, SJ, Fenton, S-JH, et al. Prediction models in first-episode psychosis: systematic review and critical appraisal. Br J Psychiatry 2022; 220.4: XXX.Google Scholar
Lee, J, akeuchi, H, Fervaha, G, Sin, GL, Foussias, G, Agid, O, et al. Subtyping schizophrenia by treatment response: antipsychotic development and the central role of positive symptoms. Can J Psychiatry 2015; 60: 515–22.10.1177/070674371506001107CrossRefGoogle ScholarPubMed
Chekroud, AM, Zotti, RJ, Shehzad, Z, Gueorguieva, R, Johnson, MK, Trivedi, MH, et al. Cross-trial prediction of treatment outcome in depression: a machine learning approach. Lancet Psychiatry 2016; 3: 243–50.10.1016/S2215-0366(15)00471-XCrossRefGoogle ScholarPubMed
Laje, G, Perlis, RH, Rush, AJ, McMahon, FJ. Pharmacogenetics studies in STAR*D: strengths, limitations, and results. Psychiatr Serv 2009; 60: 1446–57.10.1176/ps.2009.60.11.1446CrossRefGoogle ScholarPubMed
Lin, E, Lin, C-H, Lane, H-Y. Precision psychiatry applications with pharmacogenomics: artificial intelligence and machine learning approaches. Int J Mol Sci 2020; 21: 969.10.3390/ijms21030969CrossRefGoogle ScholarPubMed
Kambeitz, J, Goerigk, S, Gattaz, W, Falkai, P, Benseñor, IM, Lotufo, PA, et al. Clinical patterns differentially predict response to transcranial direct current stimulation (tDCS) and escitalopram in major depression: a machine learning analysis of the ELECT-TDCS study. J Affect Disord 2020; 265: 460–7.10.1016/j.jad.2020.01.118CrossRefGoogle ScholarPubMed
Hilbert, K, Kunas, SL, Lueken, U, Kathmann, N, Fydrich, T, Fehm, L. Predicting cognitive behavioral therapy outcome in the outpatient sector based on clinical routine data: a machine learning approach. Behav Res Ther 2020; 124: 103530.10.1016/j.brat.2019.103530CrossRefGoogle ScholarPubMed
Haas, SS, Antonucci, LA, Wenzel, J, Ruef, A, Biagianti, B, Paolini, M, et al. A multivariate neuromonitoring approach to neuroplasticity-based computerized cognitive training in recent onset psychosis. Neuropsychopharmacology 2021: 46: 828–35.10.1038/s41386-020-00877-4CrossRefGoogle ScholarPubMed
Pearson, R, Pisner, D, Meyer, B, Shumake, J, Beevers, CG. A machine learning ensemble to predict treatment outcomes following an Internet intervention for depression. Psychol Med 2019; 49: 2330–41.10.1017/S003329171800315XCrossRefGoogle ScholarPubMed
Bennemann, B, Schwartz, B, Giesemann, J, Lutz, W. Predicting patients who will drop out of out-patient psychotherapy using machine learning algorithms. Br J Psychiatry 2022; 220.4: XXX.Google Scholar
Cearns, M, Opel, N, Clark, S, Kaehler, C, Thalamuthu, A, Heindel, W, et al. Predicting rehospitalization within 2 years of initial patient admission for a major depressive episode: a multimodal machine learning approach. Transl. Psychiatry 2019; 9: 285.10.1038/s41398-019-0615-2CrossRefGoogle Scholar
Koutsouleris, N, Kahn, RS, Chekroud, AM, Leucht, S, Falkai, P, Wobrock, T, et al. Multisite prediction of 4-week and 52-week treatment outcomes in patients with first-episode psychosis: a machine learning approach. Lancet Psychiatry 2016; 3: 935–46.10.1016/S2215-0366(16)30171-7CrossRefGoogle ScholarPubMed
Torous, J, Keshavan, M. Towards precision clinical trials and personalized prevention in CHR with smartphone digital phenotyping and personal sensing tools. Schizophr Res 2021; 227: 61–2.10.1016/j.schres.2020.04.007CrossRefGoogle Scholar
Koppe, G, Meyer-Lindenberg, A, Durstewitz, D. Deep learning for small and big data in psychiatry. Neuropsychopharmacol 2021; 46: 176–90.10.1038/s41386-020-0767-zCrossRefGoogle Scholar
Durstewitz, D, Koppe, G, Meyer-Lindenberg, A. Deep neural networks in psychiatry. Mol Psychiatry 2019; 24: 1583–98.10.1038/s41380-019-0365-9CrossRefGoogle Scholar
Figure 0

Fig. 1 Individual prognosis along the disease trajectory.Black lines indicate fields with stronger translational potential due to a larger number of validation studies; grey lines indicate fields of research with currently less translational perspective, owing to a sparse number of studies and validation attempts. Diff. diagnostics, differential diagnostics.

Submit a response

eLetters

No eLetters have been published for this article.