Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

RESEARCH ARTICLE | NEUROSCIENCE OPEN ACCESS

Neuron-specific transcriptomic signatures indicate


neuroinflammation and altered neuronal activity in ASD
temporal cortex
Pan Zhanga,b,c, Alicja Omanskad,e , Bradley P. Andere,f, Michael J. Gandalc,g,h,i,1,2 , Boryana Stamovae,f,1,2 , and Cynthia M. Schumannd,e,1,2

Edited by Huda Akil, University of Michigan-Ann Arbor, Ann Arbor, MI; received July 12, 2022; accepted December 28, 2022

Autism spectrum disorder (ASD) is a highly heterogeneous disorder, yet transcriptomic


profiling of bulk brain tissue has identified substantial convergence among dysregu- Significance
lated genes and pathways in ASD. However, this approach lacks cell-specific resolution.
We performed comprehensive transcriptomic analyses on bulk tissue and laser-capture We present a comprehensive
microdissected (LCM) neurons from 59 postmortem human brains (27 ASD and 32 assessment of neuronal cell-type-
controls) in the superior temporal gyrus (STG) of individuals ranging from 2 to 73 years specific gene expression and
of age. In bulk tissue, synaptic signaling, heat shock protein-related pathways, and RNA alternative splicing changes in
splicing were significantly altered in ASD. There was age-dependent dysregulation of ASD cortex, directly comparing
genes involved in gamma aminobutyric acid (GABA) (GAD1 and GAD2) and glutamate RNA-seq results from bulk tissue
(SLC38A1) signaling pathways. In LCM neurons, AP-1-mediated neuroinflammation with isolated neurons. We
and insulin/IGF-1 signaling pathways were upregulated in ASD, while mitochondrial
observe strong signatures of cell
function, ribosome, and spliceosome components were downregulated. GABA synthesiz-
stress and neural-immune/
ing enzymes GAD1 and GAD2 were both downregulated in ASD neurons. Mechanistic
modeling suggested a direct link between inflammation and ASD in neurons, and pri- inflammatory pathway activation
oritized inflammation-associated genes for future study. Alterations in small nucleolar present within ASD neurons—a
RNAs (snoRNAs) associated with splicing events suggested interplay between snoRNA signal that is typically attributed
dysregulation and splicing disruption in neurons of individuals with ASD. Our findings to astrocyte/microglial
supported the fundamental hypothesis of altered neuronal communication in ASD, populations. Our findings also
demonstrated that inflammation was elevated at least in part in ASD neurons, and provide further evidence for the
may reveal windows of opportunity for biotherapeutics to target the trajectory of gene hypothesized imbalance of
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

expression and clinical manifestation of ASD throughout the human lifespan. excitatory to inhibitory neuronal
activity in the brains of
ASD | transcriptome | neuron-specific
individuals with ASD. Moreover,
Autism spectrum disorder (ASD) defines a heterogeneous set of complex neurodevelop- we find that the transcriptomic
mental disorders affecting 1 in 54 children in the USA according to current estimation architecture of ASD interacts
(1, 2) and confers lifelong challenges. ASD is characterized by difficulties with social substantially with age, thus
communication as well as a repetitive, restricted repertoire of behaviors and interests (3). revealing windows of opportunity
Population, family, and twin studies all indicate a strong genetic component contributing for treatments that target specific
to risk for ASDs (4, 5), with heritability estimates of ~70% (6). However, the genetic molecular pathology.
causes and pathophysiology of ASD are varied and often complex.
Despite this heterogeneity, transcriptomic analyses of postmortem human brain have
elucidated substantial convergent molecular-level pathology associated with idiopathic
and syndromic forms of ASD (7–13). Multiple studies have profiled the transcriptomes
of postmortem brain regions from individuals diagnosed with ASD (7, 8, 11, 13),
including the temporal cortex implicated due to its critical importance in speech and
language function (7, 8, 11, 13). The most consistent findings include disruption of
neuronal/synaptic activity and activation of innate immunity/glial markers (7, 8, 11). Author contributions: B.P.A., B.S., and C.M.S. designed
Alternative splicing and non-coding RNAs have also been shown to be dysregulated in research; P.Z., A.O., M.J.G., B.P.A., B.S., and C.M.S.
performed research; P.Z., A.O., M.J.G., and B.S. analyzed
ASD brains (8). data; M.J.G., B.S., and C.M.S. supervised the study; and
Most previous transcriptomic studies, however, profiled homogenate brain tissue and P.Z., B.P.A., M.J.G, B.S., and C.M.S. wrote the paper.

have therefore been unable to pinpoint the underlying specific cell types in which gene The authors declare no competing interest.

expression is altered. Single-nucleus RNA-sequencing (sn-RNAseq) datasets have started This article is a PNAS Direct Submission.

to be generate in postmortem ASD cortex (12, 13), identifying substantial changes in Copyright © 2023 the Author(s). Published by PNAS.
This open access article is distributed under Creative
upper-layer excitatory neurons and microglia, consistent with observations from bulk Commons Attribution License 4.0 (CC BY).
tissue. As such sn-RNAseq datasets currently primarily profile the 3′ end of highly 1
To whom correspondence may be addressed. Email:
expressed genes within each cell, these data characterize neither lowly expressed coding michael.gandal@pennmedicine.upenn.edu, bsstamova
@ucdavis.edu, or cschumann@ucdavis.edu.
and noncoding genes, nor splicing alterations that may contribute to altered neuronal 2
M.J.G., B.S., and C.M.S. contributed equally to this work.
function in ASD.
This article contains supporting information online at
Here, we performed the systematic study using transcriptomic profiling to directly https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.​
compare both bulk cortical tissue and laser capture microdissected (LCM) neurons from 2206758120/-/DCSupplemental.
anatomically well-defined superior temporal gyrus (STG) samples from 59 subjects (27 Published March 2, 2023.

PNAS 2023 Vol. 120 No. 10 e2206758120 https://doi.org/10.1073/pnas.2206758120 1 of 11


with ASD and 32 controls) ranging from 2 to 73 years of age changes in neurons remain unknown. This study aimed to identify
(Fig. 1 and Dataset S12). The STG modulates language processing neuron-specific transcriptomic changes in ASD brain by identi-
and social perception, thereby playing a critical role in integrating fying differentially expressed genes, differential splicing (DS)
a breadth of information to provide meaning to the surrounding events, age-related gene expression changes across the lifespan,
world (14). Structural and functional imaging studies have long and co-expression networks to reveal gene modules altered in
implicated STG in ASD (14, 15); however, molecular-level ASD. We also applied mechanistic modeling approaches to

A
Ground tissue for
bulk analysis
3A
Total RNA extraction
Superior temporal gyrus 4 and QC
sampled from Tissue flash frozen
32 Controls in liquid N2

1 27 ASD
2 Embedded in OCT
for LCM
3B

Prepare cDNA libraries,


5 QC, and pool

Sectioned on cryostat Placed on PEN membrane Neurons (100/sample)


to 12 µm slides and stained captured directly into tubes
by laser capture microscope DNA sequencing
with HistoGene
(50M paired-end
3C 3D 3E 6 150 bp reads)
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

7 Data analysis pipeline

B Diagnosis Age
1.00
Sex
30 F
60
0.75
20
count

Age

40 0.50
10
20 0.25 M

0 0 0.00
Control Autism Control Autism Control Autism

PMI RIN
8

30
6
Hour

RIN

20
4
10

2
Control Autism Control Autism

Fig. 1. Overview of Experiment Design, Data Analysis Pipeline (A), and Study Cohort (B).

2 of 11 https://doi.org/10.1073/pnas.2206758120 pnas.org
pinpoint pathways and genes directly linked to ASD in neurons. comprehensively characterized differential gene expression (DGE)
(Fig. 1). and local splicing alterations in ASD. After adjusting for known
covariates and correcting for multiple comparisons, we found 194
Results differentially expressed genes between individuals with ASD and
controls (FDR < 0.05). Of these, 143 were upregulated and 51
Global Gene Expression Changes in ASD STG. RNA sequencing were downregulated (Fig. 2A and Dataset S1), with a median
was performed on bulk tissue STG of 59 human brains, 27 absolute fold change of 1.45 (range 1.11 to 4.04, Fig. 2A). We
from individuals with ASD and 32 from neurotypical controls, observed significant concordance between our DGE results and
ranging from 2 to 73 years of age. Following quality control, we previous data of the same region from different samples (13)

A B
Number of DE genes ( FDR < 0.05)

Disease effect
25 Block-M32

*
*

*
-0.1 0 0.1
20 Block-M14 beta

*
*
*
Block-M3 GWAS enrichment

*
*

*
*
15 No. of genes upregulated: 143
Block-M19 0 0.5 1

*
*
*
-log10FDR
10 Block-M10

*
*
*
*

*
*
*
*
*
*
*
*
*
*
Cell type enrichment
5 Block-M31

*
*
*

*
*

*
0 1 2 3
Block-M1 -log10FDR

*
*

*
*
*
*
*

*
*
0

OPC

Microglia
InNeu

LGD+Missense
Presynaptic
Oligo

Endo
Astro

LGD DNMs
ExcNeu

Postsynaptic

pLI>0.99
Module-list overlap,

Syndromic ASD
Vulnerable
Constrained
ASD GWAS
Per
ASD expression

FMRP Target
CHD8 targets
5 permutation test

No. of genes downregulated: 51


10 0 5 10 15
-log10FDR
1.0 1.5 2.0 2.5 3.0 3.5 4.0
RNA-seq Fold Change

C
6 RGS8
MGA

SPEN
CRYAB

HSPB1
D chaperone-mediated protein folding
DNA packaging complex q value
FXR1 PLOD3
GAN
BAG3
nucleosome 0.018
NEMF PIGT SLC16A1SNX31HSPA1A response to unfolded protein 0.015
CLUL1
PCBP1-AS1 TRA2A HIBCH
BCOR
NKAIN1
LINC01535 CPEB1-AS1 extrinsic apoptotic signaling pathway
C20orf27 ZBTB22 CRKL
DNAJB4 SPR
regulation of intracellular pH 0.012
NR2F2-AS1 CDKN2D XPNPEP3 SAV1 HIST2H2BD HSPA1B CP
NR2F2
CDC7 KIAA1614
RBM48 JAM3 ST6GALNAC4P1 unfolded protein binding 0.009
4 MYO1E
myelination
-log10 p value

KCNH3 PTGES3 DNAJB1


regulation of transcription in response to stress
cellular response to unfolded protein
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

HSPB8
KCNIP1
synaptic transmission, glutamatergic
neurotransmitter receptor complex overlap
Golgi lumen 20
2 collagen trimer 30
multicellular organismal response to stress 40
neurotransmitter receptor activity 50
extracellular matrix structural constituent 60
G protein-coupled peptide receptor activity 70
excitatory postsynaptic potential
0 ammonium ion binding
-2 -1 0 1 2 -2 -1 0 1 2
log2 FoldChange NES

E 14
F EPB41L3
12
10 Significant DS: 308

13
GAD1 expression

GAD2 expression

SEPT8 APBA2
in block tissue

in block tissue

MED4 ANKRD20A3 COG1


8
observed

CD47 MARCH7 ROBO3


9 12 Diagnosis CLASP2 CNTNAP1
Control NALCN IQSEC1 CLASP2
PIK3AP1 EWSR1
Autism
MACF1
11
8 4

10
Diagnosis * Age interaction FDR = 0.026
7 Diagnosis * Age interaction FDR = 0.047
9 0
0 20 40 60 0 20 40 60 0 1 2 3 4 5
Age Age expected

Fig. 2. Transcriptomic Difference Between ASD Cases and Controls in Bulk Tissue STG. (A) Distribution of fold-change of differential expression for 194
differentially expressed genes. Case/control fold-changes for upregulated genes are plotted in gold (N = 143, positive values) and control/case fold-changes for
downregulated genes in blue (N = 51, negative values). (B) Co-expressed gene modules that were significantly disrupted in ASD. Modules were hierarchically
clustered by module eigengene. Module-diagnosis associations were shown on the right of each module (*FDR < 0.05). Additional enrichment analyses were also
shown for each module, including: Enrichment for ASD GWAS common variants (33)(*FDR < 0.1); Enrichment for major CNS cell types (16)(*FDR < 0.05); Enrichment
against literature-curated gene lists (*FDR < 0.05, see SI Appendix, SI Methods for details of the gene lists). Per, pericytes; OPC, oligodendrocyte progenitor cells;
InNeu, inhibitory neuron; ExcNeu, excitatory neuron; Oligo, oligodendrocytes; Endo, endothelial cells; Astro, astrocytes. (C) Volcano plot showing significantly
up- (gold) and down-regulated (blue) genes (FDR < 0.05). Genes discussed in the main text are colored red. (D) Functional enrichment of differentially expressed
genes in ASD cases compared to controls. Top 10 significantly enriched up- and down-regulated categories were shown. The color of each dot reflects FDR-
corrected q-value, and the size of each dot reflects the number of overlapped genes between our gene list and the corresponding GO category. NES, normalized
enrichment score. (E) Age trajectory of GAD1 (Left) and GAD2 (Right) gene expression, stratified by ASD diagnosis. The expression levels were on the log2CPM
(counts per million) scale. (F) Quantile-quantile plot of observed P-values vs. expected P-values for differentially spliced intron clusters. Significant DS events
(FDR < 0.05) were colored red, and overlapping gene names were labeled for top clusters.

PNAS 2023 Vol. 120 No. 10 e2206758120 https://doi.org/10.1073/pnas.2206758120 3 of 11


(SI Appendix, Fig. S1, Spearman ρ = 0.37 for t statistics among excitatory and inhibitory neurons (Fig. 2B), suggesting a broad
all genes, P value < 10−16). disruption of neuronal and synaptic processes in ASD STG.
Functional and pathway enrichment analyses indicated an Genes in the upregulated module Block-M1 and one downreg-
over-representation of heat-shock proteins (HSPs) and HSP- ulated module (Block-M10) were enriched in high-confidence
related chaperones, which were upregulated in ASD subjects. This ASD risk loci (29, 30), mutationally constrained (31), and highly
included HSP70 family members HSPA1A and HSPA1B; HSP40 intolerant to mutations (pLI > 0.99) genes (32), as well as regu-
family members DNAJB1 and DNAJB4; small HSP20 family latory target genes of CHD8, which has clear links to at least a
members HSPB1 and HSPB8; and HSP-binding chaperons BAG3 subset of ASD cases (33) (Fig. 2B, see SI Appendix, SI Methods for
and PTGES3 (Fig. 2 C and D). HSPs are involved in stress the details of all curated, hypothesis-driven gene sets. Dataset
response, immune activation, and inflammation(17, 18), all of S14). Many hub genes for module Block-M10 encoded synaptic
which were upregulated in ASD postmortem brain (7). proteins (SI Appendix, Fig. S2 and Dataset S3). This module was
Downregulated genes were mainly enriched in pathways related also enriched for ASD common risk alleles from ASD GWAS data
to synaptic function (Fig. 2D), consistent with previous findings (34). Together this suggested a causal role of synaptic dysfunction
(7). Notably, two important voltage-gated potassium channel- in ASD etiology.
related genes KCNH3 and KCNIP1 were among the most down-
regulated (Fig. 2C), which may relate to disrupted neuronal excit- Neuron-Specific Gene Expression and Splicing Alterations in ASD
ability hypothesized in ASD (19, 20). STG. To provide cell-type specificity for the observed transcriptomic
As age-dependent expression alterations have been reported in changes, we performed laser capture microdissection to capture
ASD brain (21), we employed an analytical model accounting for neurons using STG sections taken from the same subjects profiled
age and the interaction between age and diagnosis. Fourteen genes using bulk RNA-seq (SI Appendix, Figs. S6 and S7). We then
showed age-dependent DGE between ASD and control (Dataset interrogated ASD-associated gene expression and splicing alterations
S2). Gene set enrichment analysis (GSEA) indicated that genes using the same bioinformatic pipelines as above. Across 13,458
with significant diagnosis-by-age interaction were enriched in neuron-expressed genes, 83 were significantly differentially expressed
immune/inflammation pathways and synaptic-related pathways. between ASD subjects and controls at FDR <0.05, of which 52 were
(SI Appendix, Fig. S9). upregulated and 31 downregulated (Fig. 3A and Dataset S6). Median
Interestingly, genes involved in gamma aminobutyric acid (GABA) absolute fold change in expression between subjects with ASD and
synthesis (GAD1 and GAD2) (22) were downregulated in ASD only controls was 2.48 (range 1.29 to 9.72; Fig. 3A). The concordance of
during late adulthood (Fig. 2E). This may indicate an age-dependent neuronal DGE with bulk tissue DGE was low (Spearman ρ = 0.18
dysregulation of GABA signaling in ASD neurons, or a decrease in for t statistics across all genes, SI Appendix, Fig. S3), suggesting our
the proportion of GABAergic neurons in ASD brains (23). analysis captured ASD signatures unique to neurons.
Given mounting evidence suggesting potential mechanistic Upregulated genes in ASD neurons were highly enriched in
overlap between neurodevelopmental and neurodegenerative dis- pathways related to growth and differentiation (Fig. 3E).
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

orders, particularly with respect to genes involved in synapse and Specifically, the AP-1 transcription factor complex components
brain connectivity (24), we evaluated the overlap between our (35) FOS, JUN and JUNB were all upregulated in ASD neurons,
ASD findings with published results in Alzheimer’s disease (AD). among other growth/differentiation regulators such as SOX9,
Using results from Mount Sinai/JJ Peters VA Medical Center Brain S1PR1, and PPP1R16B (Fig. 3D). The AP-1 transcription factor
Bank (MSBB) cohort (25), we demonstrated that the changes we complex is known to regulate many downstream biological pro-
detected in ASD were significantly overlapped with transcriptomic cesses (35). Indeed, we found upregulated B cell signaling adap-
alterations observed in AD from the same brain region (Odds tor gene BCL10 and NFκB inhibitor delta gene NFKBID in
Ratio = 1.6, P = 1.9 × 10−8, SI Appendix, Fig. S8). ASD neurons (Fig. 3D). Up-regulation of both NFKBID and
DS events in the bulk tissue transcriptome were evaluated using AP-1 point to dysregulated inflammation in ASD neurons. In
LeafCutter (26). Among 35,505 intron clusters identified by addition, the inward-rectifier potassium ion channel gene
LeafCutter, 308 clusters (297 unique genes) showed significant KCNJ2 was also upregulated in ASD neurons (Fig. 3D).
DS between ASD patients and controls (FDR < 0.05). The 297 Interestingly, both KCNJ2 and AP-1 subunit FOS were involved
genes did not show significant functional enrichment (Fig. 2F and in regulating excitability and plasticity at the cholinergic synapse
Dataset S5). (36–38).
To place subtle changes across the ASD STG transcriptome Downregulated genes in ASD neurons are primarily enriched in
into a systems-level context, we performed weighted gene corre- mitochondrial function and oxidoreductase activity (Fig. 3E).
lation network analysis (WGCNA) to build gene co-expression Specifically, comparing to bulk tissue STG, more subunits of the
networks (27), identifying 31 modules of co-expressed genes NADH:ubiquinone oxidoreductase (complex I) were downregulated
(Methods and Dataset S3). Seven modules significantly associated in neurons, and their effect sizes were larger (SI Appendix, Fig. S4).
with ASD diagnosis, two of which were strongly enriched for Our results provide evidence that compared to bulk tissue, mito-
ASD-associated genetic risk factors (Modules Block-M1 and chondrial dysfunction is much more profound in STG neurons.
Block-M10, Fig. 2B). While LCM captured both excitatory and inhibitory neurons,
Module Block-M1 was upregulated in ASD STG and its gene we note that GAD1 and GAD2 genes are among the most down-
members were enriched in RNA splicing and mRNA metabolic regulated in ASD neurons (Fig. 3D). The coordinated down-reg-
pathways (Dataset S4). Notably, significantly upregulated HSPs ulation of both GABA synthesizing enzymes suggest that the level
were also members of the Block-M1 module (Dataset S3). HSPs of GABA neurotransmitter may be decreased in ASD neurons,
contribute to RNA splicing during stress (28). Downregulated providing support to the excitation to inhibition (E/I) imbalance
modules in ASD were mostly enriched for synaptic functions hypothesis of ASD (19, 20, 39).
(Block-M3, Block-M10, Block-M14, Block-M19, Block-M31; By testing the interaction between diagnosis and age, three
Dataset S4). Cell-type enrichment analysis also indicated these genes (HTRA2, aka OMI; ZNF765; and PCDHB18P) showed
downregulated modules were enriched in marker genes for both age-dependent differential expression in ASD neurons (Dataset S7;

4 of 11 https://doi.org/10.1073/pnas.2206758120 pnas.org
A B C
7.5
Number of DE genes ( FDR < 0.05)
10 FRYL
L
Diagnosis * Age interaction FDR = 0.009 12 Significant DS: 1292
MED23
No. of genes upregulated: 52 ANK2 CCNDBP1
CAMK1D PLEKHG3 QRSL1

HTRA2 expression
5 5.0 CFAP69 BAG4 CNOT1
ZNF384 USP53
8 NUDT6 CEP164

observed
in neuron
Diagnosis VWA3A OCEL1L1
1
RBBP8
0 2.5 Control
Autism
neuron
4 block tissue

5 0.0
No. of genes downregulated: 31

-2.5 0
1 2 3 4 5 6 7 8 9 10 20 40 60 0 1 2 3 4 5
RNA-seq Fold Change Age expected

D 8
E
growth factor binding
ELL
endothelium development
LAMB2
kidney morphogenesis
KCNC2 DAB2
CSRNP1
endothelial cell differentiation
AL365209.1 cytokine binding q value
6 TOB2
RNU1-67P UMAD1 INPP5A TACO1
AC004890.2
ureteric bud development 0.030
GAD2 FAM216A SPINDOC
KANK2
nuclear receptor transcription coactivator activity 0.025
RGS5 RPL32 SLC45A4
branching morphogenesis of an epithelial tube
-log10 p value

JAM2 BCL10
GAD1 PDHX TMCC2 JUNTPRA1 ESRRA 0.020
RFC3
NDUFA4 ROBO2
PEAK1
PPP1R16B RASD1 JUNB MT-TT response to glucocorticoid
4 NPIPB11 RPS21 FDFT1 ATP5F1B
SLC6A8
KCNJ2 AC069120.1
FAM107A TLE3 C3orf33 regulation of cytokine-mediated signaling pathway 0.015

SOX9
FOS structural constituent of ribosome
S1PR1 NFKBID
protein targeting to ER
oxidoreductase activity, acting on NAD(P)H, quinone overlap
mitochondrial respiratory chain complex I assembly 20
2 Sm-like protein family complex 40
spliceosomal snRNP complex 60
respiratory chain
inner mitochondrial membrane protein complex
respiratory chain complex
0 oxidoreductase complex
-2 0 2 -2 -1 0 1 2
log2 FoldChange NES

Fig. 3. DGE and DS Between ASD Cases and Controls in LCM Neurons from STG. (A) Distribution of fold-change of differential expression for 83 differentially
expressed genes. Case/control fold-changes for upregulated genes are plotted in gold (N = 52, positive values) and control/case fold-changes for downregulated
genes in blue (N = 31, negative values). (B) Age trajectory of HTRA2 gene expression, stratified by ASD diagnosis. The expression levels are on the log2CPM (counts
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

per million) scale. (C) Quantile–quantile plot of observed P-values vs. expected P-values for differentially spliced intron clusters. Significant DS events (FDR < 0.05)
were colored red, and overlapping gene names were labeled for top clusters. DS results from bulk tissue were also plotted for comparison. (D) Volcano plot showing
up- (gold) and down-regulated (blue) genes. Genes discussed in the main text are colored red. (E) Functional enrichment of differentially expressed genes in ASD
cases compared to controls. Top 10 significantly enriched up- and downregulated categories are shown. The color of each dot reflects FDR-corrected q-value,
and the size of each dot reflects the number of overlapped genes between our gene list and the corresponding GO category. NES, normalized enrichment score.

no pathway enrichment was detected for neuronal genes with significantly upregulated in ASD neurons, while one module was
significant diagnosis-by-age interaction). For example, age trajec- downregulated (Fig. 4A).
tories of serine peptidase HTRA2 were opposite in ASD brains The upregulated neu-M5 co-expression module was highly
compared to controls (Fig. 3B). In healthy brains, the expression represented by the DGE analysis signal. Upregulated genes
of HTRA2 was much higher before age 30 and decreases with age, JUN, JUNB, and NFKBID were all hub genes of neu-M5 mod-
while its expression levels begin lower and increase with age in ule. Neu-M5 module also captured additional AP-1 subunits
ASD STG neurons. and interactors, such as FOSL2 (40) and IRF3 (41, 42). Neu-
To further evaluate potential overlap between ASD and neuro- M5 module was enriched in immune response pathways, pro-
degenerative processes in neurons, we compared our neuronal viding further evidence that AP-1-mediated neuroinflammation
findings with results from MSBB. Changes in ASD neuronal was elevated in ASD neurons (Fig. 4C and Dataset S10). Hubs
transcriptome were also significantly overlapped with transcrip- of Neu-M5 also contained multiple ion channel-related genes,
tomic alterations observed in AD from the same brain region (odds such as sodium ion channel gene SCN1B, potassium channel
ratio = 1.4, P = 0.01, SI Appendix, Fig. S8). genes KCNJ2 and KCNJ10, and solute carrier gene SLC40A1
We also quantified local splicing events in the neuronal tran- (Fig. 4C and Dataset S9). Coordinated upregulation of various
scriptome. After adjusting for multiple testing, LeafCutter iden- ion channels suggested that membrane transport was activated
tified 1,292 significant differential spliced intron clusters (1,177 in ASD neurons, consistent with heightened excitability. Neu-
unique genes) out of 17,250 total intron clusters at FDR < 0.05 M5 was significantly overlapped with module M16 from
(Fig. 3C and Dataset S8). No functional enrichment was Voineagu et al. (7). M16 was also enriched in immune/inflam-
observed for the 1,177 genes. We observed more disruptions in matory response and was up-regulated in ASD (7). Our data
local splicing events in ASD neurons than in bulk tissue (308 refined our understanding of the neuroinflammatory changes
DS out of 35,505 events in bulk tissue, 1,292 DS out of 17,250 in ASD to include a neuronal component. Additionally, down-
events in neurons; P < 2 × 10−16 test of proportions). regulated neu-M17 module was enriched in mitochondrial
function and contained the most differentially expressed mito-
Neuron-Specific Networks Pinpoint Subtle Changes in the chondrial genes, such as ATP synthase subunits ATP5F1B and
Neuronal Transcriptome in ASD. Co-expression network analysis ATP5PF (Fig. 4F and Dataset S9).
on neuronal data identified 18 modules, each containing between Neuronal co-expression networks further captured signals that
101 and 998 co-expressed genes (Dataset S9). Four modules were were not detected by DGE analysis. Neu-M6 module was

PNAS 2023 Vol. 120 No. 10 e2206758120 https://doi.org/10.1073/pnas.2206758120 5 of 11


A B Excitatory neurons Inhibitory neurons
*

Neu-M13
Neu-M6
Neu-M7
Neu-M4
Neu-M5
Neu-M3

Neu-M9

Neu-M2
Neu-M8
Neu-M10

Neu-M12

Neu-M14
Neu-M15
Neu-M11
Neu-M16

Neu-M17
Neu-M18
Neu-M1
Disease effect 6 * Neuron Modules
* * ** *

Z-score
ASD expression -0.1 0 0.1
Neu-M5
beta
ASD_GWAS
* 4 * Neu-M11
Postsynaptic
* * ** * GWAS enrichment
* Neu-M17
Presynaptic
* ** 0 Neu-M16
* * **
0.5 1
LGD+Missense
-log10FDR 2
LGD DNMs
* * * * Neu-M6
pLI>0.99
* * ** * ** Module-list overlap,
Vulnerable
* * * * permutation test 0
Constrained
* * ** *
FMRP Target
* * * ** *

Ex5b
Ex6b

In1b

In4b
In6b
Ex3e

Ex6a

In1a
In1c
In4a
In6a
0 5 10 15
* * * **

Ex1
Ex2
Ex4

Ex8
Ex9

In3

In7
In8
CHD8_targets -log10FDR
Syndromic ASD

C Neu-M5
D Neu-M6
E Neu-M16
signal release from synapse
sequence-specific DNA binding regulation of metal ion transport
neurotransmitter secretion
positive regulation of myeloid leukocyte regulation of cell-matrix adhesion
cytokine production in immune response calcium ion regulated exocytosis
positive regulation of intracellular signal transduction regulation of carbohydrate biosynthetic
microtubule
positive regulation of immune effector process regulation of calcium ion transport
vesicle-mediated transport in synapse
nucleobase-containing small molecule interconversion positive regulation of smooth muscle cell migration
synaptic vesicle cycle
negative regulation of pri-miRNA transcription
transmembrane transporter activity regulation of axon extension involved in axon guidance
channel activity hexose transmembrane transport
0 1 2
0.0 0.5 1.0 1.5 0 1 2 -log10FDR
-log10FDR -log10FDR
AC104024.2 ERGIC1 ARHGAP39
DOCK1 AHDC1 TNK2
KCNH8 GLRX CAMK1D
RASSF8-AS1 RAI1 IQSEC3
POLD2 GPRC5B SLC6A8
DIP2A
CTPS2 SENP3 ENDOD1 AMMECR1LP1
KCNJ10 SORBS1 LAS1L WIPF3 ZNF83 MLLT6
LIX1 SCAF1
POU2F2 HMGA1 UNC5A MEGF8 PSD
NFKBID TACO1 IRF3 AL391684.1 PTK2B
CSRP1 SCN1B DHX30 PIP5K1C
EPB41L1 SIPA1L3 LRRC4B
PCDHGB4 NME6 SND1 NSD1 EPN1
CPLX2 TOM1L2
SLC7A11 ABLIM2
UBE2L6 IGF1R NAPRT LLGL1
F8 LINC-PINT HCFC1
FGF1 TNR SAMD4B PPP1R16B POM121
LRP4 SLC40A1 S1PR1 PEAK1 CIT TTC28 GNA11
BHLHE40 HECTD3 CBX6
SLC39A14 KSR2 ANXA11 KRBA1
SOX2 MT-TT AJAP1 PCDHGB
NHSL2 KCNAB2 INPP5E
USP3 FGFR2 CAMK2A
ATP1A2 IGFBP5 PTOV1 PRRT2 CLIP2
LINC00476 ELL BCL10 PRDM11 TAMM41 SYNGR1
ELK1
ZBTB7C INPP5A ARID1A WFS1 MAP1S PCSK1N
JUN KCNJ2 AC097376.2 NCLN
ALDH6A1 UBE2H SNPH USP20
TLE3 FBXL20 FAM107ASPECC1 RNF130 SIRT7
AC138393.3 PROX1 AC016831.4 CAMTA2
TOB2 FOSL2 REPIN1 IGLON5
AC004890.2BAHD1AL080317.3 RANGAP1 SCAMP5
ADGRV1 PLOD1 DDN CYHR1
SGMS2 IRS2
IBA57 BTG2 CADM3 PACSIN1 PNKD
POU3F3
LINC02210
MCU
KCTD15

F mitochondrial membrane
G Neu-M18 Neu-M7 Neu-M9 H BCL10
AC004890.2

Neu-M17 Neu-M14
MT-TT
TLE3
proteasome core complex
ZBTB7C
oxidative phosphorylation TACO1
Neu-M12 Neu-M17 Neu-M2 Neu-M3 IBA57
mitochondrial translation USP3

mitochondrial inner membrane KCNJ2


GALNT2 S1PR1
farnesyl diphosphate metabolic process Neu-M13 Neu-M6 Neu-M8 Neu-M4 BAHD1
TOB2
antigen processing and presentation of ELL
endogenous peptide antigen via MHC class I JUN
SDC2
0.0 0.5 1.0 1.5
TLNRD1
-log10FDR Neu-M1
Neu-M5 Neu-M11 Neu-M15 Neu-M16 Neu-M10 INPP5A JUNB
SFT2D1
MARK1 LINC00476 GTF2IRD2B
COX7C
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

ECHS1
BUB3
RPLP2
ASD
PDZD11 ZWILCH

I
ANAPC7
MLLT11 CCT7
HSBP1 HELLPAR
RBM11 FDPS
Upstream regulator

PAPOLG
RFC3 DDX20
PSMA1 dopamine
KCNAB1
PSMB3 IFIT1
(predicted)

CDH7 TMEM70
NTN4
MRPS9 KLHL5

NAALAD2 EPHA7
TFG ATP5F1B
Activated
C6orf120 COA7
ATG5 CBX1
beta.estradiol DRD2 CREB1 Inhibited
GATC Pka ELK1
MRPL3
ATP5PF
TMEM14A C1orf52 Affected
TXNDC16 NDUFB6 DCUN1D5 TRHDE
CYB5R4
AC104083.1 CISD2
FOXO1 FOXO3 STAT5B CREM RNA.polymerase.II
LAMTOR3 RELA MEF2D Creb ATF1 HIF1A ESR1
RNF170
C2orf69 Direct interaction
Indirect interaction
Differentially expressed gene

NDUFA4 RASD1 PDGFB


GAD1 JUN JUNB FOS FOSL2
SLC7A11
(observed)

BCL10 SOX9 CSRNP1 ALAS1 NF1 ERBB4


SLC1A3 GAD2 NFKBID
Differential expression Log2FC

ESRRA EGR4 NPAS4 GADD45B ATP5F1B


Transporter GABA synthesis -3 0 3

Transcription regulator Mitochondrial

Fig. 4. Co-expression and Mechanistic Network Analysis of the ASD Neuronal Transcriptome. (A) Hierarchical clustering of neuronal gene co-expression
modules by module eigengenes. Module–diagnosis associations were shown below each module. Enrichment for ASD GWAS common variants is shown for
each module. Enrichment against literature-curated gene lists is shown on the bottom. (B) Module enrichment for neuron subtypes. Expression profiles of
neuron subtypes were obtained from ref. 16. Red asterisks indicate significant enrichment. (C–F) Functional enrichment (Top) and top 50 hub genes (Bottom) for
module Neu-M5 (C), Neu-M6 (D), Neu-M16 (E) and Neu-M17 (F). Edges represent co-expression (Pearson correlation > 0.5). Co-expressed partners with evidence
of protein-protein interaction (PPI) were connected by solid black lines. PPI data was compiled from well-characterized PPI databases (SI Appendix, SI Methods).
(G) Bayesian network representing the relationships among modules (ellipse) and ASD diagnosis (diamond). The thickness of the arrow is proportional to the
number of times that a connection was detected during a bootstrap of 500 times. (H) The estimated gene regulatory network (Bayesian network) for 21 hub
genes in Neu-M5. (I) Dopamine network built by IPA from neuron DEGs (FDR < 0.1, 225 genes). A network of 17 upstream regulators (orange diamonds) were
predicted to drive the changes of 74 observed target genes (23 of which were shown based on node connectivity ≥ 2, ellipse). ELK1 was differentially expressed
in neurons, and it was also predicted as an upstream regulator.

upregulated in ASD, and among its hub genes were several insulin Neu-M16 was enriched in synaptic functions (Fig. 4E and Dataset
signaling pathway components, including insulin-like growth factor S10). Further, cell-type analysis showed that Neu-M16 was also
(IGF) receptor IGF1R, IGF binding protein IGFBP5, insulin recep- highly enriched in excitatory neurons (Fig. 4B), with CAMK2A
tor substrate IRS2 as well as CBL-associated SORBS1 (Fig. 4D and and CAMK2B among its hub genes (Dataset S9). The upregulation
Dataset S9). Our results provided direct molecular-level evidence of Neu-M16 suggested elevated excitatory signal in ASD
that insulin signaling was altered in ASD neurons. neurons.
Among all five significantly disrupted modules, none were We further tested whether significantly disrupted neuronal
enriched for ASD common variants and only one upregulated modules were enriched in any neuron subtypes. Upregulated neu-
module (Neu-M16) showed enrichment in highly confident ASD ronal modules in ASD were only enriched in excitatory neuron
risk genes, as well as in several other curated gene sets (Fig. 4A). subtypes, while enrichment of inhibitory neurons was only

6 of 11 https://doi.org/10.1073/pnas.2206758120 pnas.org
observed for downregulated modules (Fig. 4B). This provides a top upstream regulator with the mechanistic network presented
additional evidence for altered neuronal activity in ASD neurons, in Fig. 4I. This dopamine network contained 17 upstream regulators
consistent with the findings in our DGE analysis. predicted to drive the observed changes of 74 target genes, including
To gain mechanistic insights from our neuronal data, we used AP-1 subunits JUN, JUNB, FOS, and GABA synthetase GAD1
Bayesian network inference (43–45) to distinguish between and GAD2. Moreover, there were seven genes (BCL10, JUN, JUNB,
direct module-trait associations and indirect module-trait cor- KCNJ2, ELL, S1PR1, and SDC2) from this mechanistic network
relations. Our Bayesian network contained 19 nodes, including that overlapped with the 21 genes identified by the Bayesian net-
18 neuronal co-expressed gene modules represented by module work analysis. This further highlighted the involvement of these
eigengenes, plus a binary node representing ASD diagnosis. Our genes in ASD, since they were identified independently by different
results indicated that Neu-M5 was directly associated with ASD, modeling approaches.
conditioned on all other modules (Fig. 4G). This provided fur-
ther evidence that neuroinflammation was directly associated Small Noncoding RNAs Are Selectively Downregulated in ASD Neurons
with ASD in neurons. and Correlate with Altered Local Splicing. When investigating
We further used Bayesian network to prioritize among hub the genes downregulated in ASD neurons more closely, we
genes of Neu-M5 module for future validation. We focused on noticed a striking pattern that 51 out of the 59 neuron-expressed
21 genes within Neu-M5 that were also significant differentially small nucleolar RNA (snoRNA) (46) genes were downregulated
expressed genes (DEGs). The Bayesian network learned from these in ASD neurons, and 13 were significantly downregulated at
21 hub genes is shown in Fig. 4H. Genes with the highest degree P-value < 0.05 (Fig. 5A and Dataset S6). Dysregulation of snoRNAs
of connectivity in this network included BCL10, ELL, and was not observed in bulk tissue (Dataset S1), and snoRNAs were
GTF2IRD2B, which can be potential targets for functional undetectable in a recent ASD single-cell study (12). snoRNAs are
evaluations. involved in the modification and maturation of ribosomal RNAs
As an orthogonal approach, we identified the upstream regulators (rRNAs) and small nuclear RNAs (snRNAs) (47). Interestingly,
and built mechanistic networks in Ingenuity Pathway Analysis both ribosome and spliceosome components were among the most
(IPA® ) with the differentially expressed genes in captured neurons downregulated in ASD neurons (Fig. 3E). Moreover, snRNAs were
(FDR < 0.1, 225 genes, Dataset S13). Dopamine was predicted as also downregulated in ASD neurons, with 23 out of 24 snRNA genes

A 8 Neuon
B TFDP2
ANKRD28
DPP10
PCDH17
GRIA1
NDUFAF6
NPAS2
6 FMNL2
Pearson
PTPRN2
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

LY6H Correlation
GRK4 1.0
-log10 p value

GCC2
ENTPD4 0.5
RAB3A
IPO13
4 SNORA23 MPHOSPH9
0.0
FDR = 0.05
SCYL2 -0.5
SNORA53 PDE4B
SNORA46 SCARNA5 TLN2 -1.0
FBXO9
SNORA26 GOT1
SNORD13 SNORA63 RAPGEF4
2 SNORA63 SNORA76
ARHGEF11
p = 0.05 TXN
SNORA40 SNORA13 SCARNA13
PPP4R3B
SNORD89 GRIA2
NEMF
ARIH1
SNORA74A
SNORA20
SNORA26
SNORD115-1
SCARNA3
SNORA71B
SNORA63
SNORD89
SCARNA5
SNORA53
SCARNA1
SNORA22
SNORA46
SNORA72
SNORA54
SNORD3A
SNORA35
SCARNA6
SCARNA21
SNORA12
SNORA23
SNORD15B
SNORA76
SNORA49
SNORA71D
SNORA13
SNORD13
SNORA42
SNORA80
SNORD3B-2

0
-2 0 2
log2FoldChange

C D
10.0 p = 0.0027
3
chr10:101163631-101165513

snoRNAs
7.5
snoRNA expression

SCARNA1
2
normalized PSI

SCARNA5
GOT1 Intron

SNORA12
5.0 SNORA13 1
SNORA23
SNORA46
2.5 SNORA53 0
SNORA63
SNORD15B
0.0 -1
SNORD89

-2.5 -2
-2 -1 0 1 2 3 Control Autism
GOT1 Intron chr10:101163631-101165513 PSI

Fig. 5. Coordinated Dysregulation of snoRNAs in ASD Neurons. (A) Volcano plot showing differentially expressed genes in ASD neurons compared to control.
snoRNA genes were colored red. (B) Significant correlation between snoRNA expression (x-axis) and intron PSI (y-axis). Introns are labeled with the name of
overlapping gene locus. Gene loci that are correlated with more than 3 snoRNAs were shown. (C) Scatter plot showing the correlation between GOT1 intron 5
PSI and expression levels of multiple snoRNAs across all neuron samples. Also shown were fitted regression lines with 95% CIs. Intron coordinates were based
on GRCh37. (D), PSI of GOT1 intron 5 is downregulated in ASD neurons.

PNAS 2023 Vol. 120 No. 10 e2206758120 https://doi.org/10.1073/pnas.2206758120 7 of 11


Table 1. Summary of genes discussed in this manuscript HSPs may induce immune responses in ASD brain. They also play
a role in facilitating alternative RNA splicing (28). Previous studies
Noted Categories Genes found that both immune response and RNA splicing are upregulated
in ASD brain (7, 8, 11), and our results signify that upregulated
E/I signal CAMK2A*; CAMK2B*; GAD1*; GAD2*;
HSP-related pathways potentially contribute to these observations.
GOT1*
HSPs and HSP-related chaperones are normally induced in response
Growth / differentiation FOS*; FOSL2*; IRF3*; JUN*; JUNB*;
to stress. The upregulation of HSPs in ASD neurons may relate to
PPP1R16B*; S1PR1*; SOX9*
elevated endoplasmic reticulum (ER) stress since both UPR and
Heat shock BAG3†; DNAJB1†; DNAJB4†; HSPA1A†;
apoptosis are also upregulated in our ASD bulk data (Fig. 2D). ASD-
HSPA1B†; HSPB1†; HSPB8†;
PTGES3†
linked rare or de novo mutations in synaptic genes can lead to mis-
folded proteins and cause ER stress (55), itself coupled to heightened
Immune / inflammation BCL10*; NFKBID*
inflammation and neurotoxic cell death (56). ER stress-related genes
Insulin signaling IGF1R*; IGFBP5*; IRS2*; SORBS1* are also dysregulated in the middle frontal cortex of subjects with
Ion channel KCNH3†; KCNIP1†; KCNJ10*; KCNJ2*; ASD (57). Our data suggest that ER stress is a major response to
SCN1B*; SLC40A1* ASD genetic mutations, and ER stress activates UPR, including the
Mitochondrial ATP5F1B*; ATP5PF* production of HSPs and chaperones. UPR further induces multiple
Age-associated GAD1†; GAD2†; HTRA2*; PCDH- downstream processes such as inflammation and immune response
B18P*; ZNF765* (58). Limiting the effect of ER stress and UPR may be a promising
*
observed in STG neurons. therapeutic avenue for ASD.

observed in bulk STG.
The category of each gene is based on what we discussed in this manuscript.
ASD Neuronal Transcriptome Reveals Upregulated
Neuroinflammation and Altered Neuronal Activity. We observed a
downregulated in ASD and 13 significantly downregulated at P value
< 0.05 (SI Appendix, Fig. S5 and Dataset S6) Table 1. strong upregulation of AP-1 transcription factor components in ASD
As snoRNAs and snRNAs are known to be critical regulators of neurons. AP-1 subunits FOS, JUN and JUNB were upregulated at
alternative splicing (48–51), and splicing alterations are strongly FDR < 0.05, and FOSL2 was upregulated at nominal P < 0.05. AP-1
implicated in ASD and observed in LCM-captured neurons, we next regulated gene expression in response to various stimuli, including
examined alterations in splicing events that may be associated with cytokines, growth factors, stress signals, infections and inflammation/
snoRNA dysregulation. We calculated the correlation between neuroinflammation (35). In ASD neurons, AP-1 activation likely
snoRNA expression level and each local splicing event in neuronal induces broad inflammatory response, since several immune and
data, followed by FDR correction for multiple comparisons. We inflammation-related genes were also strongly upregulated. These
identified 835 gene loci in neurons with at least one intron whose included NFKBID and BCL10, which were involved in the NF-κB
percent spliced in (PSI) was significantly correlated with snoRNA pathway and were upregulated at FDR < 0.05. In addition, the
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

gene expression (Dataset S11). Of these 835 intron clusters, 196 interferon regulatory factor IRF3 is upregulated at P < 0.05, and IRF3
were significantly dysregulated in ASD neurons (Dataset S11). is co-expressed with AP-1 subunits. The simultaneous upregulation
Several intron clusters correlated with multiple snoRNAs and cor- of AP-1 subunits, NFκB-related genes and interferon regulatory
responded to genes involved in synaptic functions (Fig. 5B). For factors suggested that immune and inflammation responses were
example, GOT1 encodes the glutamic-oxaloacetic transaminase activated in ASD neurons. Upregulated immune response and
known to function as an essential regulator of glutamate level (52). neuroinflammation have been consistently observed in ASD
PSI of intron 5 of GOT1 was highly correlated with the expression patients by bulk tissue transcriptomic studies largely implicating
level of multiple snoRNA genes (Fig. 5C). GOT1 intron 5 was also glial cells (7, 8). However, our results demonstrated that immune/
differentially spliced between ASD and control (Fig. 5D). DS of neuroinflammatory response was clearly activated in ASD neurons,
GOT1 gene may change the level of glutamate and thus leads to an and may be mediated by transcription factor AP-1.
imbalance of E/I in neuronal communication in ASD neurons. We also observed strong downregulation in ASD STG neurons
of GAD1 and GAD2 genes, involved in the biosynthesis of the
Discussion inhibitory neurotransmitter GABA. In contrast, CAMK2A and
CAMK2B genes, which are essential for aspects of plasticity at
Bulk Tissue Transcriptome Findings Reveal Downregulated glutamatergic excitatory synapses, are upregulated at nominal
Neuronal and Synaptic Function Processes, Upregulation of significance. In addition, co-expressed gene modules that were
HSPs, and Unfolded Protein Response (UPR) in ASD Temporal upregulated in ASD neurons were mainly enriched in excitatory
Cortex. Our bulk tissue analyses revealed a potential causal role neurons, while the downregulated module was primarily enriched
of downregulated neuronal processes and synaptic functions in in inhibitory neurons. These data supported the hypothesis that
ASD etiology, consistent with findings from previous bulk-tissue ASD reflects imbalance of E/I in neuronal communication, also
transcriptomic studies on the same brain region (7). Previous studies reported in several brain regions in ASD (19, 20, 39). This is a
also reported dysregulated alternative splicing events in ASD brain report providing molecular-level evidence for imbalance of E/I in
(8, 53). DS analysis in bulk STG found several synaptic genes, neuronal communication specifically in STG neurons in ASD.
including CANCA2D1, CAMK4, CLASP2, CNTNAP1, EPHB1, Future studies will focus on the role of snoRNAs in ASD neurons,
KALRN, NRXN3, SOS2 and SYNGAP1, differentially spliced in and other long and small modulatory non-coding RNAs. Given the
ASD. SYNGAP1 isoforms have been shown to differentially regulate emerging role of snoRNAs as alternative splicing regulators (50,
synaptic plasticity and dendritic development (54). This further 51), we hypothesize that a coordinated downregulation of multiple
signifies the importance of studying alternatively spliced isoforms snoRNAs correlates with elevated dysregulation of local splicing
in ASD brain. events in ASD neurons. Our data provide evidence supporting this
We also observed a coordinated upregulation of multiple HSPs hypothesis; however, no causal relationship can be determined. It
and HSP-related chaperones in ASD STG. HSPs can serve as acti- will be critical to determine if snoRNA dysregulation plays a causal
vators and regulators of the immune system (18), and upregulated role in ASD etiology, and if so, pinpoint the underlying mechanism

8 of 11 https://doi.org/10.1073/pnas.2206758120 pnas.org
and possibly relate to transcript isoforms driving ASD brain and hypothesis that GABAergic inhibitory neurons are disproportionally
neuronal phenotype. affected in ASD throughout the lifespan.
In our study, the concordance of ASD-associated transcriptomic
changes in bulk tissue and LCM neurons was relatively low. This Mechanistic Modeling Highlighted Pathways and Genes That
is not surprising given bulk tissue contains various different neu- Were Directly Associated with ASD in Neurons. We applied two
ronal and non-neuronal cell types. Neuron-specific changes could orthogonal approaches to gain mechanistic insight from our LCM
be masked by changes in other cell types when profiling bulk neuron data. Bayesian network inference built a network from
tissue. The lack of concordance highlighted the necessity of spe- co-expressed modules and further prioritized hub genes within
cifically isolating and profiling each cell type to identify cell-type- selected module(s), identifying BCL10, ELL and GTF2IRD2B as
specific transcriptomic signatures. genes having driving roles in ASD. The approach from Ingenuity
Pathways Analyses first identified upstream regulators that align
Age-Associated Differential Expression in STG Points to Altered with the observed expression changes in our dataset and elucidated
Neuronal Activity in ASD. In bulk tissue, genes involved in that they share a predicted association with dopamine signaling
GABAergic signaling (GAD1 and GAD2) were upregulated with pathways. Expansion of the regulatory network to include another
age in controls, while downregulated with age in ASD. Multiple layer of DEGs in our dataset indicated strong overlap and regulatory
lines of evidence have pointed to deviations in the production, connections to many key genes involved in neuronal function and
transmission, and reception of GABAergic inhibitory interneurons, inflammation (GAD1, GAD2, and many transcriptional regulators).
including decreased numbers of GABAergic interneurons (especially Interestingly, the expansion of the IPA network further converged on
Parvalbumin neurons) (59) and reduced density of GABA receptors regulatory molecules BCL10 and RNA polymerase II, of which the
(60–62). Our findings indicated that the reduction of GAD1 and former was directly identified in the independent Bayesian approach,
GAD2 mRNA levels in ASD brain became more profound with and the latter associated with ELL. Together these two independent
age, consistent with the observations at cellular level. In addition, approaches highlight the role of inflammatory pathways and genes
SLC38A1, involved in neurotransmission at glutaminergic and in ASD etiology.
GABAergic synapses (63), was downregulated in ASD relative to
control STG. SLC38A1 is implicated in Rett Syndrome (64) and Limitation and Future Studies
mitochondrial disorders, and its decrease may contribute to the Although our study had a relatively large sample size of 59 human
observed alterations in synapse formation and neural connectivity. brains, there are inherent limitations to utilizing postmortem human
In LCM neurons, the expression of HTRA2 was higher below brain tissue, including variability in cause of death, co-morbid con-
age 30 and decreases with age in control neurons, while lower at ditions, medication use, postmortem interval (PMI), etc., Future
younger ages and increasing with age in ASD neurons. HTRA2 studies will have the advantage of more in-depth clinical information
is important in maintaining mitochondrial homeostasis (65) and available from newer brain tissue collections such as Autism BrainNet
inducing apoptosis. It is implicated in pathogenesis of neurode- (SFARI) and the NIH Biobanks to evaluate the relationship of cel-
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

generation, hypoxic-ischemic damage, and is proposed as a poten- lular and molecular findings with clinical characteristics. Most of
tial treatment target in neurological diseases (66). Attenuated our subjects were male, consistent with the 4:1 ratio of males/females
HTRA2 activity may lead to neuronal cell death, altered chaperon with ASD; future studies should include a larger female cohort to
activity and autophagy and has been linked to Parkinson’s disease adequately evaluate sex differences. Additionally, the development
(67). In addition, increased active form of the OMI/HTRA2 serine of novel tools for profiling the transcriptome such as scRNA-seq
protease has been positively correlated with cholinergic alterations allows for analyses of individual cell types at high resolution; however,
in AD brain (68). Thus, it is plausible that the altered expression snRNA-seq currently does not provide the depth of coverage to
of HTRA2 with age we observed in ASD brain may be associated quantify low-expressed genes and local splicing, especially in human
with neuronal alterations during development. These findings brain samples that we were able to achieve with LCM neuron sam-
further support the hypothesis of altered neuronal E/I activity, ples. Our study, however, did not distinguish neuronal subtypes, as
neuroinflammation, cell death, and mitochondrial dysfunction, we found Histogene the most reliable cell body staining method for
implicated in ASD (69) and suggest treatment windows to target maintaining tissue quality and accurate neuronal identification. Our
specific genes to alter their expression trajectories with age. future studies will expand on current findings, particularly age-re-
In summary, our age-related findings support the premise brain lated changes in GAD implicating GABAergic interneurons, to map
development in individuals with ASD deviates from that of the neu- expression in specific cell types in multiple brain regions, in relation
rotypical trajectory beginning in childhood and continues to evolve to clinical characteristics, across the human lifespan.
across the lifespan (70, 71). Although the STG remains relatively
unexplored (72–74), other brain regions display early excess followed
by reductions in volume, connectivity, and cell densities as people Methods
with ASD age through adulthood. Initial excess and overconnectivity
Bulk Tissue RNA Extraction and Library Preparation. Tissue and clinical data
may lead to hyperexcitation, rendering the brain vulnerable to age-
collection procedures were approved by the institutional review board (IRB) and
related and pro-inflammatory mechanisms contributing to later
Human and Anatomical Specimens Tissue Oversight Committee (HASTOC) at the
degenerative outcomes. Our findings of altered neuroinflammatory University of California, Davis School of Medicine. Informed consent was obtained
expression patterns, taken together with reports of excessive micro- from next-of-kin at the time of brain tissue collection for follow-up to collect donor
glial activation in STG (74, 75), implicate immune dysfunction in clinical information to confirm diagnoses by the Autism Tissue Program (now Autism
the pathophysiology of ASD that may exacerbate with age. BrainNet).
Additionally, we found significant relationships with our ASD tran- Human brain tissue was collected, sectioned coronally and flash-frozen. The
scriptome profiles and AD profiles in the same brain region (25), STG was identified anatomically according to “Atlas of the Human Brain” fourth
supporting recent theories of increased susceptibility to neurodegen- edition (Maj, Majtanik, Paxinos 2015). Total RNA was extracted using the Direct-zol
erative and cognitive decline (76, 77). Lastly, our findings of an RNA MiniPrep (Zymo Research #R2051).
age-related decrease in GAD expression in bulk STG tissue, and an Fifty nanograms of RNA from each sample were used to create strand-specific
overall downregulation of GAD in neurons, further supports the total RNA libraries with the NuGEN Ovation Universal RNA-Seq System v2 and

PNAS 2023 Vol. 120 No. 10 e2206758120 https://doi.org/10.1073/pnas.2206758120 9 of 11


processed in parallel on the Sciclone NGS automated workstation (Perkin Elmer) distinguish between modules that are directly associated with ASD and those
according to manufacturer protocol. that are indirectly associated with ASD (45).

Laser Capture Microdissection, RNA Extraction, and Library Preparation. Prioritizing Genes within Neu-M5. To prioritize genes within the Neu-M5 mod-
Using a Leica LMD-6000 laser capture microdissection system, 100 neurons ule, we first identified 21 genes within Neu-M5 that are also significant DEGs.
from each sample were laser captured directly into lysis buffer. RNA was We then built Bayesian network among the 21 genes with method described in
extracted using the PicoPure total RNA kit according to manufacturer protocol. the above section. We selected genes with the highest number of connections as
Strand-specific rRNA depleted RNA libraries were prepared from 10 μL of the nodes likely to influence the expression of many other genes within the module.
final neuronal RNA. Alzheimer’s Disease Datasets. Alzheimer's disease DGE data was obtained
from the Mount Sinai/JJ Peters VA MSBB study (25) through Synapse via accession
RNA Sequencing, Read Mapping, Quantification of Gene Expression, and
number syn30821563.
QC. RNA-Seq was performed on Illumina HiSeq4000. Libraries were sequenced to
~50 million 2 × 150 bp reads per sample. Reads were aligned to the GRCH37.p13 Data, Materials, and Software Availability. All custom code used in this man-
(hg19) reference genome using STAR. Gene-level quantifications were calculated uscript is available at https://github.com/gandallab/ASD_STG_LCM_RNAseq (85).
using featureCounts (v1.6.4). Sample outliers were defined as samples with stand- RNA-seq data from this study is available through the database for Genotypes
ardized sample network connectivity Z scores < −2 (78), and were removed. Using and Phenotypes (dbGaP Study Accession ID: phs003208) (86).
this method, five samples from bulk data, and five samples from neuron data were
removed. Batch effects were accounted for using metrics derived from PicardTools ACKNOWLEDGMENTS. We are grateful to the families of our brain donors
(v2.21.2). for their invaluable gift to autism research. We appreciate the contributions of
Carolyn Komich Hare, clinical Director of Autism BrainNet, for collection of the
DGE and Differential Alternative Splicing. DGE analyses were performed ADI-R as well as Robin Riedel, UC Davis, for figure design. Tissue was provided
using DESeq2 (1.22.2)(79) with default parameters. Local splicing analysis was by the University of Maryland Brain and Tissue Bank, Dr. Daniel Campbell
performed using LeafCutter (26) as previously described (10). at Michigan State University, Autism Tissue Program (now Autism BrainNet,
supported by the Simons Foundation), Brain Endowment for Autism Research
Co-expression Network Analysis. Weighted gene co-expression network
Sciences (BEARS) at the UC Davis MIND Institute, and the Harvard Brain Tissue
analysis (WGCNA) (27) was performed to define modules of co-expressed genes
Resource Center. Laser capture microdissection was conducted using the CAMI
from RNA-seq data.
core facility at UC Davis Center for Health and the Environment with instrumen-
Functional Enrichment Analysis. Gene ontology (GO) enrichment was per- tation funding by NIHS10RR-023555 (carried out by A.O.). RNA sequencing
formed using the gProfileR R package (80) and the fGSEA algorithm as imple- was conducted using the Vincent J. Coates Genomics Sequencing Laboratory
mented in the clusterProfiler R package (81). at UC Berkeley, supported by NIH S10 OD018174 Instrumentation Grant.
Enrichment analyses were also performed using several established, hypoth- M.J.G.’s contributions were supported by a SFARI Bridge to Independence
esis-driven gene sets. Significance was determined from permutation-derived Award, R01-MH121521, R01-MH123922 and IDDRC P50-HD103557 at UCLA.
null distribution. This research was supported by NIH grants MH108909 (MPI C.M.S. and B.S.),
R01 MH097236 and R21 MH 119650 (C.M.S), and the Intellectual and
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

Cell Type Enrichment Analysis. Cell-type enrichment analysis was performed Developmental Disabilities Research Center (IDDRC) at the UC Davis MIND
using the Expression Weighted Cell Type Enrichment (EWCE) package in R (82). Institute (1U54HD079125).
GWAS Enrichment Analysis. Stratified LD score regression (sLDSC) (83) was
used to test whether a gene set of interest is enriched for SNP-heritability in a
Author affiliations: aDepartment of Psychiatry and Biobehavioral Sciences, David Geffen
given GWAS dataset. School of Medicine, University of California, Los Angeles, CA 90095; bDepartment of
Human Genetics, David Geffen School of Medicine, University of California, Los Angeles,
Ingenuity Pathway Analysis. Ingenuity Pathway Analysis (IPA®, QIAGEN) tools CA 90095; cSemel Institute for Neuroscience and Human Behavior, University of
were used to predict the upstream regulators and to build mechanistic networks (84). California, Los Angeles, CA 90095; dDepartment of Psychiatry and Behavioral Sciences,
School of Medicine, University of California, Davis, Sacramento, CA 95817; eUniversity
of California, Davis, MIND Institute, Sacramento, CA 95817; fDepartment of Neurology,
Designing and Learning the Bayesian Network Structure. Each of the 18 University of California, Davis, School of Medicine, Sacramento, CA 95817; gDepartment
module eigengenes from our neuron co-expression data serves as one random of Psychiatry, Perelman School of Medicine, University of Pennsylvania, Philadelphia,
PA 19104; hDepartment of Genetics, Perelman School of Medicine, University of
variable (node) in our Bayesian network (BN). In addition, we also include diag- Pennsylvania, Philadelphia, PA 19104; and iLifespan Brain Institute, Penn Med and the
nosis (ASD or CTL) as a binary random variable (node). This design allows us to Children’s Hospital of Philadelphia, Philadelphia, PA 19104

1. Y. S. Kim et al., Prevalence of autism spectrum disorders in a total population sample. Am. J. 12. D. Velmeshev et al., Single-cell genomics identifies cell type-specific molecular changes in autism.
Psychiatry. 168, 904–912 (2011). Science 364, 685–689 (2019).
2. M. J. Maenner et al., Prevalence of autism spectrum disorder among children aged 8 years - autism 13. M. J. Gandal et al., Broad transcriptomic dysregulation occurs across the cerebral cortex in ASD.
and developmental disabilities monitoring network, 11 sites, United States, 2016. MMWR Surveill. Nature 611, 532–539 (2022).
Summ. 69, 1–12 (2020). 14. R. J. Jou, N. J. Minshew, M. S. Keshavan, M. P. Vitale, A. Y. Hardan, Enlarged right superior
3. C. Lord et al., Autism spectrum disorder. Nat. Rev. Dis. Primers 6, 5 (2020). temporal gyrus in children and adolescents with autism. Brain Res. 1360, 205–212
4. S. N. Hansen et al., Recurrence risk of autism in siblings and cousins: A multinational, population- (2010).
based study. J. Am. Acad. Child Adolesc. Psychiatry 58, 866–875 (2019). 15. D. G. Amaral, C. M. Schumann, C. W. Nordahl, Neuroanatomy of autism. Trends Neurosci. 31,
5. L. Castelbaum, C. M. Sylvester, Y. Zhang, Q. Yu, J. N. Constantino, On the nature of monozygotic twin 137–145 (2008).
concordance and discordance for autistic trait severity: A quantitative analysis. Behav. Genet. 50, 16. D. Wang et al., Comprehensive functional genomic resource and integrative model for the human
263–272 (2020). brain. Science 362, 8464 (2018).
6. T. Gaugler et al., Most genetic risk for autism resides with common variation. Nat. Genet 46, 17. M. E. Feder, G. E. Hofmann, Heat-shock proteins, molecular chaperones, and the stress response:
881–885 (2014). Evolutionary and ecological physiology. Annu. Rev. Physiol. 61, 243–282 (1999).
7. I. Voineagu et al., Transcriptomic analysis of autistic brain reveals convergent molecular pathology. 18. R. J. Binder, Functions of heat shock proteins in pathways of the innate and adaptive immune
Nature 474, 380–384 (2011). system. J. Immunol. 193, 5765–5771 (2014).
8. N. N. Parikshak et al., Genome-wide changes in lncRNA, splicing, and regional gene expression 19. V. S. Sohal, J. L. R. Rubenstein, Excitation-inhibition balance as a framework for investigating
patterns in autism. Nature 540, 423–427 (2016). mechanisms in neuropsychiatric disorders. Mol. Psychiatry. 24, 1248–1257 (2019).
9. M. J. Gandal et al., Shared molecular neuropathology across major psychiatric disorders parallels 20. S. B. Nelson, V. Valakh, Excitatory/inhibitory balance and circuit homeostasis in autism spectrum
polygenic overlap. Science 359, 693–697 (2018). disorders. Neuron 87, 684–698 (2015).
10. M. J. Gandal et al., Transcriptome-wide isoform-level dysregulation in ASD, schizophrenia, and 21. M. L. Chow et al., Age-dependent brain gene expression and copy number anomalies in autism suggest
bipolar disorder. Science 362, eaat8127 (2018). distinct pathological processes at young versus mature ages. PLoS Genet. 8, e1002592 (2012).
11. S. Gupta et al., Transcriptome analysis reveals dysregulation of innate immune response genes and 22. B. P. Grone, K. P. Maruska, Three distinct glutamate decarboxylase genes in vertebrates. Sci. Rep. 6,
neuronal activity-dependent genes in autism. Nat. Commun. 5, 5748 (2014). 30507 (2016).

10 of 11 https://doi.org/10.1073/pnas.2206758120 pnas.org
23. E. Patrick et al., Deconvolving the contributions of cell-type heterogeneity on cortical gene 57. A. Crider, A. O. Ahmed, A. Pillai, Altered expression of endoplasmic reticulum stress-related genes in
expression. PLoS Comput. Biol. 16, e1008120 (2020). the middle frontal cortex of subjects with autism spectrum disorder. Mol. Neuropsychiatry 3, 85–91
24. E. Magnin, Neurodevelopmental and neurodegenerative similarities and interactions: A point of (2017).
view about lifelong neurocognitive trajectories. J. Alzheimers Dis. 79, 1397–1407 (2021). 58. M. I. Timberlake, Y. Dwivedi, Linking unfolded protein response to inflammation and depression:
25. M. Wang et al., The Mount Sinai cohort of large-scale genomic, transcriptomic and proteomic data in Potential pathologic and therapeutic implications. Mol Psychiatry 24, 987–994 (2019).
Alzheimer’s disease. Sci. Data 5, 1–16 (2018). 59. E. Hashemi, J. Ariza, H. Rogers, S. C. Noctor, V. Martínez-Cerdeño, The number of parvalbumin-
26. Y. I. Li et al., Annotation-free quantification of RNA splicing using LeafCutter. Nat. Genetics 50, expressing interneurons is decreased in the prefrontal cortex in autism. Cerebral Cortex 27,
151–158 (2018). 1931–1943 (2017).
27. P. Langfelder, S. Horvath, WGCNA: An R package for weighted correlation network analysis. BMC 60. G. J. Blatt et al., Density and distribution of hippocampal neurotransmitter receptors in autism: An
Bioinformatics 9, 559 (2008). autoradiographic study. J. Autism Dev. Disorders 31, 537–543 (2001).
28. H. J. Yost, S. Lindquist, RNA splicing is interrupted by heat shock and is rescued by heat shock 61. A. L. Oblak, T. T. Gibbs, G. J. Blatt, Reduced GABAA receptors and benzodiazepine binding sites in the
protein synthesis. Cell 45, 185–193 (1986). posterior cingulate cortex and fusiform gyrus in autism. Brain Res. 1380, 218–228 (2011).
29. F. K. Satterstrom et al., Large-scale exome sequencing study implicates both developmental and 62. A. L. Oblak, T. T. Gibbs, G. J. Blatt, Decreased GABAB receptors in the cingulate cortex and fusiform
functional changes in the neurobiology of autism. Cell 180, 568–84.e23 (2020). gyrus in autism. J. Neurochem. 114, 1414–1423 (2010).
30. B. S. Abrahams et al., SFARI Gene 2.0: A community-driven knowledgebase for the autism spectrum 63. T. Qureshi et al., The glutamine transporter Slc38a1 regulates GABAergic neurotransmission and
disorders (ASDs). Mol. Autism 4, 36 (2013). synaptic plasticity. Cereb. Cortex 29, 5166–5179 (2019).
31. K. E. Samocha et al., A framework for the interpretation of de novo mutation in human disease. Nat. 64. L.-W. Jin et al., Dysregulation of glutamine transporter SNAT1 in Rett syndrome microglia: A
Genet 46, 944–950 (2014). mechanism for mitochondrial dysfunction and neurotoxicity. J. Neurosci. 35, 2516–2529 (2015).
32. K. J. Karczewski et al., The ExAC browser: Displaying reference data information from over 60 000 65. L. V. Walle, M. Lamkanfi, P. Vandenabeele, The mitochondrial serine protease HtrA2/Omi: An
exomes. Nucleic Acids Res. 45, D840–D855 (2017). overview. Cell Death Differ. 15, 453–460 (2008).
33. B. Wilkinson et al., The autism-associated gene chromodomain helicase DNA-binding protein 8 66. X. J. Su, L. Huang, Y. Qu, D., Mu Progress in research on the role of Omi/HtrA2 in neurological
(CHD8) regulates noncoding RNAs and autism-related genes. Transl. Psychiatry. 5, e568 (2015). diseases. Rev. Neurosci. 30, 279–287 (2019).
34. J. Grove et al., Identification of common genetic risk variants for autism spectrum disorder. Nat. 67. H. Plun-Favreau et al., The mitochondrial protease HtrA2 is regulated by Parkinson’s disease-
Genet 51, 431–444 (2019). associated kinase PINK1. Nat Cell Biol. 9, 1243–1252 (2007).
35. J. Hess, P. Angel, M. Schorpp-Kistner, AP-1 subunits: Quarrel and harmony among siblings. J. Cell 68. T. Darreh-Shori et al., Increased active OMI/HTRA2 serine protease displays a positive correlation
Sci. 117, 5965–5973 (2004). with cholinergic alterations in the Alzheimer’s disease brain. Mol. Neurobiol. 56, 4601–4619
36. E. C. Warburton et al., Cholinergic neurotransmission is essential for perirhinal cortical plasticity and (2019).
recognition memory. Neuron 38, 987–996 (2003). 69. K. Kawada, N. Kuramoto, S. Mimori, Possibility that the onset of autism spectrum disorder is induced
37. A. Binda et al., A novel KCNJ2 mutation identified in an autistic proband affects the single channel by failure of the glutamine-glutamate cycle. Curr. Mol. Pharmacol. 14, 170–174 (2021).
properties of Kir2.1. Front. Cell Neurosci. 12, 76 (2018). 70. C. Ecker, S. Y. Bookheimer, D. G. Murphy, Neuroimaging in autism spectrum disorder: Brain structure
38. E. Ambrosini et al., Genetically induced dysfunctions of Kir2.1 channels: Implications for short QT3 and function across the lifespan. Lancet Neurol. 14, 1121–1134 (2015).
syndrome and autism-epilepsy phenotype. Hum. Mol. Genet. 23, 4875–4886 (2014). 71. T. A. Avino et al., Neuron numbers increase in the human amygdala from birth to adulthood, but not
39. J. L. R. Rubenstein, M. M. Merzenich, Model of autism: Increased ratio of excitation/inhibition in key in autism. Proc. Natl. Acad. Sci. U.S.A. 115, 3710–3715 (2018).
neural systems. Genes Brain Behav. 2, 255–267 (2003). 72. B. P. Ander, N. Barger, B. Stamova, F. R. Sharp, C. M. Schumann, Atypical miRNA expression in
40. F. Renoux et al., The AP1 transcription factor Fosl2 promotes systemic autoimmunity and temporal cortex associated with dysregulation of immune, cell cycle, and other pathways in autism
inflammation by repressing treg development. Cell Rep. 31, 107826 (2020). spectrum disorders. Mol. Autism 6, 1–13 (2015).
41. S. E. Sweeney, T. B. Kimbler, G. S. Firestein, Synoviocyte innate immune responses: II. Pivotal role of 73. B. Stamova, B. P. Ander, N. Barger, F. R. Sharp, C. M. Schumann, Specific regional and age-related
IFN regulatory factor 3. J. Immunol. 184, 7162–7168 (2010). small noncoding RNA expression patterns within superior temporal gyrus of typical human brains
42. M. Iwanaszko, M. Kimmel, NF-κB and IRF pathways: Cross-regulation on target genes promoter are less distinct in autism brains. J. Child Neurol. 30, 1930–1946 (2015).
level. BMC Genomics 16, 307 (2015). 74. K. Garbett et al., Immune transcriptome alterations in the temporal cortex of subjects with autism.
43. N. Friedman, M. Linial, I. Nachman, D. Pe’er, Using Bayesian networks to analyze expression data. J. Neurobiol. Dis. 30, 303–311 (2008).
Comput. Boil. 7, 601–620 (2000). 75. A. S. Lee, E. C. Azmitia, P. M. Whitaker-Azmitia, Developmental microglial priming in postmortem
44. B. Zhang et al., Integrated systems approach identifies genetic nodes and networks in late-onset autism spectrum disorder temporal cortex. Brain Behav. Immun. 62, 193–202 (2017).
Alzheimer’s disease. Cell 153, 707–720 (2013). 76. G. Vivanti, S. Tao, K. Lyall, D. L. Robins, L. L. Shea, The prevalence and incidence of early-onset
Downloaded from https://www.pnas.org by 177.159.17.251 on May 3, 2024 from IP address 177.159.17.251.

45. S. Mostafavi et al., A molecular network of the aging human brain provides insights into the dementia among adults with autism spectrum disorder. Autism Res. 14, 2189–2199 (2021).
pathology and cognitive decline of Alzheimer’s disease. Nat. Neurosci. 21, 811–819 (2018). 77. B. N. Hand, A. M. Angell, L. Harris, L. A. Carpenter, Prevalence of physical and mental health
46. T. Kiss, Small nucleolar RNAs: An abundant group of noncoding RNAs with diverse cellular functions. conditions in Medicare-enrolled, autistic older adults. Autism 24, 755–764 (2020).
Cell 109, 145–148 (2002). 78. M. C. Oldham, P. Langfelder, S. Horvath, Network methods for describing sample relationships in
47. T. Bratkovič, J. Božič, B. Rogelj, Functional diversity of small nucleolar RNAs. Nucleic Acids Res. 48, genomic datasets: Application to Huntington’s disease. BMC Syst. Biol. 6, 63 (2012).
1627–1651 (2020). 79. M. I. Love, W. Huber, S. Anders, Moderated estimation of fold change and dispersion for RNA-seq
48. H. Dvinge, J. Guenthoer, P. L. Porter, R. K. Bradley, RNA components of the spliceosome regulate data with DESeq2. Genome Biol. 15, 1–21 (2014).
tissue- and cancer-specific alternative splicing. Genome Res. 29, 1591–1604 (2019). 80. U. Raudvere et al., g: Profiler: A web server for functional enrichment analysis and conversions of
49. S. Valadkhan, L. S. Gunawardane, Role of small nuclear RNAs in eukaryotic gene expression. Essays gene lists (2019 update). Nucleic Acids Res. 47, W191–W198 (2019).
Biochem. 54, 79–90 (2013). 81. G. Yu, L.-G. Wang, Y. Han, Q.-Y., He, clusterProfiler: An R package for comparing biological themes
50. S. Kishore, S. Stamm, The snoRNA HBII-52 regulates alternative splicing of the serotonin receptor among gene clusters. OMICS 16, 284–287 (2012).
2C. Science 311, 230–232 (2006). 82. N. G. Skene, S. G. Grant, Identification of vulnerable cell types in major brain disorders using
51. M. Falaleeva et al., Dual function of C/D box small nucleolar RNAs in rRNA modification and single cell transcriptomes and expression weighted cell type enrichment. Front Neurosci. 10, 16
alternative pre-mRNA splicing. Proc. Natl. Acad. Sci. U.S.A. 113, E1625–E1634 (2016). (2016).
52. A. D’Aniello et al., Amino acids and transaminases activity in ventricular CSF and in brain of normal 83. H. K. Finucane et al., Partitioning heritability by functional annotation using genome-wide
and Alzheimer patients. Neurosci. Lett. 388, 49–53 (2005). association summary statistics. Nat. Genet. 47, 1228–1235 (2015).
53. B. S. Stamova et al., Evidence for differential alternative splicing in blood of young boys with autism 84. A. Krämer, J. Green, J. Pollard Jr., S. Tugendreich, Causal analysis approaches in ingenuity pathway
spectrum disorders. Mol. Autism 4, 1–16 (2013). analysis. Bioinformatics 30, 523–530 (2014).
54. Y. Araki et al., SynGAP isoforms differentially regulate synaptic plasticity and dendritic development. 85. P. Zhang, M. J. Gandal, ASD_STG_LCM_RNAseq_paper_code. Github. https://github.com/
Elife 9, e56273 (2020). gandallab/ASD_STG_LCM_RNAseq. Deposited 27 January 2023.
55. D. Comoletti et al., The Arg451Cys-neuroligin-3 mutation associated with autism reveals a defect in 86. A. Omanska, B. Stamova, C. M. Schumann, B. P. Ander, Transcriptomic profiling of bulk tissue and
protein processing. J. Neurosci. 24, 4889–4893 (2004). laser-capture micro dissected neurons of postmortem human brains in the superior temporal gyrus.
56. N. T. Sprenkle, S. G. Sims, C. L. Sánchez, G. P. Meares, Endoplasmic reticulum stress and NCBI’s Database of Genotypes and Phenotypes (dbGaP). https://www.ncbi.nlm.nih.gov/projects/gap/
inflammation in the central nervous system. Mol. Neurodegener. 12, 42 (2017). cgi-bin/study.cgi?study_id=phs003208.v1.p1 (Deposited 17 February 2023).

PNAS 2023 Vol. 120 No. 10 e2206758120 https://doi.org/10.1073/pnas.2206758120 11 of 11

You might also like