Machine learning identifies candidates for drug repurposing in Alzheimer's disease

Steve Rodriguez, Clemens Hug, Petar Todorov, Nienke Moret, Sarah A Boswell, Kyle Evans, George Zhou, Nathan T Johnson, Bradley T Hyman, Peter K Sorger, Mark W Albers, Artem Sokolov, Steve Rodriguez, Clemens Hug, Petar Todorov, Nienke Moret, Sarah A Boswell, Kyle Evans, George Zhou, Nathan T Johnson, Bradley T Hyman, Peter K Sorger, Mark W Albers, Artem Sokolov

Abstract

Clinical trials of novel therapeutics for Alzheimer's Disease (AD) have consumed a large amount of time and resources with largely negative results. Repurposing drugs already approved by the Food and Drug Administration (FDA) for another indication is a more rapid and less expensive option. We present DRIAD (Drug Repurposing In AD), a machine learning framework that quantifies potential associations between the pathology of AD severity (the Braak stage) and molecular mechanisms as encoded in lists of gene names. DRIAD is applied to lists of genes arising from perturbations in differentiated human neural cell cultures by 80 FDA-approved and clinically tested drugs, producing a ranked list of possible repurposing candidates. Top-scoring drugs are inspected for common trends among their targets. We propose that the DRIAD method can be used to nominate drugs that, after additional validation and identification of relevant pharmacodynamic biomarker(s), could be readily evaluated in a clinical trial.

Conflict of interest statement

The authors declare the following competing interests. P.K.S. is a member of the SAB or Board of Directors of Applied Biomath, RareCyte, NanoString and Glencoe Software and has equity in some of these companies. In the last 5 years, the Sorger lab has received research funding from Novartis and Merck. P.K.S. declares that none of these relationships are directly or indirectly related to the content of this manuscript. B.T.H. has stock in Novartis and Dewpoint. N.T.J. is an employee of H3 Biomedicine, a subsidiary of Eisai Inc. that develops therapies for Alzheimer’s. S.R., P.K.S., M.W.A., and A.S. are inventors on a patent application (WO/2017/173451) for novel targets in neurodegenerative diseases. All other authors (C.H., P.T., N.M., S.B., K.E., G.Z.) declare no competing interests.

Figures

Fig. 1. The definition and validation of…
Fig. 1. The definition and validation of the DRIAD framework.
a Overview of the machine learning framework used to establish potential associations between gene lists and Alzheimer’s disease. (i) The framework accepts as input gene lists derived from experimental data or extracted from database resources or literature. (ii) Given a gene expression matrix, the framework subsamples it to a particular gene list of interest, and (iii) subsequently trains and evaluates through cross-validation a predictor of Braak stage of disease. (iv) The process is repeated for randomly selected gene lists of equal lengths to determine whether predictor performance associated with the gene list of interest is significantly higher than what is expected by chance. b AMP-AD datasets used by the machine learning framework. The three datasets used to evaluate the predictive power of gene lists are provided by The Religious Orders Study and Memory and Aging Project (ROSMAP), The Mayo Clinic Brain Bank (MAYO), and The Mount Sinai/JJ Peters VA Medical Center Brain Bank (MSBB). The schematic highlights regions of the brain that are represented in each dataset. The MSBB dataset spans four distinct regions, which are designated using Brodmann (BM) area codes. c Performance of predictors trained on gene lists reported in previous studies of AMP-AD datasets. The predictors are evaluated for their ability to distinguish early-vs-late disease stages with performance reported as area under the ROC curve (AUC). The vertical line on each row denotes predictor performance associated with a gene list reported in the literature, while the background distribution is constructed over randomly selected lists of matching lengths. Each row is annotated with the pubmed ID of the study, the supplemental resource that contained the gene list, and a short keyphrase providing functional context. Shown unadjusted p-values were computed with a one-sided empirical test, by counting the fraction of randomly selected lists in the background distribution that outperformed the corresponding literature lists.
Fig. 2. Collection and evaluation of drug-associated…
Fig. 2. Collection and evaluation of drug-associated gene lists.
a Overview of the 3′ DGE experimental protocol used to derive drug-associated gene expression signatures. ReNcell VM human neural progenitor cells were plated and differentiated for 10 days, resulting in a mixed cell population of neurons, glia, and oligodendrocytes. The mixed culture was subsequently treated with a panel of drugs (Supplementary Data 3) at 10 µM for 24 h and frozen in a lysis buffer until library preparation. RNA was extracted and reverse transcribed into cDNA in each well of the plate, followed by pooling and preparation of mRNA libraries. After sequencing, mRNA reads were demultiplexed according to well barcodes, and the resulting gene expression profiles were processed by a standard differential expression method to derive drug-associated gene lists. b A highlight of two compounds whose gene lists consistently yield improved performance over the randomly selected lists of equal length. Shown is performance associated with predicting early-vs-late disease stages in several AMP-AD datasets. Each row corresponds to an evaluation of gene lists in a single dataset; MSBB evaluation is subdivided into four brain regions, specified as Brodmann Area. The vertical line denotes performance of the drug-associated list, while the background distribution shows performance of gene lists randomly selected from the same dataset. The drugs are annotated with their nominal targets. The unadjusted p-values were computed with a one-sided empirical test, by counting the fraction of randomly selected lists that outperformed the corresponding drug-associated lists.
Fig. 3. Top 15 FDA-approved (left) and…
Fig. 3. Top 15 FDA-approved (left) and experimental/investigational (right) drugs, sorted by harmonic mean p-value.
Each heatmap shows unadjusted empirical p-values associated with a drug’s predictive performance across two AMP-AD datasets, ROSMAP and MSBB. The MSBB analysis is further subdivided by the brain region, specified as Brodmann Area. The empirical p-values were computed by counting the proportion of randomly selected lists that outperformed the gene lists of interest (i.e., a one-sided test). The p-values were then aggregated across the datasets by computing the harmonic mean p-value (HMP), which is shown in the last column of each heatmap. The rows are annotated with the name of the drug/compound, its nominal target, and the index of the corresponding DGE experiment. Additional annotations include information about each compound’s approval status (approved/investigational/experimental) and whether compounds were found to be toxic in neuronal cell cultures.
Fig. 4. Analysis of target affinity among…
Fig. 4. Analysis of target affinity among the top-scoring drugs.
a Overview of target affinity spectrum (TAS) score computation from raw drug binding data. Three types of drug binding data were sourced from ChEMBL and from the internal Laboratory Systems of Pharmacology dataset that have not yet been incorporated into ChEMBL. Empirically derived thresholds for the different data types were used to assign TAS scores to each drug–target pair. Multiple measurements for the same drug–target combination were aggregated along the first quartile to define the final TAS value. b Binding affinity of compounds in the ranked list to every member of the Janus Kinase family. The compounds are sorted in increasing order by the harmonic mean p-value (as defined in Fig. 3) along the x-axis. The top heatmap shows the binding affinity of each compound to the selected targets, explicitly naming the FDA-approved drugs. Colored and gray tiles denote confirmed binders and non-binders, respectively; missing entries correspond to unknown affinity values. The combined affinity is defined as the strongest binding (lowest TAS score) among all four JAK targets. The bottom plot shows the breakdown of the combined affinity values by TAS-specific empirical cumulative distribution functions (ECDFs). Each line shows ECDFs for all drugs that bind the corresponding target with a TAS score of 1 (dark orange), 2 (orange), or 3 (light orange). c Top targets whose binding affinity correlates most strongly with the compound ranking. The ECDFs of confirmed non-binders (TAS = 10) are shown as gray dashed lines for reference. Area under ECDF can be interpreted as a summary statistic that captures the position of drugs binding to that target with the corresponding affinity in the ranked list. Correlation between the drug ranking and TAS values was computed using the one-sided Kendall’s Tau test, with the associated unadjusted p-value displayed in the bottom right corner of each plot.
Fig. 5. Analysis of polypharmacology effects among…
Fig. 5. Analysis of polypharmacology effects among the top-scoring drugs.
a An example polypharmacology test with a focus on RPS6KA1 and TYK2. The drugs are ranked by the harmonic mean p-value (as in Figs. 3 and 4), and the distributions of drugs bindings to both RPS6KA1 and TYK2 (left), those binding to RPS6KA1 but not TYK2 (middle) and, conversely, TYK2 but not RPS6KA1 (right) are shown along this ranking. Individual drugs that bind those targets are annotated by vertical tick marks directly below the corresponding distribution. b Top ten positive and top ten negative interactions between pairs of targets. The distributions in each plot are compared using Wilcoxon Rank Sum test, with the resulting p-value presented in the bottom right corner. If compounds that bind both targets appear significantly closer to the top of the ranked list (left side of the x axis), we define the target pair to be a positive interaction. Conversely, a pair of targets with an explicit non-binding interaction observed among the top-ranking compounds is defined to be antagonistic. A set of five neutral target pairs (i.e., no significant positive or negative effect) is included for reference.

References

    1. Hebert LE, Weuve J, Scherr PA, Evans DA. Alzheimer disease in the United States (2010-2050) estimated using the 2010 census. Neurology. 2013;80:1778–1783. doi: 10.1212/WNL.0b013e31828726f5.
    1. Alzheimer’s Association. 2019 Alzheimer’s disease facts and figures. Alzheimers Dement. 15, 321–387 (2019).
    1. Mehta D, Jackson R, Paul G, Shi J, Sabbagh M. Why do trials for Alzheimer’s disease drugs keep failing? A discontinued drug perspective for 2010–2015. Expert Opin. Investig. Drugs. 2017;26:735–739. doi: 10.1080/13543784.2017.1323868.
    1. Pushpakom S, et al. Drug repurposing: progress, challenges and recommendations. Nat. Rev. Drug Discov. 2019;18:41–58. doi: 10.1038/nrd.2018.168.
    1. Hernandez, J. J. et al. Giving drugs a second chance: overcoming regulatory and financial hurdles in repurposing approved drugs as cancer therapeutics. Front. Oncol. 7, 273 (2017).
    1. Mallikarjun V, Swift J. Therapeutic manipulation of ageing: repurposing old dogs and discovering new tricks. EBioMedicine. 2016;14:24–31. doi: 10.1016/j.ebiom.2016.11.020.
    1. Hara Y, McKeehan N, Fillit HM. Translating the biology of aging into novel therapeutics for Alzheimer disease. Neurology. 2019;92:84–93. doi: 10.1212/WNL.0000000000006745.
    1. Nevado-Holgado AJ, Lovestone S. Determining the molecular pathways underlying the protective effect of non-steroidal anti-inflammatory drugs for Alzheimer’s disease: a bioinformatics approach. Comput. Struct. Biotechnol. J. 2017;15:1–7. doi: 10.1016/j.csbj.2016.10.003.
    1. Corbett A, et al. Drug repositioning for Alzheimer’s disease. Nat. Rev. Drug Discov. 2012;11:833–846. doi: 10.1038/nrd3869.
    1. Shoaib M, Kamal MA, Rizvi SMD. Repurposed drugs as potential therapeutic candidates for the management of Alzheimer’s disease. Curr. Drug Metab. 2017;18:842–852. doi: 10.2174/1389200218666170607101622.
    1. Brown AS, Patel CJ. MeSHDD: literature-based drug-drug similarity for drug repositioning. J. Am. Med. Inform. Assoc. 2017;24:614–618. doi: 10.1093/jamia/ocw142.
    1. Gottlieb A, Stein GY, Ruppin E, Sharan R. PREDICT: a method for inferring novel drug indications with application to personalized medicine. Mol. Syst. Biol. 2011;7:496. doi: 10.1038/msb.2011.26.
    1. Himmelstein, D. S. et al. Systematic integration of biomedical knowledge prioritizes drugs for repurposing. eLife6, e26726 (2017).
    1. Lamb J, et al. The Connectivity Map: using gene-expression signatures to connect small molecules, genes, and disease. Science. 2006;313:1929–1935. doi: 10.1126/science.1132939.
    1. Cheng F, et al. Network-based approach to prediction and population-based validation of in silico drug repurposing. Nat. Commun. 2018;9:2691. doi: 10.1038/s41467-018-05116-5.
    1. Zhang, M. et al. Drug repositioning for Alzheimer’s disease based on systematic ‘omics’ data mining. PLoS ONE11, e0168812 (2016).
    1. Regan, K., Moosavinasab, S., Payne, P. & Lin, S. Drug repurposing hypothesis generation using the ‘RE:fine Drugs’ System. J. Vis. Exp.10.3791/54948 (2016).
    1. Ryan J, Fransquet P, Wrigglesworth J, Lacaze P. Phenotypic heterogeneity in dementia: a challenge for epidemiology and biomarker studies. Front. Public Health. 2018;6:181. doi: 10.3389/fpubh.2018.00181.
    1. Boyle PA, et al. Attributable risk of Alzheimer’s dementia attributed to age-related neuropathologies. Ann. Neurol. 2019;85:114–124. doi: 10.1002/ana.25380.
    1. Hodes RJ, Buckholtz N. Accelerating medicines partnership: Alzheimer’s Disease (AMP-AD) Knowledge Portal aids Alzheimer’s drug discovery through open data sharing. Expert Opin. Ther. Targets. 2016;20:389–391. doi: 10.1517/14728222.2016.1135132.
    1. Dönertaş HM, Fuentealba Valenzuela M, Partridge L, Thornton JM. Gene expression-based drug repurposing to target aging. Aging Cell. 2018;17:e12819. doi: 10.1111/acel.12819.
    1. Williams G, et al. Drug repurposing for Alzheimer’s disease based on transcriptional profiling of human iPSC-derived cortical neurons. Transl. Psychiatry. 2019;9:220. doi: 10.1038/s41398-019-0555-x.
    1. Duan Q, et al. L1000CDS 2: LINCS L1000 characteristic direction signatures search engine. NPJ Syst. Biol. Appl. 2016;2:1–12. doi: 10.1038/npjsba.2016.15.
    1. Wang Z, Lachmann A, Keenan AB, Ma’ayan A. L1000FWD: fireworks visualization of drug-induced transcriptomic signatures. Bioinformatics. 2018;34:2150–2152. doi: 10.1093/bioinformatics/bty060.
    1. Keenan AB, et al. Connectivity Mapping: methods and applications. Annu. Rev. Biomed. Data Sci. 2019;2:69–92. doi: 10.1146/annurev-biodatasci-072018-021211.
    1. Braak H, Braak E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 1991;82:239–259. doi: 10.1007/BF00308809.
    1. Moret N, et al. Cheminformatics tools for analyzing and designing optimized small-molecule collections and libraries. Cell Chem. Biol. 2019;26:765–777.e3. doi: 10.1016/j.chembiol.2019.02.018.
    1. Long JM, Holtzman DM. Alzheimer disease: an update on pathobiology and treatment strategies. Cell. 2019;179:312–339. doi: 10.1016/j.cell.2019.09.001.
    1. Jouanne M, Rault S, Voisin-Chiret A-S. Tau protein aggregation in Alzheimer’s disease: an attractive target for the development of novel therapeutic agents. Eur. J. Med. Chem. 2017;139:153–167. doi: 10.1016/j.ejmech.2017.07.070.
    1. Li, X., Long, J., He, T., Belshaw, R. & Scott, J. Integrated genomic approaches identify major pathways and upstream regulators in late onset Alzheimer’s disease. Sci. Rep. 5, 12393 (2015).
    1. Miller JA, Oldham MC, Geschwind DH. A systems level analysis of transcriptional changes in Alzheimer’s disease and normal aging. J. Neurosci. 2008;28:1410–1420. doi: 10.1523/JNEUROSCI.4098-07.2008.
    1. Seyfried NT, et al. A multi-network approach identifies protein-specific co-expression in asymptomatic and symptomatic Alzheimer’s disease. Cell Syst. 2017;4:60–72.e4. doi: 10.1016/j.cels.2016.11.006.
    1. Mizuno S, et al. AlzPathway: a comprehensive map of signaling pathways of Alzheimer’s disease. BMC Syst. Biol. 2012;6:52. doi: 10.1186/1752-0509-6-52.
    1. Mostafavi S, et al. A molecular network of the aging human brain provides insights into the pathology and cognitive decline of Alzheimer’s disease. Nat. Neurosci. 2018;21:811–819. doi: 10.1038/s41593-018-0154-9.
    1. Allen M, et al. Conserved brain myelination networks are altered in Alzheimer’s and other neurodegenerative diseases. Alzheimers Dement. 2018;14:352–366. doi: 10.1016/j.jalz.2017.09.012.
    1. Allen M, et al. Divergent brain gene expression patterns associate with distinct cell-specific tau neuropathology traits in progressive supranuclear palsy. Acta Neuropathol. 2018;136:709–727. doi: 10.1007/s00401-018-1900-5.
    1. Sekiya M, et al. Integrated biology approach reveals molecular and pathological interactions among Alzheimer’s Aβ42, Tau, TREM2, and TYROBP in Drosophila models. Genome Med. 2018;10:26. doi: 10.1186/s13073-018-0530-9.
    1. Wang M, et al. Integrative network analysis of nineteen brain regions identifies molecular signatures and networks underlying selective regional vulnerability to Alzheimer’s disease. Genome Med. 2016;8:104. doi: 10.1186/s13073-016-0355-3.
    1. Raj T, et al. Integrative transcriptome analyses of the aging brain implicate altered splicing in Alzheimer’s disease susceptibility. Nat. Genet. 2018;50:1584–1592. doi: 10.1038/s41588-018-0238-1.
    1. McKenzie AT, et al. Multiscale network modeling of oligodendrocytes reveals molecular components of myelin dysregulation in Alzheimer’s disease. Mol. Neurodegener. 2017;12:82. doi: 10.1186/s13024-017-0219-3.
    1. Bennett RE, et al. Tau induces blood vessel abnormalities and angiogenesis-related gene expression in P301L transgenic mice and human Alzheimer’s disease. Proc. Natl Acad. Sci. USA. 2018;115:E1289–E1298. doi: 10.1073/pnas.1710329115.
    1. Johnson WE, Li C, Rabinovic A. Adjusting batch effects in microarray expression data using empirical Bayes methods. Biostatistics. 2007;8:118–127. doi: 10.1093/biostatistics/kxj037.
    1. Leek JT, Storey JD. Capturing heterogeneity in gene expression studies by surrogate variable analysis. PLoS Genet. 2007;3:e161. doi: 10.1371/journal.pgen.0030161.
    1. Gagnon-Bartsch, J. A., Jacob, L. & Speed, T. P. Removing Unwanted Variation From High Dimensional Data With Negative Controls. Berkeley Technical Report 1–112 (Department of Statistics, University of California, 2013).
    1. Song Y, et al. A dynamic view of the proteomic landscape during differentiation of ReNcell VM cells, an immortalized human neural progenitor line. Sci. Data. 2019;6:190016. doi: 10.1038/sdata.2019.16.
    1. Ye C, et al. DRUG-seq for miniaturized high-throughput transcriptome profiling in drug discovery. Nat. Commun. 2018;9:1–9. doi: 10.1038/s41467-017-02088-w.
    1. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26:139–140. doi: 10.1093/bioinformatics/btp616.
    1. Wilson DJ. The harmonic mean p-value for combining dependent tests. Proc. Natl Acad. Sci. USA. 2019;116:1195–1200. doi: 10.1073/pnas.1814092116.
    1. Marsilje TH, et al. Synthesis, structure–activity relationships, and in vivo efficacy of the novel potent and selective anaplastic lymphoma kinase (ALK) inhibitor 5-chloro-N2-(2-isopropoxy-5-methyl-4-(piperidin-4-yl)phenyl)-N4-(2-(isopropylsulfonyl)phenyl)pyrimidine-2,4-diamine (LDK378) currently in phase 1 and phase 2 clinical trials. J. Med. Chem. 2013;56:5675–5690. doi: 10.1021/jm400402q.
    1. Hafner M, et al. Multiomics profiling establishes the polypharmacology of FDA-approved CDK4/6 inhibitors and the potential for differential clinical activity. Cell Chem. Biol. 2019;26:1067–1080.e8. doi: 10.1016/j.chembiol.2019.05.005.
    1. Taylor JM, Moore Z, Minter MR, Crack PJ. Type-I interferon pathway in neuroinflammation and neurodegeneration: focus on Alzheimer’s disease. J. Neural Transm. (Vienna) 2018;125:797–807. doi: 10.1007/s00702-017-1745-4.
    1. Tang H-W, et al. Atg1-mediated myosin II activation regulates autophagosome formation during starvation-induced autophagy. EMBO J. 2011;30:636–651. doi: 10.1038/emboj.2010.338.
    1. Zachari M, Ganley IG. The mammalian ULK1 complex and autophagy initiation. Essays Biochem. 2017;61:585–596. doi: 10.1042/EBC20170021.
    1. Lee E-J, Tournier C. The requirement of uncoordinated 51-like kinase 1 (ULK1) and ULK2 in the regulation of autophagy. Autophagy. 2011;7:689–695. doi: 10.4161/auto.7.7.15450.
    1. Zare-Shahabadi A, Masliah E, Johnson GVW, Rezaei N. Autophagy in Alzheimer’s disease. Rev. Neurosci. 2015;26:385–395. doi: 10.1515/revneuro-2014-0076.
    1. Funderburk SF, Marcellino BK, Yue Z. Cell ‘self-eating’ (autophagy) mechanism in Alzheimer’s disease. Mt. Sinai J. Med. 2010;77:59–68. doi: 10.1002/msj.20161.
    1. Abd-Elrahman KS, Hamilton A, Vasefi M, Ferguson SSG. Autophagy is increased following either pharmacological or genetic silencing of mGluR5 signaling in Alzheimer’s disease mouse models. Mol. Brain. 2018;11:19. doi: 10.1186/s13041-018-0364-9.
    1. Wang L, Jin G, Yu H, Li Q, Yang H. Protective effect of Tenuifolin against Alzheimer’s disease. Neurosci. Lett. 2019;705:195–201. doi: 10.1016/j.neulet.2019.04.045.
    1. Patir A, Shih B, McColl BW, Freeman TC. A core transcriptional signature of human microglia: Derivation and utility in describing region-dependent alterations associated with Alzheimer’s disease. Glia. 2019;67:1240–1253. doi: 10.1002/glia.23572.
    1. Ellison EM, Bradley-Whitman MA, Lovell MA. Single-base resolution mapping of 5-hydroxymethylcytosine modifications in hippocampus of Alzheimer’s disease subjects. J. Mol. Neurosci. 2017;63:185–197. doi: 10.1007/s12031-017-0969-y.
    1. Chao MV, Rajagopal R, Lee FS. Neurotrophin signalling in health and disease. Clin. Sci. Lond. 2006;110:167–173. doi: 10.1042/CS20050163.
    1. Mortezaei Z, Lanjanian H, Masoudi-Nejad A. Candidate novel long noncoding RNAs, MicroRNAs and putative drugs for Parkinson’s disease using a robust and efficient genome-wide association study. Genomics. 2017;109:158–164. doi: 10.1016/j.ygeno.2017.02.004.
    1. Chang J, Baloh RH, Milbrandt J. The NIMA-family kinase Nek3 regulates microtubule acetylation in neurons. J. Cell Sci. 2009;122:2274–2282. doi: 10.1242/jcs.048975.
    1. Pei J-J, et al. P70 S6 kinase mediates tau phosphorylation and synthesis. FEBS Lett. 2006;580:107–114. doi: 10.1016/j.febslet.2005.11.059.
    1. Pei J-J, Björkdahl C, Zhang H, Zhou X, Winblad B. p70 S6 kinase and tau in Alzheimer’s disease. J. Alzheimers Dis. 2008;14:385–392. doi: 10.3233/JAD-2008-14405.
    1. Zhou X-W, Tanila H, Pei J-J. Parallel increase in p70 kinase activation and tau phosphorylation (S262) with Abeta overproduction. FEBS Lett. 2008;582:159–164. doi: 10.1016/j.febslet.2007.11.078.
    1. Wu M, et al. Identification of key genes and pathways for Alzheimer’s disease via combined analysis of genome-wide expression profiling in the hippocampus. Biophys. Rep. 2019;5:98–109. doi: 10.1007/s41048-019-0086-2.
    1. Caccamo A, et al. Reducing ribosomal protein S6 kinase 1 expression improves spatial memory and synaptic plasticity in a mouse model of Alzheimer’s disease. J. Neurosci. 2015;35:14042–14056. doi: 10.1523/JNEUROSCI.2781-15.2015.
    1. Hernandez, I. et al. A farnesyltransferase inhibitor activates lysosomes and reduces tau pathology in mice with tauopathy. Sci. Transl. Med. 11, eaat3005 (2019).
    1. Subramanian A, et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA. 2005;102:15545–15550. doi: 10.1073/pnas.0506580102.
    1. Seddighi S, Houck AL, Rowe JB, Pharoah PDP. Evidence of a causal association between cancer and alzheimer’s disease: a mendelian randomization analysis. Sci Rep. 2019;18:13548. doi: 10.1038/s41598-019-49795-6.
    1. Ospina-Romero M, et al. Rate of memory change before and after cancer diagnosis. JAMA Netw. Open. 2019;2:e196160. doi: 10.1001/jamanetworkopen.2019.6160.
    1. Zhou Y, et al. Human and mouse single-nucleus transcriptomics reveal TREM2-dependent and TREM2-independent cellular responses in Alzheimer’s disease. Nat. Med. 2020;26:131–142. doi: 10.1038/s41591-019-0695-9.
    1. Tam OH, et al. Postmortem cortex samples identify distinct molecular subtypes of ALS: retrotransposon activation, oxidative stress, and activated glia. Cell Rep. 2019;29:1164–1177.e5. doi: 10.1016/j.celrep.2019.09.066.
    1. Saldi TK, Gonzales PK, LaRocca TJ, Link CD. Neurodegeneration, heterochromatin, and double-stranded RNA. J. Exp. Neurosci. 2019;13:1179069519830697. doi: 10.1177/1179069519830697.
    1. Bai B, et al. U1 small nuclear ribonucleoprotein complex and RNA splicing alterations in Alzheimer’s disease. Proc. Natl Acad. Sci. USA. 2013;110:16562–16567. doi: 10.1073/pnas.1310249110.
    1. Gjoneska E, et al. Conserved epigenomic signals in mice and humans reveal immune basis of Alzheimer’s disease. Nature. 2015;518:365–369. doi: 10.1038/nature14252.
    1. Liddelow SA, et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature. 2017;541:481–487. doi: 10.1038/nature21029.
    1. Boyle PA, et al. Person-specific contribution of neuropathologies to cognitive loss in old age. Ann. Neurol. 2018;83:74–83. doi: 10.1002/ana.25123.
    1. Park J, et al. A 3D human triculture system modeling neurodegeneration and neuroinflammation in Alzheimer’s disease. Nat. Neurosci. 2018;21:941–951. doi: 10.1038/s41593-018-0175-4.
    1. Dickerman BA, García-Albéniz X, Logan RW, Denaxas S, Hernán MA. Avoidable flaws in observational analyses: an application to statins and cancer. Nat. Med. 2019;25:1601–1606. doi: 10.1038/s41591-019-0597-x.
    1. Dong X. Current strategies for brain drug delivery. Theranostics. 2018;8:1481–1493. doi: 10.7150/thno.21254.
    1. Srivastava A, Sarkar H, Gupta N, Patro R. RapMap: a rapid, sensitive and accurate tool for mapping RNA-seq reads to transcriptomes. Bioinformatics. 2016;32:i192–i200. doi: 10.1093/bioinformatics/btw277.
    1. Svensson V, et al. Power analysis of single cell RNA-sequencing experiments. Nat. Methods. 2017;14:381–387. doi: 10.1038/nmeth.4220.
    1. Airola A, Pahikkala T, Waegeman W, De Baets B, Salakoski T. An experimental comparison of cross-validation techniques for estimating the area under the ROC curve. Comput Stat. Data Anal. 2011;55:1828–1844. doi: 10.1016/j.csda.2010.11.018.
    1. Gaulton A, et al. The ChEMBL database in 2017. Nucleic Acids Res. 2017;45:D945–D954. doi: 10.1093/nar/gkw1074.
    1. Rodriguez, S. et al. Data: machine learning identifies candidates for drug repurposing in Alzheimer’s disease. 10.7303/syn18488020 (2021).
    1. Rodriguez, S. et al. Machine learning identifies candidates for drug repurposing in Alzheimer’s disease. 10.5281/zenodo.4432966 (2021).
    1. Potter, S. & Murrell, P. grImport2: importing ‘SVG’ graphics. (2019).
    1. Potter, S. & Murrell, P. gridSVG: export ‘grid’ graphics as SVG. (2019).

Source: PubMed

3
Subskrybuj