ArticlePDF Available

Abstract and Figures

Previous genome-wide association studies (GWAS), conducted by our group and others, have identified loci that harbor risk variants for neurodegenerative diseases, including Alzheimer's disease (AD). Human disease variants are enriched for polymorphisms that affect gene expression, including some that are known to associate with expression changes in the brain. Postulating that many variants confer risk to neurodegenerative disease via transcriptional regulatory mechanisms, we have analyzed gene expression levels in the brain tissue of subjects with AD and related diseases. Herein, we describe our collective datasets comprised of GWAS data from 2,099 subjects; microarray gene expression data from 773 brain samples, 186 of which also have RNAseq; and an independent cohort of 556 brain samples with RNAseq. We expect that these datasets, which are available to all qualified researchers, will enable investigators to explore and identify transcriptional mechanisms contributing to neurodegenerative diseases.
Content may be subject to copyright.
Data Descriptor: Human whole
genome genotype and transcriptome
data for Alzheimers and other
neurodegenerative diseases
Mariet Allen
1,*
, Minerva M. Carrasquillo
1,*
, Cory Funk
2
, Benjamin D. Heavner
2
,
Fanggeng Zou
1
, Curtis S. Younkin
3
, Jeremy D. Burgess
1
, High-Seng Chai
4
, Julia Crook
2
,
James A. Eddy
2
, Hongdong Li
2
, Ben Logsdon
5
, Mette A. Peters
5
, Kristen K. Dang
5
,
Xue Wang
3
, Daniel Serie
3
, Chen Wang
4
, Thuy Nguyen
1
, Sarah Lincoln
1
, Kimberly Malphrus
1
,
Gina Bisceglio
1
,MaLi
1
, Todd E. Golde
6
, Lara M. Mangravite
5
, Yan Asmann
2
,
Nathan D. Price
2
, Ronald C. Petersen
7
, Neill R. Graff-Radford
8
, Dennis W. Dickson
1
,
Steven G. Younkin
1
& Nilüfer Ertekin-Taner
1,8
Previous genome-wide association studies (GWAS), conducted by our group and others, have identied loci
that harbor risk variants for neurodegenerative diseases, including Alzheimer's disease (AD). Human disease
variants are enriched for polymorphisms that affect gene expression, including some that are known to
associate with expression changes in the brain. Postulating that many variants confer risk to
neurodegenerative disease via transcriptional regulatory mechanisms, we have analyzed gene expression
levels in the brain tissue of subjects with AD and related diseases. Herein, we describe our collective
datasets comprised of GWAS data from 2,099 subjects; microarray gene expression data from 773 brain
samples, 186 of which also have RNAseq; and an independent cohort of 556 brain samples with RNAseq.
We expect that these datasets, which are available to all qualied researchers, will enable investigators to
explore and identify transcriptional mechanisms contributing to neurodegenerative diseases.
Design Type disease state design individual genetic characteristics comparison design
Measurement Type(s) genetic sequence variation analysis transcription proling by array assay
Technology Type(s) Whole Genome Association Study RNA-seq assay
Factor Type(s) regional part of brain diagnosis
Sample Characteristic(s) Homo sapiens cerebellum temporal cortex
1
Mayo Clinic, Department of Neuroscience, 4500 San Pablo Road, Jacksonville, Florida 32224, USA.
2
Institute for
Systems Biology, 401 Terry Ave N., Seattle, Washington 98109, USA.
3
Mayo Clinic, Department of Health
Sciences Research, 4500 San Pablo Road, Jacksonville, Florida 32224, USA.
4
Mayo Clinic, Department of Health
Sciences Research, 200 First Street, Rochester, Minnesota 55905, USA.
5
Sage Bionetworks, 1100 Fairview Ave. N.,
Seattle, Washington 98109, USA.
6
University of Florida, Center for Translational Research in Neurodegenerative
Diseases, 1275 Center Dr, Gainesville, Florida 32611, USA.
7
Mayo Clinic, Department of Neurology, 200 First
Street, Rochester, Minnesota 55905, USA.
8
Mayo Clinic, Department of Neurology, 4500 San Pablo Road,
Jacksonville, Florida 32224, USA. *These authors contributed equally to this work. Correspondence and requests
for materials should be addressed to N.E.-T. (email: taner.nilufer@mayo.edu).
OPEN
SUBJECT CATEGORIES
» Neurodegeneration
» Genetics of the nervous
system
» Genome-wide
association studies
» RNA sequencing
Received: 8April 2016
Accepted: 31 August 2016
Published: 11 October 2016
www.nature.com/scientificdata
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 1
Background & Summary
In the past decade GWAS identied risk loci for human diseases, including AD
17
and other
neurodegenerative diseases
8,9
. Despite this progress, a comprehensive understanding of the molecular
mechanisms underlying these complex conditions remains elusive. This is partly due to the inability of
the disease GWAS approach to identify the actual disease gene and the functional disease risk variants.
We
10
and others
11,12
utilized combined gene expression GWAS (eGWAS) and disease GWAS to identify
loci which harbor regulatory variants that confer disease risk and to nominate the actual disease genes at
these loci. The underlying premise of these studies is that genetic variants that modulate expression levels
of genes, which encode critical members of disease molecular pathways, will also inuence disease risk
13
.
If this is correct, then there should be signicant overlap between disease GWAS and eGWAS variants,
especially if assessed in the disease-relevant tissue. Indeed, in an eGWAS of brain tissue from subjects
with AD and non-AD, comprised largely of other neurodegenerative diagnoses, we identied signicant
enrichment for disease GWAS variants for AD and other diseases
10
.We
1418
and others
8,1922
determined
that many of the risk variants for AD and other neurodegenerative diseases inuence brain levels of genes
that are nearby in the genome. These studies implicate the genes that are likely to be involved in disease
pathways, nominate regulatory variants as the functional disease risk factors and provide testable
hypotheses for their downstream effects.
Most large-scale gene expression studies in human brains published to date
10,19,20,23
utilize
microarray-based gene or exon arrays. Despite the versatility, cost-effectiveness and large-scale utility,
this approach has limitations, including restricted dynamic range, lack of probes for all known gene
isoforms and connement of assays to known transcripts. RNA sequencing (RNAseq) provides an
attractive alternative that can surpass these limitations and provide much more in-depth information
about the human transcriptome in a high-throughput manner
24
. To expand our prior work on the
human transcriptome based on microarray approaches and to evaluate gene/exon/isoform levels in a
comparative fashion between AD and other neurodegenerative diseases, we have generated RNAseq data
on brain samples from both a subset of the subjects that underwent microarray transcriptome studies
18
and also an independent cohort. These datasets will be of utility in performing expression quantitative
trait loci (eQTL), expression proling and network analyses to facilitate interpretation of genetic
associations and further understanding of disease-mediated changes in transcriptional regulation.
The present report is a description of the large-scale human genetic, and both microarray- and
RNAseq-based transcriptome datasets we generated. The datasets described in this report have been
made available to the research community through the Accelerating Medicines Partnership in
Alzheimers Disease (AMP-AD) Knowledge Portal (Data Citation 1). The portal is hosted in the
Synapse software platform
25
from Sage Bionetworks as part of a series of datasets developed in support of
the AMP-AD Target Identication and Preclinical Validation Project. The AMP-AD consortium includes
six academic teams that will be generating genomic data from human brain or blood samples collected
from more than 10 cohorts. Datasets are hosted in a common environment with standardized meta-data
and annotations to facilitate cross-cohort query, access, and analysis. Each dataset provides a unique
perspective on AD; therefore, datasets differ in types, generation protocols, and underlying patient
characteristics. Together, this collection represents to date the most comprehensive collection of human
genomic data in the eld and, as such, it will be invaluable to a broad set of researchers.
The datasets described herein include the following: (1) late-onset AD GWAS
1
(Mayo LOAD GWAS)
on 2,099 subjects (Data Citation 2); (2) Mayo eGWAS
10
on 773 samples from the cerebellum (CER) and
temporal cortex (TCX) brain regions from a subset of Mayo LOAD GWAS participants (Data
Citations 3,4); (3) Mayo Pilot RNAseq
18
generated on a subset of 186 TCX samples from the Mayo
eGWAS (Data Citation 5); (4) Mayo RNAseq on an independent cohort of 556 TCX
26
(Data Citation 6)
and CER (Data Citation 7) samples from subjects with AD, progressive supranuclear palsy (PSP),
pathologic aging and elderly controls without neurodegenerative diseases. This report provides a
comprehensive understanding of these cohorts, a detailed description of subjects, samples, data
generation, and quality control (QC) as well as instructions to access these rich datasets by the scientic
community.
Methods
The repository of human whole genome genotype and transcriptome data described herein (Table 1,
Fig. 1) consist of the following resources some of which have previously been published: Previously
published datasets include whole genome genotype data from the Mayo LOAD GWAS
1
(Data Citation 2)
and microarray-based whole transcriptome data from the Mayo eGWAS
10
(Data Citations 3,4).
Next-generation RNA-sequencing (RNAseq) data from a subset of the patients from the Mayo Clinic
eGWAS, referred to as the Mayo Pilot RNAseq(Data Citation 5), was published in part
18
.
A non-overlapping cohort with RNAseq-based transcriptome data named Mayo RNAseq(Data
Citations 6,7) has also been published in part
26
. For a comprehensive description of the overall repository,
the data from the published studies are also described herein, albeit in an abbreviated fashion. These four
study cohorts will be referred to by their names as mentioned above, preceded by letters A-D (Table 1)
henceforth.
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 2
Study Populations
All of this work was approved by the Mayo Clinic Institutional Review Board. All human subjects or their
next of kin provided informed consent. The characteristics of the four study populations are as follows:
Mayo LOAD GWAS. The characteristics of the cohort for this study (Data Citation 2) were previously
described in detail
1
. Briey, this is a LOAD case versus control study composed in total of 2,099 subjects
sourced from three different series, namely: Mayo Clinic Jacksonville, Mayo Clinic Rochester and Mayo
Clinic Brain Bank series. These series are respectively termed as JS, RS and AUT in the GWAS
publication
1
(Table 1). Subjects in the Mayo Clinic Jacksonville and Mayo Clinic Rochester series were
diagnosed clinically. These series consisted of 353 LOAD cases versus 331 controls; and 245 LOAD cases
versus 701 controls. The Mayo Clinic Brain Bank series is a post-mortem cohort that consists of 246
LOAD cases versus 223 controls. All subjects were North American Caucasians. All clinical LOAD
subjects were diagnosed as probable or possible AD, according to NINCDS-ADRDA criteria
27
. All
clinical controls had a clinical dementia rating score of 0. LOAD subjects in the Mayo Clinic Brain Bank
series met neuropathologic criteria for denite AD and had a Braak score of 4.0 (ref. 28), while controls
did not meet neuropathologic criteria for AD, and each had Braak score of 2.5, which is an intermediary
level of neurobrillary tangle pathology between Braak score of 2 and 3; but most controls had
neuropathologies unrelated to AD, including vascular dementia, frontotemporal dementia, dementia with
Lewy bodies, multi-system atrophy, amyotrophic lateral sclerosis, and progressive supranuclear palsy.
Ages, APOE ε4genotype and sex distribution for the Mayo LOAD GWAS cohort are shown in Table 2.
This study only included subjects with ages between 60 and 80 years, based on the assumption that much
of the genetic risk for LOAD will be concentrated in this age group, especially given the
age-dependent effects of the strongest AD risk variant apolipoprotein E ε4(APOE4)
28
. Age for the
clinically diagnosed LOAD cases is dened as age at rst diagnosis of AD, since age at onset is not always
available. Age at entry into the study is used for the clinically diagnosed controls. Age at death is utilized
for the cases and controls in the postmortem Mayo Clinic Brain Bank series, given that for this
cohort, age at clinical diagnosis/ evaluation is not always available. Illumina Hap300 microarray genotypes
from the subjects in these three case-control series were utilized to conduct a GWAS of LOAD risk
1
.
Mayo eGWAS. This cohort was previously described in detail
10
. All subjects in the Mayo eGWAS
(Data Citations 3,4) are a subset of the Mayo Clinic Brain Bank series from the Mayo LOAD GWAS
Study Name Brief Description Study Cohort/
Sample type
N Cohort Characteristics Datatype Platform Reference
A. Mayo LOAD
GWAS (Data
Citation 2)
LOAD Case control GWAS. Uses
samples from 3 cohorts: Total 2,099
subjects (Post-QC). This data is used to
identify loci associated with LOAD risk.
Mayo Clinic
Jacksonville (JS)/
Antemortem
N=353 cases, 331
controls
Clinical: AD Cases and Controls,
collected at Mayo Clinic Jacksonville.
Age at rst diagnosis of AD or age at
study entry: 6080.
LOAD GWAS
Genotypes,
demographics
Illumina Hap 300
Carrasquillo
et al.
1
,
Nature
Genetics
Mayo Clinic
Rochester (RS)/
Antemortem
N=245 cases, 701
controls
Clinical: AD Cases and Controls,
collected at Mayo Clinic Rochester.
Age at rst diagnosis of AD or age at
study entry: 6080.
Mayo Clinic Brain
Bank (AUT)/
Postmortem
N=246 cases, 223
controls
Post-mortem: AD Cases (Braak 4.0)
and Other Pathologies (Braak 2.5).
Age at death: 6080.
B. Mayo eGWAS
(Data Citations 3,4)
WG-DASL gene expression measures
for a subset of Mayo Brain Bank
subjects that were included in the Mayo
LOAD GWAS: RNA was isolated from
two brain regions: TCX and CER. This
data is utilized to identify loci associated
with brain gene expression in subjects
with AD, subjects with Other brain
pathologies that do not meet criteria for
AD (Non-AD), and the combined
cohort.
Mayo Brain Bank/
Temporal Cortex
N=202 AD, 197
Non-AD controls
Post-mortem: AD Cases (Braak 4.0)
and Other Pathologies (Braak 2.5).
Age at death: 6080.
Gene expression
phenotypes,
eGWAS results,
covariates
Illumina
WG-DASL
Zou et al.
10
,
PLoS
Genetics
Mayo Brain Bank/
Cerebellum
N=197 AD, 177
Non-AD controls
C. Mayo Pilot
RNAseq (Data
Citation 5)
RNAseq gene expression measures for a
subset of Mayo Brain Bank subjects that
were included in the Mayo LOAD
GWAS: RNA was isolated from TCX.
This data is utilized to identify loci
associated with brain gene expression in
subjects with AD and subjects with PSP.
Mayo Brain Bank/
Temporal Cortex
N=94 AD, 92 PSP Post-mortem: AD Cases (Braak 4.0)
and pathologic diagnosis of PSP
(Braak 2.5). Age at death: 6080.
Gene expression
phenotypes,
covariates
IlluminaHiSeq2000,
50 bp, paired end
RNAseq
Allen
et al.
18
,
Neurology:
Genetics
D. Mayo RNAseq
(Data Citations 6,7)
RNAseq gene expression measures for
subjects from the Mayo Brain Bank
non-overlapping with the Mayo LOAD
GWAS, and also from Banner Sun
Health Institute. RNA was isolated from
two brain regions: TCX and CER. This
data is utilized to compare brain gene
expression between different pairwise
diagnostic groups.
Mayo Brain Bank
and Banner Sun
Health/Temporal
Cortex
N=84 AD, 84 PSP,
30 pathologic
aging, 80 controls Post-mortem: AD Cases (Braak 4.0),
pathologic diagnoses of PSP
(Braak 3), pathologic aging
(Braak 3) and elderly control brains
(Braak 3) without neurodegenerative
diagnoses. Age at death 60.
Gene expression
phenotypes,
covariates
IlluminaHiSeq2000,
101 bp, paired end
RNAseq
NA
Mayo Brain Bank
and Banner Sun
Health/Cerebellum
N=86 AD, 84
PSP, 28
pathologic aging,
80 controls
Table 1. Meta-data for each of the four studies.
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 3
(Data Citation 2) (Fig. 1). The Mayo eGWAS is a whole transcriptome expression study in which brain
samples from two different regions were analyzed, namely cerebellum (CER), which is relatively spared in
AD, and temporal cortex (TCX), which is typically one of the rst regions to be affected with AD
neuropathology
29
. Transcriptome measurements were obtained from TCX of 202 AD subjects and from
CER of 197 AD (Table 1). This study also included subjects without AD neuropathology, which are
referred to as non-AD, given that many of these subjects had other neuropathologies. There were 197
non-AD subjects with TCX transcriptome measurements with the following neuropathologic diagnoses:
progressive supranuclear palsy (PSP, n=107); Lewy body disease (LBD, n=25); corticobasal
degeneration (CBD, n=22); frontotemporal lobar degeneration (FTLD, n=16); multiple system atrophy
(MSA, n=11), vascular dementia (VaD, n=6); other (n=10). There were 177 non-AD subjects with
CER transcriptome measurements that had the following neuropathologies: PSP (n=98); LBD (n=23);
CBD (n=22); FTLD (n=15); MSA (n=7); VaD (n=4); other (n=8). Eighty-ve percent of the subjects
in the TCX cohort overlapped with those in the CER cohort. Demographics for the Mayo eGWAS
subjects and samples, including RNA quality as assessed by RNA Integrity Numbers (RIN) are shown in
Table 2.
Mayo Pilot RNAseq. All subjects in the Mayo Pilot RNAseq study (Data Citation 5) are a subset of the
Mayo eGWAS (Data Citations 3,4), and are therefore also participants of the Mayo Clinic Brain Bank
series that was included in the Mayo LOAD GWAS (Data Citation 2) (Fig. 1). The diagnostic categories
in the Mayo Pilot RNAseq consist of 94 subjects with AD neuropathology and 92 PSP subjects, previously
described
18,26
. PSP is a primary tauopathy characterized neuropathologically by neurobrillary tangles
(NFT) and tau-positive glial lesions
29,30
; and often presents clinically as a parkinsonian disorder. All PSP
A. Mayo LOAD GWAS
(n=2,099)
(Data Citation 2)
1 post-mortem cohort:
Mayo Clinic Brain Bank
(n=469)
2 ante-mortem cohorts:
Mayo Clinic Jacksonville
(n=684) and Rochester
(n=946).
TCX samples with WG-
DASL gene expression
(n=197 AD, 177 non-AD)
CER samples with WG-
DASL gene expression
(n=202 AD, 197 non-AD)
B. Mayo eGWAS
(n=773)
(Data Citation 3,4)
C. Mayo Pilot RNAseq
(n=94 AD, 92 PSP)
(Data Citation 5)
D. Mayo RNAseq
(n=556)
(Data Citation 6,7)
TCX samples
(n=84 AD, 84 PSP, 30
pathologic aging, 80
controls)
CER samples
(n=86 AD, 84 PSP, 28
pathologic aging, 80
controls)
Figure 1. Overview of the relationship of the four genomic datasets herein described.
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 4
subjects were diagnosed neuropathologically by a single neuropathologist (DWD). For this study, only
TCX samples were assessed (Table 2).
Mayo RNAseq. The subjects from this cohort are non-overlapping with the cohorts described above.
The Mayo RNAseq cohort was utilized to generate RNAseq-based whole transcriptome data from 278
TCX
26
(Data Citation 6) and 278 CER (Data Citation 7) samples. Two hundred thirty-eight subjects had
both CER and TCX RNAseq and the rest had either CER or TCX RNAseq measurements based on tissue
availability. CER samples were from the following diagnostic categories: 86 AD, 84 PSP, 28 pathologic
aging and 80 controls without neurodegenerative diagnoses. TCX samples had the following diagnostic
groups: 84 AD, 84 PSP, 30 pathologic aging and 80 controls. Control subjects each had Braak
28
NFT stage
of 3.0 or less, CERAD
31
neuritic and cortical plaque densities of 0 (none) or 1 (sparse) and lacked any of
the following pathologic diagnoses: AD, Parkinsons disease (PD), DLB, VaD, PSP, motor neuron disease
(MND), CBD, Picks disease (PiD), Huntingtons disease (HD), FTLD, hippocampal sclerosis (HipScl) or
dementia lacking distinctive histology (DLDH). Subjects with pathologic aging also lacked the above
diagnoses and had Braak NFT stage of 3.0 or less, but had CERAD neuritic and cortical plaque densities
of 2 or more. None of the pathologic aging subjects had a clinical diagnosis of dementia or mild cognitive
impairment. Given the presence of amyloid plaques, but not tangles and the absence of dementia,
pathologic aging is considered to be either a prodrome of AD or a condition, in which there is resistance
to the development of NFT and/or dementia
32
.
Within the Mayo RNAseq cohort (Data Citations 6,7), all AD and PSP subjects were from the Mayo
Clinic Brain Bank, and all pathologic aging subjects were obtained from the Banner Sun Health Institute.
Thirty-four control CER and 31 control TCX samples were from the Mayo Clinic Brain Bank, and the
remaining control tissue was from the Banner Sun Health Institute. All subjects were North American
Caucasians. All but control subjects, had ages at death 60, and a more relaxed lower age cutoff of 50
was applied for normal controls to achieve sample sizes similar to that of AD and PSP subjects. No upper
age limit was imposed on this cohort, however when subjects had ages at death of 90, their ages were
recorded as 90_or_aboveand shown as 90in Table 2 to protect patient condentiality.
Table 2 details the demographic characteristics of the Mayo RNAseq cohort (Data Citations 6,7).
PSP subjects tended to be younger than the other diagnostic groups. As expected, there was a greater
frequency of APOE4 positive subjects in the AD group, followed by pathologic aging, then PSP and
control subjects. AD and pathologic aging subjects had greater female sex frequency (57%), followed by
controls (49%), then PSP subjects (39%). RIN for all samples were selected to be 5.0. Pathologic aging
and control samples had slightly lower RINs than AD and PSP samples, due to limitations in availability
of samples in these former diagnostic categories.
Molecular Data
Sample collection and processing. For the Mayo LOAD GWAS (A) (Data Citation 2), DNA samples
were collected and processed as previously described
1
. For the antemortem Mayo Clinic Jacksonville and
Mayo Clinic Rochester series, whole blood samples were collected in 10 ml EDTA tubes followed by DNA
A. Mayo LOAD GWAS (Data Citation 2) B. Mayo eGWAS (Data Citations 3,4) C. Mayo Pilot RNAseq (Data Citation
5)
TCX CER TCX
Variables AD (n=844) CON (1,255) AD (n=202) NON-AD
(n=197)
AD (n=197) NON-AD (n=177) AD (n=94) PSP (N=92)
Mean Age ±s.d. (Range) 74.0 ±4.8 (6080) 73.2 ±4.4 (6080) 73.6 ±5.5 (6080) 71.6 ±5.6 (6080) 73.6 ±5.6 (6080) 71.7 ±5.5 (6080) 74.1 ±5.7 (6080) 71.9 ±5.4 (6080)
APOE4 positive/negative/
null
(%APOE4 positive)
549/277/18 (65%) 344/889/22 (27%) 123/79/0 (61%) 49/146/2 (25%) 126/71/0 (64%) 45/130/2 (25%) 58/36/0 (62%) 20/72/0 (22%)
Female (%) 482 (57%) 641 (51%) 108 (53%) 78 (40%) 101 (51%) 63 (36%) 41 (44%) 37 (40%)
Mean RIN ±s.d. (Range) NA NA 6.3±0.9 (59) 6.9 ±1.0 (59.3) 7.2 ±1.0 (59.4) 7.2 ±1.0 (59) 7.0 ±0.7 (6.29) 7.0 ±0.9 (5.79.3)
D. Mayo RNAseq (Data Citations 6,7)
TCX CER
Variables AD (n=84) PSP (n=84) Path Aging
(n=30)
Control (n=80) AD (n=86) PSP (n=84) Path Aging
(n=28)
Control (n=80)
Mean Age ±s.d. (Range) 82.4 ±7.7 (6090) 74.0 ±6.5 (6189) 85.2 ±4.3 (7690) 82.6 ±8.8 (5390) 82.5 ±7.7 (6090) 74.0 ±6.5 (6189) 84.7 ±4.3 (7690) 82.5 ±8.3 (5890)
APOE4 positive/negative
(%APOE4 positive)
43/41 (51%) 13/71 (15%) 10/20 (33%) 10/70 (13%) 43/43 (50%) 13/71 (15%) 9/19 (32%) 11/69 (14%)
Female (%) 48 (57%) 33 (39%) 17 (57%) 39 (49%) 49 (57%) 33 (39%) 16 (57%) 39 (49%)
Mean RIN ±s.d. (Range) 8.6 ±0.5 (7.710.0) 8.5 ±0.5 (7.810.0) 7.4 ±1.0 (5.38.9) 7.6 ±1.0 (5.39.7) 8.3 ±0.8 (5.710.0) 8.4 ±0.9 (5.510.0) 7.5±1.0 (5.79.0) 7.6 ±1.0 (5.59.7)
Table 2. Demographics for the cohorts included in each of the four studies.
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 5
extraction using AutoGenFlex STAR instrument (AutoGen), whereas cerebellar tissue was used for DNA
extraction from the postmortem Mayo Clinic Brain Bank series using the Wizard Genomic DNA
purication kit (Promega). Given limited amounts of DNA from samples in the Mayo Clinic Rochester
series and Mayo Clinic Brain Bank series, whole genome amplication (WGA) was applied using the
Illustra GenomiPhi V2 DNA Amplication Kit (GE Healthcare Bio-Sciences), in four 5 ml reactions that
utilized 515 ng genomic DNA as a template. Subsequent to the pooling of these reaction products, WGA
DNA was subjected to quality control (QC) using SNP genotyping as previously described.
RNA extraction methods for the Mayo eGWAS
10
(B) (Data Citations 3,4) and Mayo Pilot RNAseq
18
(C) (Data Citation 5) were previously described. Total RNA was extracted from frozen brain samples
using the Ambion RNAqueous kit (Life Technologies, Grand Island, NY) according to the manufacturers
instructions. Brain samples for the Mayo RNAseq (D) (Data Citations 6,7) study underwent RNA
extractions via the Trizol/chloroform/ethanol method, followed by DNase and Cleanup of RNA using
Qiagen RNeasy Mini Kit and Qiagen RNase -Free DNase Set. The quantity and quality of all RNA
samples were determined by the Agilent 2100 Bioanalyzer using the Agilent RNA 6000 Nano Chip
(Agilent Technologies, Santa Clara, CA). Samples had to have an RNA Integrity Number (RIN) 5.0 for
inclusion in either study (Table 2).
Data generation. The genotype data for the Mayo LOAD GWAS (A) (Data Citation 2) was generated
using HumanHap300-Duo Genotyping BeadChips
1
, which were processed with an Illumina BeadLab
station at the Mayo Clinic Genotyping Shared Resource (currently Mayo Clinic Medical Genome
Facility =MGF, Rochester, Minnesota) according to the manufacturers protocols. Two samples were
genotyped per chip for 318,237 SNPs across the genome. Genotype calls were made using the auto-calling
algorithm in Illuminas BeadStudio 2.0 software.
For the Mayo eGWAS study (B) (Data Citations 3,4), transcript levels were measured using the Whole
Genome DASL assay (Illumina, San Diego, CA) as previously described
10
. Probe annotations were done
based on NCBI RefSeq, Build 36.2. The RNA samples were randomized across the chips and plates using
a stratied approach to ensure balance with respect to diagnosis, age, gender, RIN and APOE genotype.
Raw probe mRNA expression data were exported from GenomeStudio software (Illumina Inc.) and
preprocessed for background correction, variance stabilizing transformation, quantile normalization and
probe ltering using the lumi package of BioConductor
33
.
Samples for both Mayo Pilot RNAseq (C) (Data Citation 5) and Mayo RNAseq (D) (Data
Citations 6,7) studies were randomized prior to transfer to the Mayo Clinic MGF Gene Expression Core
for library preparation and then the Sequencing Core for RNA sequencing. Mayo Pilot RNAseq (C)
(Data Citation 5) AD and PSP samples were randomized across owcells, taking into account age at
death, sex and RIN. These samples underwent library preparation and sequencing at different times and
therefore should be considered as separate datasets. Likewise, Mayo RNAseq (D) of TCX
26
and CER
samples (Data Citations 6,7, respectively) underwent RNAseq at different times. These samples were
randomized across owcells, taking into account age at death, sex, RIN, Braak stage and diagnosis. The
TruSeq RNA Sample Prep Kit (Illumina, San Diego, CA) was used for library preparation from all
samples. The library concentration and size distribution was determined on an Agilent Bioanalyzer DNA
1000 chip. All samples were run in triplicates using barcoding (3 samples per owcell lane). For Mayo
Pilot RNAseq (C) (Data Citation 5) samples, 50 base-pair, paired-end sequencing was done, whereas
Mayo RNAseq (D) (Data Citations 6,7) samples underwent 101 bp, paired-end sequencing.
Data Processing. Mayo LOAD GWAS (A) (Data Citation 2) genotypes from Illumina BeadStudio 2.0
software were utilized to generate lgen, map and fam les that were imported into PLINK
34
and
converted to binary ped (.bed) and map (.bim) les, which are deposited together with PLINK format fam
and covariate les (DOI and descriptions for each these les are provided in Table 3 (available online
only)).
The Mayo eGWAS WG-DASL microarray expression dataset from TCX and CER (B) includes
covariates and probe expression levels (Data Citation 3), which are preprocessed as published
10
and
described above. The Mayo eGWAS eSNP Results(Data Citation 4) are the eQTL results from the test of
association between the Mayo LOAD GWAS (Data Citation 2) genotypes and the WG-DASL gene
expression measures analyzed by multivariable linear regression using an additive model in PLINK
34
,as
published previously
10
(DOI and descriptions for each these les are provided in Table 3 (available online
only)). These analysis used preprocessed probe transcript levels as traits, SNP minor allele dosage as the
independent variable, and adjusted for the following covariates: APOE ε4 dosage (0, 1, 2), age at death,
sex, PCR plate, RIN and adjusted RIN squared (RIN-RINmean)
2
. Analyses were limited to SNP-probe
pairs that were in-cis,dened as +/ 100 kb of the targeted gene according to NCBI Build 36. The ADs
and nonADs were analyzed both separately and jointly. The joint analyses included diagnosis as an
additional covariate (AD =1, nonAD =0). Results of analyses for both the genotyped SNPs as well as
genotypes imputed to HapMap2 reference are provided. HapMap2 imputations were done as described
10
.
The eGWAS results were previously made available through the NIAGADS repository (https://www.
niagads.org/datasets/ng00025).
The Mayo Pilot RNAseq
18
(Data Citation 5), Mayo RNAseq TCX
26
and CER data (Data Citations 6,7,
respectively) were processed using the same analytic pipeline. Read alignments were done using the
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 6
SNAPR software
35
, an RNA sequence aligner based on SNAP, using GRCh38 reference and Ensembl v77
gene models. Outputs include per-sample gene and transcript counts, which are merged into a single le
per data type (gene or transcript) that contains data for all samples across all genes/transcripts (DOI and
descriptions for each these les are provided in Table 3 (available online only)). Alignment with SNAPR
starts with the creation of hash indices built from both a reference genome GRCh38 and transcriptome
GRCh38.77. SNAPR lters fastq reads by Phred score (>80% of the read must have a Phred score
>= 20) and simultaneously aligns each read (or read pair) to both the genome and transcriptome. The
best alignment is written to a sorted BAM le with read counts simultaneously tabulated and written for
each sample. Read counts are given by gene ID and transcript ID (two separate les). We have previously
tested the read counts generated by SNAPR to the read counts generated by HT-Seq and found them to
be very comparable.
Post-processing was also performed using the same pipeline for these three RNAseq datasets as follows:
The individual read count les produced by SNAPR are merged into a single le using two scripts:
merge_count_les.R and a dataset-specic read-count merge script. These scripts generate the
corresponding _counts.txt.gz les. The merged count les are normalized with the normalize_
readcounts.R script, which uses the edgeR implementation of the trimmed mean of M-values (TMM)
normalization method to calculate counts per million (CPM). These normalized counts are saved for both
gene and transcript levels (DOI and descriptions for each these les are provided in Table 3 (available
online only)).
Code Availability. The R script called merge_count_les.R
36
was used to merge the RNAseq read
count les produced by SNAPR into a single le, and can be found at https://github.com/CoryFunk/
AMP-AD-scripts/blob/master/combine_count_les.pl. Also, the R script used to normalize the merged
RNAseq read counts, called normalize_readcounts.R
36
, can be found at https://github.com/CoryFunk/
AMP-AD-scripts/blob/master/tmm_normalization.R.
Data Records
Data available for studies A-D (Data Citations 27; Table 3 (available online only)) consists of a set of
les that contain genomic, genetic or covariate data for a dened set of samples; analytic results are also
provided when available. Data les can be found in the Sage Bionetworks AMP-AD Knowledge Portal
(Data Citation 1) in study specic folders (and subfolders). Users can identify and search for data les
and data descriptions using the unique Synapse ID and corresponding DOI provided in Table 3 (available
online only). Each sample within a study has a unique sample ID, this sample ID is consistent across all
les within the study, and les in other studies where applicable. The relationship between studies
and sample overlaps is illustrated in Fig. 1. The samples in study C (Data Citation 5) are a subset of
the samples in study B (Data Citation 3) which are likewise a subset of the samples in study A (Data
Citation 2); the samples in study D (Data Citations 6,7) are independent of those in studies A-C. The
Usage Notes section describes the data accession conditions, and the steps for requesting access.
Technical Validation
Data QC
Mayo LOAD GWAS (A) (Data Citation 2) QC methods were previously published
1
. Briey, using
PLINK
34
, subjects with genotyping call rates of o90%, duplicate genotyping and/or sex-mismatches
between recorded and deduced sex were eliminated from the dataset. All SNPs with genotyping call rates
o90%, minor allele frequencies o0.01, and/or Hardy-Weinberg p values o0.001 were also eliminated.
Prior to QC, 318,237 SNPs were genotyped in 2,465 subjects. The available data includes the 313,504 SNP
genotypes from 2,099 subjects that passed these QC parameters.
The Mayo eGWAS
10
(B) (Data Citations 3,4) data was generated as follows: We annotated probes for
presence of genetic variants by comparing their positions according to NCBI RefSeq, Build 36.3 to those
of all variants within dbSNP131 and identied the list of probes that have 1 variants within their
sequence. We depict this information in the les for the Mayo eGWAS, eSNP Results(Data Citation 4)
(Table 3 (available online only)), by including SNP-In-Probecolumn, which has TRUEif the probe
sequence harbors 1 SNP, and FALSE, otherwise. We also calculated for each probe within each analytic
group, percent detection rate above background. Probes that are detected in >12.5%, >25%, >50% and
>75% of the subjects in each analytic group are annotated by four separate columns within the eSNP
Results(Data Citation 4) from the eGWAS that included HapMap2 imputed genotypes, described below.
The purpose of these annotation columns is to enable others the exibility to impose cutoffs based on
presence/absence of variants within probe sequence and/or probe detection rates while providing the full
dataset for completeness. The Mayo eGWAS (Data Citation 3,Data Citation 4) also included replicate
samples as described for QC and to estimate intraclass coefcients (ICC), which is the between-subject
variance, as a percentage of the total variance in probe expression
10
. There were 4 AD and 4 non-AD
temporal cortex samples that were measured in 5 replicates; and 10 AD and 5 non-AD cerebellar sample
replicates across ve plates. Universal human RNA (UHR) samples were also run on each PCR plate as
part of QC. The expression phenotypes include results from only one of the replicate subjects selected
randomly and exclude UHR results. It should be noted that 3 AD and 9 non-AD subjects for TCX, and
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 7
4 AD subjects for CER, do not have associated GWAS genotypes as they did not pass 1 GWAS QC
parameter described above.
For the Mayo Pilot RNAseq
18
(C) (Data Citation 5) data principal components analysis (PCA)
identied 2 outliers in the AD and 4 in the PSP cohort. The covariates for these subjects were set to
missing ( =NA) in the respective covariate les (DOI and descriptions for these les are provide in
Table 3 (available online only)). Hence, although 96 AD and 96 PSP subjects underwent sequencing in
the Mayo Pilot RNAseq study, 94 AD and 92 PSP subjects were retained for analyses. It should be noted
that of these subjects, 1 AD and 7 PSP subjects lack GWAS data due to either having genotype counts
o90% or failing sex checks. PCA identied no outliers in the Mayo RNAseq (D) of TCX
26
samples (Data
Citation 6) but 2 such subjects in the CER analyses (Data Citation 7). The covariate data in the relevant
CER les for these two subjects were set to missing. We likewise assessed the RNASeq data for sex
discrepancies based on Y chromosome gene expression and documented sex and identied 2 subjects
with mis-matched sex for both TCX and CER, plus a third subject in the CER cohort. These were also set
to missing in the covariate les. At the time of this publication, the Mayo RNAseq subjects did not have
GWAS genotypes deposited on Synapse.
Usage Notes
The data described herein is available for use by the research community and has been deposited in the
AMP-AD Knowledge Portal (Data Citation 1). Table 3 (available online only) provides a detailed
description of the les deposited for the four studies, their specic Synapse identiers (IDs), DOIs, the
types of les and denitions of the column headers. These les (Data Citations 27), and their assigned
DOIs will be maintained in perpetuity in the AMP-AD Knowledge Portal (Data Citation 1). Access to all
of these les is enabled through the Sage Bionetworks, Synapse repository; and a subset of the les for the
Mayo LOAD GWAS (Data Citation 2) and the Mayo eGWAS (Data Citations 3,4) are also available via
NIAGADS (www.niagads.org).
The AMP-AD Knowledge Portal hosts data derived from multiple cohorts that were generated as part
of or used in support of the AMP-AD Target Identication and Preclinical Validation project
(Data Citation 1). The portal uses the Synapse software platform
25
for backend support, providing users
with both web-based and programmatic access to data les. All data les in the portal are annotated using
a standard vocabulary to enable users to search for relevant content across the AMP-AD datasets using
programmatic queries. Data is stored in a cloud based manner hosted by Amazon web services (AWS),
which enables user to execute cloud-based compute. Detailed descriptions including data processing,
QC metrics, and assay and cohort specic variables are provided for each le as applicable.
Access for the data described herein is controlled in a manner set forth by the institutional review
board (IRB) at the Mayo Clinic. All data use terms include: (1) maintenance of data in a secure and
condential manner, (2) respect for the privacy of study participants, (3) citation of the data contributors
in any publications resulting from data use, and (4) informing data contributors of resultant publications.
Specic data use terms are provided for each dataset (Data Citations 36) under the header Terms of
use; users must register for a Synapse account and provide electronic agreement to these terms prior to
accessing the study les. Access to the Mayo LOAD GWAS data (A) (Data Citation 2) requires a data use
certicate (doi:10.7303/syn2954402.2). User approvals are managed by the Synapse Access and
Compliance Team (ACT).
Data on the AMP-AD Knowledge Portal are annotated with a common dictionary of terms
(doi:10.7303/syn5478487.2) to enable querying of the data using the Synapse analytical clients (R client:
syn1834618, python client: syn1768504, command line client: syn2375225). Fields, their allowable values
specic to the datasets described herein and the dictionary of annotations are shown in Table 3 (available
online only). These annotations can be used to identify les of interest within the available datasets and to
lter on any of the elds using the allowable values from the dictionary (an example is shown here:
doi:10.7303/syn5585666.1).
References
1. Carrasquillo, M. M. et al. Genetic variation in PCDH11X is associated with susceptibility to late-onset Alzheimer's disease. Nat
Genet 41, 192198 (2009).
2. Harold, D. et al. Genome-wide association study identies variants at CLU and PICALM associated with Alzheimer's disease.
Nature genetics 41, 10881093 (2009).
3. Lambert, J. C. et al. Genome-wide association study identies variants at CLU and CR1 associated with Alzheimer's disease.
Nature genetics 41, 10941099 (2009).
4. Seshadri, S. et al. Genome-wide analysis of genetic loci associated with Alzheimer disease. Jama 303, 18321840 (2010).
5. Naj, A. C. et al. Common variants at MS4A4/MS4A6E, CD2AP, CD33 and EPHA1 are associated with late-onset Alzheimer's
disease. Nature genetics 43, 436441 (2011).
6. Hollingworth, P. et al. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with
Alzheimer's disease. Nature genetics 43, 429435 (2011).
7. Lambert, J. C. et al. Meta-analysis of 74,046 individuals identies 11 new susceptibility loci for Alzheimer's disease. Nature
genetics 45, 14521458 (2013).
8. Hoglinger, G. U. et al. Identication of common variants inuencing risk of the tauopathy progressive supranuclear palsy. Nature
genetics 43, 699705 (2011).
9. Simon-Sanchez, J. et al. Genome-wide association study reveals genetic risk underlying Parkinson's disease. Nature genetics 41,
13081312 (2009).
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 8
10. Zou, F. et al. Brain expression genome-wide association study (eGWAS) identies human disease-associated variants. PLoS Genet
8, e1002707 (2012).
11. Dixon, A. L. et al. A genome-wide association study of global gene expression. Nature genetics 39, 12021207 (2007).
12. Emilsson, V. et al. Genetics of gene expression and its effect on disease. Nature 452, 423428 (2008).
13. Saykin, A. J. et al. Genetic studies of quantitative MCI and AD phenotypes in ADNI: Progress, opportunities, and plans.
Alzheimers Dement 11, 792814 (2015).
14. Zou, F. et al. Gene expression levels as endophenotypes in genome-wide association studies of Alzheimer disease. Neurology 74,
480486 (2010).
15. Allen, M. et al. Novel late-onset Alzheimer disease loci variants associate with brain gene expression. Neurology 79,
221228 (2012).
16. Allen, M. et al. Glutathione S-transferase omega genes in Alzheimer and Parkinson disease risk, age-at-diagnosis and brain gene
expression: an association study with mechanistic implications. Mol Neurodegener 7, 13 (2012).
17. Allen, M. et al. Association of MAPT haplotypes with Alzheimer's disease risk and MAPT brain gene expression levels.
Alzheimers Res Ther 6, 39 (2014).
18. Allen, M et al. Late-onset Alzheimer disease risk variants mark brain regulatory loci. Neurology: Genetics 1, e15 (2015).
19. Myers, A. J. et al. A survey of genetic human cortical gene expression. Nature genetics 39, 14941499 (2007).
20. Webster, J. A. et al. Genetic control of human brain transcript expression in Alzheimer disease. Am J Hum Genet 84,
445458 (2009).
21. Chapuis, J. et al. Increased expression of BIN1 mediates Alzheimer genetic risk by modulating tau pathology. Mol Psychiatry 18,
12251234 (2013).
22. Hazrati, L. N. et al. Genetic association of CR1 with Alzheimer's disease: a tentative disease mechanism. Neurobiol Aging 33,
2949 e52949 e12 (2012).
23. Ramasamy, A. et al. Genetic variability in the regulation of gene expression in ten regions of the human brain. Nat Neurosci 17,
14181428 (2014).
24. Montgomery, S. B. et al. Transcriptome genetics using second generation sequencing in a Caucasian population. Nature 464,
773777 (2010).
25. Derry, J. M. et al. Developing predictive molecular maps of human disease through community-based modeling. Nature genetics
44, 127130 (2012).
26. Allen, M. et al. Gene expression, methylation and neuropathology correlations at progressive supranuclear palsy risk loci. Acta
Neuropathol 132, 197211 (2016).
27. McKhann, G et al. Clinical diagnosis of Alzheimer's disease: report of the NINCDS-ADRDA Work Group under the auspices of
Department of Health and Human Services Task Force on Alzheimer's Disease. Neurology 34, 939944 (1984).
28. Farrer, LA et al. Effects of age, sex, and ethnicity on the association between apolipoprotein E genotype and Alzheimer disease.
A meta-analysis. APOE and Alzheimer Disease Meta Analysis Consortium. Jama 278, 13491356 (1997).
29. Braak, H. & Braak, E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol (Berl) 82, 239259 (1991).
30. Hauw, J. J. et al. Preliminary NINDS neuropathologic criteria for Steele-Richardson-Olszewski syndrome (progressive
supranuclear palsy). Neurology 44, 20152019 (1994).
31. Mirra, S. S. et al. Interlaboratory comparison of neuropathology assessments in Alzheimer's disease: a study of the Consortium to
Establish a Registry for Alzheimer's Disease (CERAD). J Neuropathol Exp Neurol 53, 303315 (1994).
32. Wang, J., Dickson, D. W., Trojanowski, J. Q. & Lee, V. M. The levels of soluble versus insoluble brain Abeta distinguish
Alzheimer's disease from normal and pathologic aging. Exp Neurol 158, 328337 (1999).
33. Du, P., Kibbe, W. A. & Lin, S. M. lumi: a pipeline for processing Illumina microarray. Bioinformatics 24, 15471548 (2008).
34. Purcell, S. et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am J Hum Genet 81,
559575 (2007).
35. Magis, A. T., Funk, C. C. & Price, N. D. SNAPR: A Bioinformatics Pipeline for Efcient and Accurate RNA-Seq Alignment and
Analysis. IEEE Life Sciences Letters 1, 2225 (2015).
36. Funk, C. AMP-AD-scripts: AMP-AD Fl-Mayo-ISB. in Zenodo https://dx.doi.org/10.5281/zenodo.56828 (2016).
Data Citations
1. Synapse http://dx.doi.org/10.7303/syn2580853 (2016).
2. Carrasquillo, M. M. et al. Synapse http://dx.doi.org/10.7303/syn2910256 (2016).
3. Zou, F. et al. Synapse http://dx.doi.org/10.7303/syn3157225 (2016).
4. Zou, F. et al. Synapse http://dx.doi.org/10.7303/syn3157249 (2016).
5. Allen, M. et al. Synapse http://dx.doi.org/10.7303/syn3157268 (2016).
6. Allen, M. et al. Synapse http://dx.doi.org/10.7303/syn3163039 (2016).
7. Allen, M. et al. Synapse http://dx.doi.org/10.7303/syn5049298 (2016).
Acknowledgements
We thank the patients and their families for the sample and tissue donations. Without their generosity,
this research would not be possible. The Mayo Clinic Alzheimer's Disease Genetic Studies were led by
Dr Nilüfer Ertekin-Taner and Dr Steven G. Younkin, Mayo Clinic, Jacksonville, FL using samples from
the Mayo Clinic Study of Aging, the Mayo Clinic Alzheimer's Disease Research Center, and the Mayo
Clinic Brain Bank. Data collection was supported through funding by NIA grants P50 AG016574, R01
AG032990, U01 AG046139, R01 AG018023, U01 AG006576, U01 AG006786, R01 AG025711, R01
AG017216, R01 AG003949, NINDS grant R01 NS080820, the GHR foundation, CurePSP Foundation,
and support from Mayo Foundation. Samples collected through the Sun Health Research Institute Brain
and Body Donation Program of Sun City, Arizona. The Brain and Body Donation Program is supported
by the National Institute of Neurological Disorders and Stroke (U24 NS072026 National Brain and Tissue
Resource for Parkinsons Disease and Related Disorders), the National Institute on Aging (P30 AG19610
Arizona Alzheimers Disease Core Center), the Arizona Department of Health Services (contract 211002,
Arizona Alzheimers Research Center), the Arizona Biomedical Research Commission (contracts 4001,
0011, 05-901 and 1001 to the Arizona Parkinson's Disease Consortium) and the Michael J. Fox
Foundation for Parkinsons Research. We thank Mrs. Kelly Viola for her assistance with revisions of this
manuscript.
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 9
Author Contributions
M.A. helped with draft of the manuscript, analyzed data, contributed to the Mayo eGWAS and oversaw
the Mayo Pilot RNAseq and Mayo RNAseq studies; M.M.C. helped with draft of manuscript, analyzed
data, co-led the Mayo LOAD GWAS, and oversaw the Mayo Pilot RNAseq and Mayo RNAseq studies;
C.F. analyzed data for Mayo Pilot RNAseq and Mayo RNAseq; B.D.H. analyzed data for Mayo Pilot
RNAseq and Mayo RNAseq; F.Z. analyzed data and oversaw the Mayo eGWAS; C.S.Y. analyzed and
databased data for all studies; J.D.B. analyzed data for Mayo eGWAS, Mayo Pilot RNAseq and Mayo
RNAseq; H.-S.C. analyzed data for Mayo eGWAS; J.C. provided statistical support; J.A.E. analyzed data
for Mayo Pilot RNAseq and Mayo RNAseq; H.L. analyzed data for Mayo Pilot RNAseq and Mayo
RNAseq; B.L. architected the data repository, deposited these data into the public portal and manage data
dissemination; M.A.P. architected the data repository, deposited these data into the public portal and
manage data dissemination; K.K.D architected the data repository, deposited these data into the public
portal and manage data dissemination; X.W. analyzed data for Mayo Pilot RNAseq and Mayo RNAseq;
D.S. analyzed data for Mayo eGWAS, Mayo Pilot RNAseq and Mayo RNAseq; C.W. analyzed data for
Mayo eGWAS; T.N. generated data; S.L. generated data; K.M. generated data; G.B. generated data;
M.L. generated data; T.E.G. provided comments for the manuscript; L.M.M. architected the data
repository, deposited these data into the public portal and manage data dissemination; Y.A. analyzed data
for Mayo Pilot RNAseq and Mayo RNAseq; N.P. oversaw bioinformatics analysis of Mayo Pilot RNAseq
and Mayo RNAseq; R.C.P. provided patient material and data; N.R.G.-R. provided patient material and
data; D.W.D. provided patient material and data; S.G.Y. analyzed data, designed and led the Mayo
GWAS, wrote the manuscript; N.E.-T. analyzed data, designed and led the Mayo eGWAS, Mayo Pilot
RNAseq and Mayo RNAseq studies and wrote the manuscript.
Additional information
Table 3 is only available in the online version of this paper.
Competing nancial interests: Below are the disclosures for R.C.P.: Pzer, Inc., and Janssen Alzheimer
Immunotherapy: Chair, Data Monitoring Committee. Hoffman-La Roche, Inc.: Consultant. Merck, Inc.:
Consultant. Genentech, Inc.: Consultant. Biogen, Inc.: Consultant. Eli Lilly & Co.: Consultant. N.R.G.-R.
has multicenter treatment study grants from Lilly and TauRx and consulted for Cytox. N.E.-T. has
consulted for Cytox. The remaining authors declare no competing nancial interests.
How to cite this article: Allen, M et al. Human whole genome genotype and transcriptome data for
Alzheimer's and other neurodegenerative diseases. Sci. Data 3:160089 doi: 10.1038/sdata.2016.89 (2016).
This work is licensed under a Creative Commons Attribution 4.0 International License. The
images or other third party material in this article are included in the articles Creative
Commons license, unless indicated otherwise in the credit line; if the material is not included under the
Creative Commons license, users will need to obtain permission from the license holder to reproduce the
material. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0
Metadata associated with this Data Descriptor is available at http://www.nature.com/sdata/ and is released
under the CC0 waiver to maximize reuse.
© The Author(s) 2016
www.nature.com/sdata/
SCIENTIFIC DATA |3:160089 |DOI: 10.1038/sdata.2016.89 10
... We examined the levels of 250 genes in the MayoPilot RNAseq dataset 55 to identify cellular and molecular processes associated with chronic inflammation in AD. Among cytokines, chemokines, enzymes, protein kinases, transcriptional factors, and cell markers, we identified 40 differentially expressed genes in the TCX of AD versus non-AD subjects (Fig. 1a, Supplementary Table 1). ...
... www.nature.com/scientificreports/ 55 . Differentially expressed genes were determined by unpaired t-test followed by Holm-Sidak's multiple comparisons tests. ...
Article
Full-text available
Astrocytes play a role in healthy cognitive function and Alzheimer’s disease (AD). The transcriptional factor nuclear factor-κB (NF-κB) drives astrocyte diversity, but the mechanisms are not fully understood. By combining studies in human brains and animal models and selectively manipulating NF-κB function in astrocytes, we deepened the understanding of the role of astrocytic NF-κB in brain health and AD. In silico analysis of bulk and cell-specific transcriptomic data revealed the association of NF-κB and astrocytes in AD. Confocal studies validated the higher level of p50 NF-κB and phosphorylated-p65 NF-κB in glial fibrillary acidic protein (GFAP)⁺-astrocytes in AD versus non-AD subjects. In the healthy mouse brain, chronic activation of astrocytic NF-κB disturbed the proteomic milieu, causing a loss of mitochondrial-associated proteins and the rise of inflammatory-related proteins. Sustained NF-κB signaling also led to microglial reactivity, production of pro-inflammatory mediators, and buildup of senescence-related protein p16INK4A in neurons. However, in an AD mouse model, NF-κB inhibition accelerated β-amyloid and tau accumulation. Molecular biology studies revealed that astrocytic NF-κB activation drives the increase in GFAP and inflammatory proteins and aquaporin-4, a glymphatic system protein that assists in mitigating AD. Our investigation uncovered fundamental mechanisms by which NF-κB enables astrocytes' neuroprotective and neurotoxic responses in the brain.
... The copyright holder for this preprint this version posted June 16, 2024. ; built utilizing the 5xFAD model in addition to human samples, creating a much more targeted disease profile [31][32][33]. We identified 15 differentially expressed genes in 5xFAD;cTFEB;HSACre hippocampal RNA lysates when compared to their 5xFAD littermate controls ( Figure 3A). ...
... We identified 15 differentially expressed genes in 5xFAD;cTFEB;HSACre hippocampal RNA lysates when compared to their 5xFAD littermate controls ( Figure 3A). Of these, 12 were down-regulated and only 3 were up-regulated, reflecting the high degree of disease-association built into the panel footprint [31][32][33]. Two of these genes were associated with protein aggregation (Kctd1, potassium channel tetramerization domain containing 1) [34] and Aβ amyloidosis formation (Phyhip, phytanoyl-CoA 2-hydroxylase interacting protein) [35], consistent with our findings of reduced plaque abundance in the CNS of 5xFAD;HSACre mice ( Figure 1C-F). ...
Preprint
Full-text available
INTRODUCTION Skeletal muscle regulates central nervous system (CNS) function and health, activating the muscle-to-brain axis through the secretion of skeletal muscle originating factors (‘myokines’) with neuroprotective properties. However, the precise mechanisms underlying these benefits in the context of Alzheimer’s disease (AD) remain poorly understood. METHODS To investigate muscle-to-brain axis signaling in response to amyloid β (Aβ)- induced toxicity, we generated 5xFAD transgenic female mice with enhanced skeletal muscle function (5xFAD;cTFEB;HSACre) at prodromal (4-months old) and late (8-months old) symptomatic stages. RESULTS Skeletal muscle TFEB overexpression reduced Aβ plaque accumulation in the cortex and hippocampus at both ages and rescued behavioral neurocognitive deficits in 8- months-old 5xFAD mice. These changes were associated with transcriptional and protein remodeling of neurotrophic signaling and synaptic integrity, partially due to the CNS-targeting myokine prosaposin (PSAP). DISCUSSION Our findings implicate the muscle-to-brain axis as a novel neuroprotective pathway against amyloid pathogenesis in AD.
... RNA-Seq data of the Mayo Clinic cohort were obtained from the AD Knowledge Portal. Detailed information was described in the previous reports on the sample collection, post-mortem sample data, tissue and RNA preparation, library preparation and sequencing, and sample quality controls [1]. The whole transcriptome data based on RNA-Seq were generated from temporal cortex (TCX) human samples obtained from the Mayo Clinic TCX RNA-Seq dataset (https:// www. ...
... synap se. org/# !Synap se: syn46 50265) [1]. The normalized RNA expression of elderly controls and AD were analysed from datasets according to the criteria of Braak staging [14]. ...
Article
Full-text available
A recent large genome-wide association study has identified EGFR (encoding the epidermal growth factor EGFR) as a new genetic risk factor for late-onset AD. SHIP2, encoded by INPPL1, is taking part in the signalling and interactome of several growth factor receptors, such as the EGFR. While INPPL1 has been identified as one of the most significant genes whose RNA expression correlates with cognitive decline, the potential alteration of SHIP2 expression and localization during the progression of AD remains largely unknown. Here we report that gene expression of both EGFR and INPPL1 was upregulated in AD brains. SHIP2 immunoreactivity was predominantly detected in plaque-associated astrocytes and dystrophic neurites and its increase was correlated with amyloid load in the brain of human AD and of 5xFAD transgenic mouse model of AD. While mRNA of INPPL1 was increased in AD, SHIP2 protein undergoes a significant solubility change being depleted from the soluble fraction of AD brain homogenates and co-enriched with EGFR in the insoluble fraction. Using FRET-based flow cytometry biosensor assay for tau-tau interaction, overexpression of SHIP2 significantly increased the FRET signal while siRNA-mediated downexpression of SHIP2 significantly decreased FRET signal. Genetic association analyses suggest that some variants in INPPL1 locus are associated with the level of CSF pTau. Our data support the hypothesis that SHIP2 is an intermediate key player of EGFR and AD pathology linking amyloid and tau pathologies in human AD. Supplementary Information The online version contains supplementary material available at 10.1007/s00401-024-02745-7.
... This study contains RNA sequencing data from cerebellum and temporal cortex samples from North American Caucasian subjects with neuropathological diagnosis of AD, progressive supranuclear palsy (PSA), pathologic aging, or elderly controls without neurodegenerative diseases. 19 Only temporal cortex samples that were homozygous for the risk or protective haplotypes at SNP rs3173615 and rs1990622 were included in our analysis. Using the IGV, we visualized each RNA-seq Binary Alignment Map (BAM) file. ...
Article
Full-text available
INTRODUCTION Sequence variants in TMEM106B have been associated with an increased risk of developing dementia. METHODS As part of our efforts to generate a set of mouse lines in which we replaced the mouse Tmem106b gene with a human TMEM106B gene comprised of either a risk or protective haplotype, we conducted an in‐depth sequence analysis of these alleles. We also analyzed transcribed TMEM106B sequences using RNA‐seq data (AD Knowledge portal) and full genome sequences (1000 Genomes). RESULTS We identified an AluYb8 insertion in the 3' untranslated region (3'UTR) of the TMEM106B risk haplotype. We found this AluYb8 insertion in every risk haplotype analyzed, but not in either protective haplotypes or in non‐human primates. DISCUSSION We conclude that this risk haplotype arose early in human development with a single Alu‐insertion event within a unique haplotype context. This AluYb8 element may act as a functional variant in conferring an increased risk of developing dementia. Highlights We conducted an in‐depth sequence analysis of (1) a risk and (2) a protective haplotype of the human TMEM106B gene. We also analyzed transcribed TMEM106B sequences using RNA‐seq data (AD Knowledge Portal) and full genome sequences (1000 Genomes). We identified an AluYb8 insertion in the 3' untranslated region (3'UTR) of the TMEM106B risk haplotype. We found this AluYb8 insertion in every risk haplotype analyzed, but not in either protective haplotypes or in non‐human primates. This AluYb8 element may act as a functional variant in conferring an increased risk of developing dementia.
... In this study, we used the iMAT algorithm to generate personalized genome-scale metabolic models for each individual in three different large AD study cohorts; The Religious Orders Study (ROS) and Rush Memory and Aging Project (MAP) (De Jager, et al., 2018), Mayo Clinic (Allen, et al., 2016) and Mount Sinai Brain Bank (MSBB) (Wang, et al., 2016), which collectively cover a total of 643 AD individuals. For the first time in the literature, we extracted both transcriptomic and genomic information from the same RNA-seq sample to generate personalized metabolic models. ...
Preprint
Motivation: Alzheimer's disease (AD) is known to cause alterations in brain metabolism. Furthermore, genomic variants in enzyme-coding genes may exacerbate AD-linked metabolic changes. Generating condition-specific metabolic models by mapping gene expression data to genome-scale metabolic models is a routine approach to elucidate disease mechanisms from a metabolic perspective. RNAseq data provides both gene expression and genomic variation information. Integrating variants that perturb enzyme functionality from the same RNAseq data may enhance model accuracy, offering insights into genome-wide AD metabolic pathology. Results: Our study pioneers the extraction of both transcriptomic and genomic data from the same RNA-seq data to reconstruct personalized metabolic models. We mapped genes with significantly higher load of pathogenic variants in AD onto a human genome-scale metabolic network together with the gene expression data. Comparative analysis of the resulting personalized patient metabolic models with the control models showed enhanced accuracy in detecting AD-associated metabolic pathways compared to the case where only expression data was mapped on the metabolic network. Besides, several otherwise would-be missed pathways were annotated in AD by considering the effect of genomic variants.
... 4,5 Sporadic or late-onset AD (LOAD) is more common (> 95% prevalence), with symptoms arising after age 65. 5,6 Its risk factors include age, the apolipoprotein E (APOE) ε4 allele, and point mutations in triggering receptor expressed on myeloid cells 2 (TREM2). 3,[7][8][9][10] Females are at higher risk 11,12 and comprise the majority of cases, 13 and female APOE ε4 carriers are at greater risk of developing AD compared to male carriers. [14][15][16] Pathophysiological changes associated with AD begin decades before clinical symptoms. ...
Article
Full-text available
INTRODUCTION Increasing evidence suggests that metabolic impairments contribute to early Alzheimer's disease (AD) mechanisms and subsequent dementia. Signals in metabolic pathways conserved across species can facilitate translation. METHODS We investigated differences in serum and brain metabolites between the early‐onset 5XFAD and late‐onset LOAD1 (APOE4.Trem2*R47H) mouse models of AD to C57BL/6J controls at 6 months of age. RESULTS We identified sex differences for several classes of metabolites, such as glycerophospholipids, sphingolipids, and amino acids. Metabolic signatures were notably different between brain and serum in both mouse models. The 5XFAD mice exhibited stronger differences in brain metabolites, whereas LOAD1 mice showed more pronounced differences in serum. DISCUSSION Several of our findings were consistent with results in humans, showing glycerophospholipids reduction in serum of apolipoprotein E (apoE) ε4 carriers and replicating the serum metabolic imprint of the APOE ε4 genotype. Our work thus represents a significant step toward translating metabolic dysregulation from model organisms to human AD. Highlights This was a metabolomic assessment of two mouse models relevant to Alzheimer's disease. Mouse models exhibit broad sex‐specific metabolic differences, similar to human study cohorts. The early‐onset 5XFAD mouse model primarily alters brain metabolites while the late‐onset LOAD1 model primarily changes serum metabolites. Apolipoprotein E (apoE) ε4 mice recapitulate glycerophospolipid signatures of human APOE ε4 carriers in both brain and serum.
Article
Full-text available
INTRODUCTION Tropomyosin related kinase B (TrkB) and C (TrkC) receptor signaling promotes synaptic plasticity and interacts with pathways affected by amyloid beta (Aβ) toxicity. Upregulating TrkB/C signaling could reduce Alzheimer's disease (AD)‐related degenerative signaling, memory loss, and synaptic dysfunction. METHODS PTX‐BD10‐2 (BD10‐2), a small molecule TrkB/C receptor partial agonist, was orally administered to aged London/Swedish‐APP mutant mice (APPL/S) and wild‐type controls. Effects on memory and hippocampal long‐term potentiation (LTP) were assessed using electrophysiology, behavioral studies, immunoblotting, immunofluorescence staining, and RNA sequencing. RESULTS In APPL/S mice, BD10‐2 treatment improved memory and LTP deficits. This was accompanied by normalized phosphorylation of protein kinase B (Akt), calcium‐calmodulin–dependent kinase II (CaMKII), and AMPA‐type glutamate receptors containing the subunit GluA1; enhanced activity‐dependent recruitment of synaptic proteins; and increased excitatory synapse number. BD10‐2 also had potentially favorable effects on LTP‐dependent complement pathway and synaptic gene transcription. DISCUSSION BD10‐2 prevented APPL/S/Aβ‐associated memory and LTP deficits, reduced abnormalities in synapse‐related signaling and activity‐dependent transcription of synaptic genes, and bolstered transcriptional changes associated with microglial immune response. Highlights Small molecule modulation of tropomyosin related kinase B (TrkB) and C (TrkC) restores long‐term potentiation (LTP) and behavior in an Alzheimer's disease (AD) model. Modulation of TrkB and TrkC regulates synaptic activity‐dependent transcription. TrkB and TrkC receptors are candidate targets for translational therapeutics. Electrophysiology combined with transcriptomics elucidates synaptic restoration. LTP identifies neuron and microglia AD‐relevant human‐mouse co‐expression modules.
Preprint
Full-text available
Multi-omic data, i.e., genomics, epigenomics, transcriptomics, proteomics, characterize cellular complex signaling systems from multi-level and multi-view and provide a holistic view of complex cellular signaling pathways. However, it remains challenging to integrate and interpret multi-omics data. Graph neural network (GNN) AI models have been widely used to analyze graph-structure datasets and are ideal for integrative multi-omics data analysis because they can naturally integrate and represent multi-omics data as a biologically meaningful multi-level signaling graph and interpret multi-omics data by node and edge ranking analysis for signaling flow/cascade inference. However, it is non-trivial for graph-AI model developers to pre-analyze multi-omics data and convert them into graph-structure data for individual samples, which can be directly fed into graph-AI models. To resolve this challenge, we developed mosGraphGen (multi-omics signaling graph generator), a novel computational tool that generates multi-omics signaling graphs of individual samples by mapping the multi-omics data onto a biologically meaningful multi-level background signaling network. With mosGraphGen, AI model developers can directly apply and evaluate their models using these mos-graphs. We evaluated the mosGraphGen using both multi-omics datasets of cancer and Alzheimer’s disease (AD) samples. The code of mosGraphGen is open-source and publicly available via GitHub: https://github.com/Multi-OmicGraphBuilder/mosGraphGen
Article
The interplay between selenoproteins, oxidative stress, and cell death pathways holds promise in unravelling novel therapeutic targets for Alzheimer’s disease (AD) in the future. Nonetheless, further comprehensive investigations are warranted to fully comprehend the precise contributions of selenoproteins in the aetiology and potential therapeutic strategies for Alzheimer’s disease. Previous work into gene expression networks in AD has included analysis of the entire transcriptome and, as of yet, has not yielded consistent insight into pathological pathways.1 Despite the comprehensive assessment of the transcriptome enabled by current technologies, one drawback of the whole transcriptome analysis is the risk of overlooking subtle yet significant variations in metabolic pathways.2 Thus, we aimed to assess gene expression of known selenoprotein and selenium-containing pathways in two different brain regions (dorsolateral prefrontal cortex (DPC) and posterior cingulate cortex (PCC)) across the AD spectrum. We used RNA sequencing data from The Rush University’s Religious Orders Study and Memory and Aging Project (ROSMAP) cohort available in the AD Knowledge Portal (https://www.synapse.org/).3 This study included data available for a total of 889 DPC and 647 PCC samples. Four pathological phenotypes were determined based on pathology (CERAD) and clinical (CDR) status: AD ([(+) pathology, (+) clinical], prodromal disease, corresponding to donors that have not received a clinical diagnosis despite the presence of pathological alterations ([(+) pathology, (−) clinical], controls ([(−) pathology, (−) clinical] and non-AD dementia [(+) pathology, (+) clinical]. This last group was excluded from the analysis as it is assumed they may have been misdiagnosed or presented with non-AD dementia. Six selenium or AD-related pathways were assessed, accounting for 421 unique genes. Group comparisons were performed using linear mixed modelling adjusted for age, sex, APOEe4 status and batch via DESeq2 package with Benjamini-Hochberg adjustment for multiple testing. A total of 18 genes significantly differed between AD and controls in both brain areas (same direction in both brain areas; P < 0.05), including eight selenoprotein genes or genes directly associated with selenoprotein synthesis. Fifteen of them were also different (same direction) in PCC (seven selenoprotein/selenoprotein synthesis genes), and four were different in DPC (four selenoprotein/selenoprotein synthesis genes) between AD and prodromal. Only three genes significantly differed between prodromal and control samples (DPC), including the selenoprotein DIO3 and the transcription factor SP3. Our findings indicate a progressive change in gene expression across the different stages of AD. These findings shed light on critical genes involved in selenoprotein synthesis that play a role in AD pathogenesis. Restricting the analysis to a subset of pathways enabled the detection of smaller alterations between groups, which is particularly appropriate in trace element homeostasis, where small alterations may have significant downstream effects.
Article
INTRODUCTION Genome‐wide association studies have identified over 70 genetic loci associated with late‐onset Alzheimer's disease (LOAD), but few candidate polymorphisms have been functionally assessed for disease relevance and mechanism of action. METHODS Candidate genetic risk variants were informatically prioritized and individually engineered into a LOAD‐sensitized mouse model that carries the AD risk variants APOE ε4/ε4 and Trem2*R47H. The potential disease relevance of each model was assessed by comparing brain transcriptomes measured with the Nanostring Mouse AD Panel at 4 and 12 months of age with human study cohorts. RESULTS We created new models for 11 coding and loss‐of‐function risk variants. Transcriptomic effects from multiple genetic variants recapitulated a variety of human gene expression patterns observed in LOAD study cohorts. Specific models matched to emerging molecular LOAD subtypes. DISCUSSION These results provide an initial functionalization of 11 candidate risk variants and identify potential preclinical models for testing targeted therapeutics. Highlights A novel approach to validate genetic risk factors for late‐onset AD (LOAD) is presented. LOAD risk variants were knocked in to conserved mouse loci. Variant effects were assayed by transcriptional analysis. Risk variants in Abca7 , Mthfr , Plcg2, and Sorl1 loci modeled molecular signatures of clinical disease. This approach should generate more translationally relevant animal models.
Article
Full-text available
To determine the effects of single nucleotide polymorphisms (SNPs) identified in a genome-wide association study of progressive supranuclear palsy (PSP), we tested their association with brain gene expression, CpG methylation and neuropathology. In 175 autopsied PSP subjects, we performed associations between seven PSP risk variants and temporal cortex levels of 20 genes in-cis, within ±100 kb. Methylation measures were collected using reduced representation bisulfite sequencing in 43 PSP brains. To determine whether SNP/expression associations are due to epigenetic modifications, CpG methylation levels of associated genes were tested against relevant variants. Quantitative neuropathology endophenotypes were tested for SNP associations in 422 PSP subjects. Brain levels of LRRC37A4 and ARL17B were associated with rs8070723; MOBP with rs1768208 and both ARL17A and ARL17B with rs242557. Expression associations for LRRC37A4 and MOBP were available in an additional 100 PSP subjects. Meta-analysis revealed highly significant associations for PSP risk alleles of rs8070723 and rs1768208 with higher LRRC37A4 and MOBP brain levels, respectively. Methylation levels of one CpG in the 3' region of ARL17B associated with rs242557 and rs8070723. Additionally, methylation levels of an intronic ARL17A CpG associated with rs242557 and that of an intronic MOBP CpG with rs1768208. MAPT and MOBP region risk alleles also associated with higher levels of neuropathology. Strongest associations were observed for rs242557/coiled bodies and tufted astrocytes; and for rs1768208/coiled bodies and tau threads. These findings suggest that PSP variants at MAPT and MOBP loci may confer PSP risk via influencing gene expression and tau neuropathology. MOBP, LRRC37A4, ARL17A and ARL17B warrant further assessment as candidate PSP risk genes. Our findings have implications for the mechanism of action of variants at some of the top PSP risk loci.
Article
Full-text available
Objective: To investigate the top late-onset Alzheimer disease (LOAD) risk loci detected or confirmed by the International Genomics of Alzheimer's Project for association with brain gene expression levels to identify variants that influence Alzheimer disease (AD) risk through gene expression regulation. Methods: Expression levels from the cerebellum (CER) and temporal cortex (TCX) were obtained using Illumina whole-genome cDNA-mediated annealing, selection, extension, and ligation assay (WG-DASL) for ∼400 autopsied patients (∼200 with AD and ∼200 with non-AD pathologies). We tested 12 significant LOAD genome-wide association study (GWAS) index single nucleotide polymorphisms (SNPs) for cis association with levels of 34 genes within ±100 kb. We also evaluated brain levels of 14 LOAD GWAS candidate genes for association with 1,899 cis-SNPs. Significant associations were validated in a subset of TCX samples using next-generation RNA sequencing (RNAseq). Results: We identified strong associations of brain CR1, HLA-DRB1, and PILRB levels with LOAD GWAS index SNPs. We also detected other strong cis-SNPs for LOAD candidate genes MEF2C, ZCWPW1, and SLC24A4. MEF2C and SLC24A4, but not ZCWPW1 cis-SNPs, also associate with LOAD risk, independent of the index SNPs. The TCX expression associations could be validated with RNAseq for CR1, HLA-DRB1, ZCWPW1, and SLC24A4. Conclusions: Our results suggest that some LOAD GWAS variants mark brain regulatory loci, nominate genes under regulation by LOAD risk variants, and annotate these variants for their brain regulatory effects.
Article
Full-text available
The process of converting raw RNA sequencing (RNA-seq) data to interpretable results can be circuitous and time-consuming, requiring multiple steps. We present an RNA-seq mapping algorithm that streamlines this process. Our algorithm utilizes a hash table approach to leverage the availability and the power of high memory machines. SNAPR, which can be run on a single library or thousands of libraries, can take compressed or uncompressed FASTQ and BAM files, and output a sorted BAM file, individual read counts, and gene fusions, and can identify exogenous RNA species in a single step. SNAPR also does native Phred score filtering of reads. SNAPR is also well suited for future sequencing platforms that generate longer reads. We show how we can analyze data from hundreds of TCGA samples in a matter of hours while identifying gene fusions and viral events at the same time. With the reference genome and transcriptome undergoing periodic updates and the need for uniform parameters when integrating multiple data sets, there is great need for a streamlined process for RNA-seq analysis. We demonstrate how SNAPR does this efficiently and accurately.
Article
Full-text available
Genetic data from the Alzheimer's Disease Neuroimaging Initiative (ADNI) have been crucial in advancing the understanding of Alzheimer's disease (AD) pathophysiology. Here, we provide an update on sample collection, scientific progress and opportunities, conceptual issues, and future plans. Lymphoblastoid cell lines and DNA and RNA samples from blood have been collected and banked, and data and biosamples have been widely disseminated. To date, APOE genotyping, genome-wide association study (GWAS), and whole exome and whole genome sequencing data have been obtained and disseminated. ADNI genetic data have been downloaded thousands of times, and >300 publications have resulted, including reports of large-scale GWAS by consortia to which ADNI contributed. Many of the first applications of quantitative endophenotype association studies used ADNI data, including some of the earliest GWAS and pathway-based studies of biospecimen and imaging biomarkers, as well as memory and other clinical/cognitive variables. Other contributions include some of the first whole exome and whole genome sequencing data sets and reports in healthy controls, mild cognitive impairment, and AD. Numerous genetic susceptibility and protective markers for AD and disease biomarkers have been identified and replicated using ADNI data and have heavily implicated immune, mitochondrial, cell cycle/fate, and other biological processes. Early sequencing studies suggest that rare and structural variants are likely to account for significant additional phenotypic variation. Longitudinal analyses of transcriptomic, proteomic, metabolomic, and epigenomic changes will also further elucidate dynamic processes underlying preclinical and prodromal stages of disease. Integration of this unique collection of multiomics data within a systems biology framework will help to separate truly informative markers of early disease mechanisms and potential novel therapeutic targets from the vast background of less relevant biological processes. Fortunately, a broad swath of the scientific community has accepted this grand challenge. Copyright © 2015 The Authors. Published by Elsevier Inc. All rights reserved.
Article
Eighty-three brains obtained at autopsy from nondemented and demented individuals were examined for extracellular amyloid deposits and intraneuronal neurofibrillary changes. The distribution pattern and packing density of amyloid deposits turned out to be of limited significance for differentiation of neuropathological stages. Neurofibrillary changes occurred in the form of neuritic plaques, neurofibrillary tangles and neuropil threads. The distribution of neuritic plaques varied widely not only within architectonic units but also from one individual to another. Neurofibrillary tangles and neuropil threads, in contrast, exhibited a characteristic distribution pattern permitting the differentiation of six stages. The first two stages were characterized by an either mild or severe alteration of the transentorhinal layer Pre-alpha (transentorhinal stages I-II). The two forms of limbic stages (stages III-IV) were marked by a conspicuous affection of layer Pre-alpha in both transentorhinal region and proper entorhinal cortex. In addition, there was mild involvement of the first Ammon's horn sector. The hallmark of the two isocortical stages (stages V-VI) was the destruction of virtually all isocortical association areas. The investigation showed that recognition of the six stages required qualitative evaluation of only a few key preparations.
Article
Germ-line genetic control of gene expression occurs via expression quantitative trait loci (eQTLs). We present a large, exon-specific eQTL data set covering ten human brain regions. We found that cis-eQTL signals (within 1 Mb of their target gene) were numerous, and many acted heterogeneously among regions and exons. Co-regulation analysis of shared eQTL signals produced well-defined modules of region-specific co-regulated genes, in contrast to standard coexpression analysis of the same samples. We report cis-eQTL signals for 23.1% of catalogued genome-wide association study hits for adult-onset neurological disorders. The data set is publicly available via public data repositories and via http://www.braineac.org/. Our study increases our understanding of the regulation of gene expression in the human brain and will be of value to others pursuing functional follow-up of disease-associated variants.