scispace - formally typeset
Open AccessPosted ContentDOI

Isoform specificity in the mouse primary motor cortex

Reads0
Chats0
TLDR
In a comprehensive analysis of 6,160 mouse primary motor cortex cells assayed with SMART-Seq, numerous examples of isoform specificity in cell types are found, including isoform shifts between cell types that are masked in gene-level analysis.
Abstract
Full-length SMART-Seq single-cell RNA-seq can be used to measure gene expression at isoform resolution, making possible the identification of gene isoform markers for cell types. In a comprehensive analysis of 6,160 mouse primary motor cortex cells assayed with SMART-Seq, we find numerous examples of isoform specificity in cell types, including isoform shifts between cell types that are masked in gene-level analysis. These findings can be used to refine spatial gene expression information to isoform resolution. Our results highlight the utility of full-length single-cell RNA-seq when used in conjunction with other single-cell RNA-seq technologies.

read more

Content maybe subject to copyright    Report

Nature | Vol 598 | 7 October 2021 | 195
Article
Isoform cell-type specificity in the mouse
primary motor cortex
A. Sina Booeshaghi
1
, Zizhen Yao
2
, Cindy van Velthoven
2
, Kimberly Smith
2
, Bosiljka Tasic
2
,
Hongkui Zeng
2
& Lior Pachter
3,4
 ✉
Full-length SMART-seq
1
single-cell RNAsequencing can be used to measure gene
expression at isoform resolution, making possible the identication of specic
isoform markers for dierent cell types. Used in conjunction with spatial RNA capture
and gene-tagging methods, this enables the inference of spatially resolved isoform
expression for dierent cell types. Here, in a comprehensive analysis of 6,160 mouse
primary motor cortex cells assayed with SMART-seq, 280,327 cells assayed with
MERFISH
2
and 94,162 cells assayed with 10x Genomics sequencing
3
, we nd examples
of isoform specicity in cell types—including isoform shifts between cell types that
are masked in gene-level analysis—as well as examples of transcriptional regulation.
Additionally, we show that isoform specicity helps to rene cell types, and that a
multi-platform analysis of single-cell transcriptomic data leveraging multiple
measurements provides a comprehensive atlas of transcription in the mouse primary
motor cortex that improves on the possibilities oered by any single technology.
Transcriptional and post-transcriptional control of individual iso-
forms of genes is crucial for neuronal differentiation
4–8
, and isoforms
of genes have been shown to be specific to cell types in mouse and
human brains
9–14
. It is therefore not surprising that dysregulation of
splicing has been shown to be associated with several neurodevel-
opmental and neuropsychiatric diseases
6,15,16
. Thus, it is of interest to
study gene expression in the brain at single-cell and isoform resolution.
Nevertheless, current single-cell studies aiming to characterize
cell types in the brain using single-cell RNA sequencing (scRNA-seq)
have relied mostly on gene-level analysis. This is, in part, owing to the
nature of the data produced by the highest-throughput single-cell
methods. Popular high-throughput assays such as Drop-seq
17
, 10x
Genomics Chromium
3
and inDrops
18
produce 3′-end reads that are,
in initial pre-processing, collated by gene to produce per-cell gene
counts. SMART-seq
19
is an scRNA-seq method that produces full-length
reads, enabling the quantification of individual isoforms of genes with
the expectation-maximization algorithm
20
. However, such increased
resolution comes at the cost of throughput. SMART-seq requires cells
to be deposited in wells, thus limiting the throughput of the assay. In
addition, SMART-seq requires more sequencing per cell
21
.
The trade-offs are evident in analysis of scRNA-seq data from the pri-
mary motor cortex (MOp) produced by the BRAIN Initiative Cell Census
Network (BICCN)
22
. We examined 6,160 (filtered) SMART-seq v4 cells
and 94,162 (filtered) 10x Genomics Chromium (10xv3) cells (Extended
Data Fig.1, Fig.2a, b) and found that while 10xv3 and SMART-seq are
equivalent in defining broad classes of cell types, 3′-end technol-
ogy that can assay more cells can identify some rare cell types that
are missed at lower cell coverage (Extended Data Fig.2a). Overall, 56
clusters with gene markers could be identified in the 10xv3 data but
not in the SMART-seq data, whereas only 39 clusters with gene markers
could be identified in the SMART-seq data and not the 10xv3 data—this
differential is consistent with previously reported comparisons of
10x Genomics Chromium and SMART-seq clusters
21,23
. However, while
SMART-seq has lower throughput than some other technologies, it has
a notable advantage: because it probes transcripts across their entire
length, SMART-seq makes possible isoform quantification and the
detection of isoform markers for cell types that cannot be detected
with 3′-end technologies (Extended Data Figs.2b, c). Moreover, the
uniformity of read coverage of SMART-seq data
1
and its quantifica-
tion with state-of-the-art tools
24
yields higher sensitivity than other
methods, which can make possible refined cell-type classification.
To take advantage of the complementary strengths of these differ-
ent platforms, we introduce an approach to scRNA-seq that links the
SMART-seq resolved isoforms to the 10x Chromium defined cell types,
and merges this information with spatial transcriptomic measurements
obtained by MERFISH
25
(Fig.1). In addition to revealing extensive iso-
form diversity and cell-type specificity in the MOp, we find evidence
for previously missed transcriptionally distinct cell subtypes in the
MOp. Our results extend the notion of a single-cell database beyond a
list of gene markers, and we produce a gene-isoform-space single-cell
atlas of the MOp using the combined 10xv3, SMART-seq and MERFISH
data. Our methods are open source, reproducible, easy to use and con-
stitute an effective workflow for leveraging full-length scRNA-seq data
in combination with data from other technologies.
Isoform markers for cell types
To identify isoform markers of cell types, we first sought to visualize
our SMART-seq data using gene-derived cluster labels from the BICCN
analysis (Methods). Rather than layering cluster labels on cells mapped
https://doi.org/10.1038/s41586-021-03969-3
Received: 9 March 2020
Accepted: 27 August 2021
Published online: 6 October 2021
Open access
Check for updates
1
Department of Mechanical Engineering, California Institute of Technology, Pasadena, CA, USA.
2
Allen Institute for Brain Science, Seattle, WA, USA.
3
Division of Biology and Biological
Engineering, California Institute of Technology, Pasadena, CA, USA.
4
Department of Computing and Mathematical Sciences, California Institute of Technology, Pasadena, CA, USA.
e-mail: lpachter@caltech.edu

196 | Nature | Vol 598 | 7 October 2021
Article
to 2D with an unsupervised dimensionality-reduction technique such
as t-distributed stochastic neighbour embedding
26
(t-SNE) or uniform
manifold approximation and projection
27
(UMAP), we used a super-
vised learning approach to project cells so that they are best separated
according to BICCN consortium
22
annotations using neighbourhood
component analysis (NCA). This method produces meaningful rep-
resentation of the global structure of the data (Fig.2b), without over-
fitting (Supplementary Fig.1a). Analysis of the projections revealed
a batch effect in the 10xv3 data—which we addressed by restricting
analysis to a single batch—and minimal evidence of a batch effect in
the MERFISH data (Methods, Supplementary Fig.2a, b).
Next, motivated by the discovery of genes exhibiting differential
exon usage between glutamatergic and GABAergic (γ-aminobutyric
acid-producing) neurons in the primary visual cortex
14
, we performed a
differential analysis between these two classes of neurons. We searched
for significant shifts in isoform abundances in genes whose expression
was stable across cell types (Methods). We discovered 398 such isoform
markers belonging to 310 genes (Supplementary Table1). Figure2c
shows an example of such an isoform from the oxidative resistance 1
(Oxr1) gene, which is known to be essential for protection against oxi-
dative stress-induced neurodegeneration
28,29
. While we see no change
in gene expression of Oxr1 between these two neuron types, we find
that among the 16 isoforms of the gene, one of them, Oxr1-204, is more
highly expressed in glutamatergic neurons. The Oxr1 gene undergoes
an isoform shift in GABAergic neurons, where the expression of the
Oxr1-204 isoform is significantly lower, suggesting distinct subcellular
isoform localization in the two neuron types
30
. A gene-level analysis is
blind to this isoform shift (Fig.2c, top right).
We hypothesized that there are genes exhibiting cell-type isoform
specificity at all levels of the MOp cell ontology. However, detection of
such genes and their associated isoforms requires meaningful cell-type
assignments and accurate isoform quantifications. To assess the reli-
ability of the SMART-seq clusters produced by the BICCN
31
, we exam-
ined the correlation in gene expression by cluster with an orthogonal
scRNA-seq technology, the 10xv3 3′-end assay. We clustered 94,162
10xv3 cells, also derived from the MOp, using the same method as the
SMART-seq cells (Methods). The clustering method generates three
hierarchies of cells: classes, subclasses and clusters. The SMART-seq
data have 2 major classes (glutamatergic and GABAergic), 18 subclasses
that subdivide the classes, and 62 clusters that subdivide the subclasses.
The 10x data similarly contain three hierarchies of cells: two major
classes (glutamatergic and GABAergic), 21 subclasses and 85 clusters.
We found high correlation of gene expression between the two assays
at the subclass and cluster levels (Extended Data Fig.3).
Next, we assessed the accuracy of the SMART-seq isoform quantifica-
tion and its concordance with 10xv3 quantifications of isoforms. Since
not all isoforms can be quantified from 10xv3 3′-end data, we examined
only isoforms containing some unique 3′ UTR sequence. This enabled
us to validate the isoform quantifications using a different method
(Methods). To extract isoform quantifications from 10xv3 data in cases
where there was a unique 3′ sequence, we relied on transcript compat-
ibility counts
32
produced by pseudoalignment with kallisto
24
. We were
able to validate the SMART-seq isoform shift predictions at both the
subclass and cluster levels (Extended Data Fig.4). The isoform abun-
dance correlations are slightly lower than those for gene abundance
estimates (Extended Data Fig.3), but sufficiently accurate to identify
significant isoform shifts, consistent with benchmarks showing that
isoforms can be quantified accurately from full-length bulk RNA-seq
33
.
Having validated the cluster assignments and isoform abundance
estimates, we tested for isoform switches for 16 cell subclasses exclud-
ing low quality cells (example in Fig.2d), and then for 48 distinct clusters
for subclasses that have more than one cluster (example in Fig.2e) and
more than 5 cells per cluster (Methods). At the higher level of 16 cell
subclasses, we found a total of 654 isoforms from 550 genes within
the glutamatergic class and 381 isoforms from 332 genes within the
GABAergic class exhibiting isoform shifts among the 16 cell subclasses
despite constant gene abundance (Supplementary Table2a, b). There
are several notable examples of isoform shifts at this level. For exam-
ple, we find a shift in the Snap25-202 isoform, whose expression has
been specifically shown to be correlated with age and to differentially
3
Gene tagging
Drop-seq, inDrops or 10x Chromium
Smart-seq, Oxford Nanopore or
Pacic Biosciences
Detector
3
5
Isoform sequencing
Spatial capture
seqFISH, MERFISH or 10x Visium
Genes Isoforms
Gene
space
Isoform
space
Physical
space
Cell type
identication
Isoform
renement
Spatial
tethering
Fig. 1 | Measuring RNA with multiple platforms. RNA is measured using
gene-tagging techniques such as the 10x Chromium scRNA-seq protocol,
isoform sequencing techniques such as SMART-seq, and spatial RNA-capture
techniques such as MERFISH. High-cell-throughput gene tagging enables
cell-type identification with marker genes and deep full-length isoform
sequencing enables cell-type marker refinement at the isoform level. Spatial
RNA capture coupled with gene tagging and isoform sequencing enables
spatial resolution of cell-type-specific isoform markers. The multi-method
procedure for sampling RNA enables inference of spatially resolved cell-type-
specific isoforms that no single technique could achieve independently
51
.

Nature | Vol 598 | 7 October 2021 | 197
regulate synaptic transmission and synaptic plasticity at central syn-
apses
34,35
. This isoform marks the L6b subclass (Fig.2d). At the cluster
level, we found 923 isoforms from 823 genes exhibiting isoform shifts
among the 48 clusters passing filter despite constant gene abundance
(Supplementary Table3). Another isoform of notethat marks the L6b
Ror1_ 1 cluster, a subset of cells in the L6b subclass, is the Stxbp2-207
isoform whose gene Stxbp2 has previously been detected in the sub-
thalamic nucleus and the posterior hypothalamus
36
.
Assaying both male and female mice enabled us to examine
sex-specific effects in all subclasses except for the L5 IT, which was
excluded owing to batch effect (Supplementary Fig.4). In total, these
subclasses exhibited 418 sex-specific isoforms, averaging 40 iso-
forms per subclass (Supplementary Table7). Unlike a recent study
that reported a sex-specific cell type in the ventromedial nucleus of
the hypothalamus
37
, we do not find any sex-specific subclasses. How-
ever, we observed several autosomal isoforms that were differentially
expressed between male and female mice. Among these, the Shank1-203
isoform is differential in Vip neurons, a finding that refines previous
data showing that Shank1, which has been shown to localize in Purkinje
cells in the cortex
38
, is a sex-specific gene whose expression is regulated
by sex hormones
39
.
We also investigated instances where clusters could be refined
according to isoform expression. After reclustering each 10xv3-derived
cluster using SMART-seq isoform quantifications (seeMethods), we
found that 12 clusters can be split by isoforms. Examining the L6 CT
Grp_1 cluster, we find that the average effect size for differential iso-
forms that split the cluster into two sub-clusters is higher than that for
genes (Extended Data Fig.5). One isoform in particular, which splits
the L6 CT Grp_1 cluster, is a protein-coding isoform of the amyloid
precursor protein gene (App). Dysregulation of splicing for isoforms
of App have been associated with disease pathogenesis in Alzheimer
disease models
40
. Our findings show that isoform-level expression can
help refine cell types in the mouse MOp beyond what is possible using
gene-level expression estimates.
Along with isoforms detectable as differential between cell types
without change in gene abundance, we identified isoform markers
for the classes, subclasses, and clusters in the MOp ontology that are
differential regardless of gene expression. We found 5,658 isoforms
belonging to 3,132 genes that are specific to the glutamatergic and
GABAergic classes (Fig.3, Supplementary Table4), 7,588 isoforms
belonging to 4,171 genes within the glutamatergic class and 4,359 iso-
forms belonging to 2,614 genes within the GABAergic class exhibiting
isoform shifts specific to subclasses (Supplementary Table5a, b), and
for the 48 clusters passing filter, 3,171 isoforms belonging to 2,461 genes
exhibiting isoform shifts in clusters (Supplementary Table6). Together,
these form an isoform catalogue of the MOp (Supplementary Fig.5a, b).
Spatial isoform specificity
Spatial scRNA-seq methods are not currently well suited to directly
probing isoforms of genes owing to the number and lengths of probes
required—however, spatial analysis at the gene level can be refined
to yield isoform-level results by extrapolating SMART-seq isoform
quantifications (Fig.3, Supplementary Fig.5c).
Figure4a, b shows an example of a gene, Pvalb, for which the
SMART-seq quantification reveals that of the two isoforms of the gene,
only one, Pvalb-201, is expressed. Moreover, thiseffect is specific to
the Pvalb cell subclass (Fig.3). In an examination of MERFISH spa-
tial single-cell RNA-seq derived from 64 slices from the MOp region
(Extended Data Fig.6a), the Pvalb subclass, for which Pvalb is a marker,
can be seen to be dispersed throughout the motor cortex spanning all
layers (Extended Data Fig.6b). While MERFISH probes only measure
abundance of Pvalb at the gene level (Fig.4c), extrapolation from the
SMART-seq quantifications can be used to refine the MERFISH result to
reveal the spatial expression pattern of the Pvalb-201 isoform.
This extrapolation can be done systematically. To build a spatial
isoform atlas of the MOp, we identified differentially expressed genes
from the MERFISH data (Supplementary Table8a, b) and for each of
them checked whether there were SMART-seq isoform markers (from
Supplementary Table5a, b). An example of the result is shown in Fig.3,
which displays one gene for each cluster, together with the isoform
label specific to that cluster and the spatial location of the specific
cluster within a slice of the mouse MOp.
b
SMART-Seq10xv3 MERFISH
Reads
15,229,289,828 22,696,236,495 NA
Cells
6,295 94,162 280,327
Genes
10,333 5,891 72
Isoforms
20,319 NA NA
Classes
443
Subclasses
18 22 25
Clusters
62 14799
Cluster ClassSubclass
a
Average detected per cell
I
so
Differential isoform
c
d
e
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
log(TPM + 1)
log(TPM + 1)
log(TPM + 1)
Lamp5: 360
Sncg: 77
Vip: 607
Sst: 406
Pvalb: 521
L5 IT: 1,561
L5/6 NP: 207
L5 PT: 12
L2/3 IT: 482
L6b: 570
VLMC: 6
L6 IT: 395
L6 CT: 904
L6 IT Car3: 5
Endo: 7
Astro: 9
SMC: 19
Low quality : 12
GlutamatergicGABAergic
Oxr1-204 Oxr1
6
4
2
0
6
4
2
0
5
0
5
0
Snap25-202 Snap25
L2/3 IT L5 PT L5/6 NP L6 CT L6 IT L6 IT Car3 L6b
6
4
2
0
6
4
2
0
Stxbp2-207
Stxbp2
L6b
Kcnip1
_1
L6b
Kcnip1
_2
L6b Ror1_1
L6b Ror1_2
L6b Rprm
L6b Shisa6
GABAergic
Glutamatergic
NN
Fig. 2 | Isoform specificity in the absence of gene specificity. a, Overview of
the data analysed. The clustering method used by the BICCN consortium
generates three hierarchies of cells: classes, subclasses and clusters. NA, not
applicable. b, A t-SNE of 10 neighbourhood components of 6,160 SMART-seq
cells coloured according to subclass. Astro, astrocytes; CT, corticothalamic;
endo, endothelial; IT, intratelencephalic; NN, non-neuronal; NP,
near-projecting; PT, pyramidal tract; SMC, smooth muscle cells; VLMC,vascular
lepotomeningeal cells. c, Example of a gene with an isoform specific to the
glutamatergic class. The Oxr1-204 isoform distribution in log1p(transcripts per
million (TPM)) across cells (left) superimposed on the t-SNE of the cells. The
cells belonging to the glutamatergic class are circled. The violin plots of
the gene and isoform distributions show that the gene is not differentially
expressed but the isoform is (right). d, Example of a gene with an isoform
specific to the L6b subclass. The Snap25-202 isoform distribution across cells
(left) superimposed on the t-SNE of the cells. The cells belonging to the L6b
subclass are circled. The violin plots of the gene and isoform distributions show
that the gene is not differentially expressed but the isoform is (right).
e, Example of a gene with an isoform specific to the L6b Ror1_1 cluster. The
Stxbp2-207 isoform distribution in log1p(TPM) across cells (left) superimposed
on the t-SNE of the cells. The cells belonging to the L6b Ror1_1 cluster are
circled. The violin plots of the gene and isoform distributions show that the
gene is not differentially expressed but the isoform is (right). *P < 0.01 between
the group and its complement. In violin plots, white circles represent the mean
and white bars represent the s.d.

198 | Nature | Vol 598 | 7 October 2021
Article
We hypothesized that the mouse MOp exhibits changes in isoform
expression associated with the physical location of cells
41,42
. To deter-
mine whether there are isoforms that increase or decrease in expression
along the depth of the motor cortex, we first estimated the position
of the various layers in the glutamatergic subclasses (Extended Data
Fig.7a, Methods), performed weighted least-squares regression on
the centroids of the subclasses and inferred isoform expression from
the SMART-seq data. While we find many isoforms that exhibit a sig-
nificant change in expression across the depth (Extended Data Fig.7b,
c, Supplementary Table8c), none of the isoforms that pass our filter
exhibit a monotonic change with respect to the mean. This suggests
that non-linear models may be better suited to study isoform variability
across the depth of the mouse MOp.
While direct measurement of isoform abundance may be possible
with spatial RNA-seq technologies such as SEQFISH
37
or MERFISH
2
,
such resolution would require dozens of probes to be assayed per
gene (Supplementary Fig.6), each of which is typically tens of base
pairs in length. Thus, while isoforms can theoretically be detected
in cases where they contain large stretches of unique sequence, the
technology is currently prohibitive for assaying most isoforms, mak-
ing the extrapolation procedure described here of practical relevance
(Supplementary Table9).
Splicing markers
Isoform quantification of RNA-seq can be used to distinguish shifts in
expression between transcripts that share transcriptional start sites
and shifts due to the use of distinct transcription start sites (TSSs).
Investigating such differences can, in principle, shed light on tran-
scriptional versus post-transcriptional regulation of detected isoform
shifts
43,44
. For example, in the glutamatergic class, Ptk2b (Extended
Data Fig.8a) exhibits differential expression of transcripts between
start sites (Extended Data Fig.8b). This gene is known to be associ-
ated with Alzheimer’s disease and its transcript usage is mediated
by genetic variation
45
. We find that isoforms sharing the preferential
start site exhibit no discernible difference in expression (Extended
Data Fig.8c), suggesting that the observed differences result from
cell-type-specific transcriptional, rather than post-transcriptional
regulation. We identified 1,971 isoforms from 128 groups of TSSs where
the TSSs are preferentially expressed in either GABAergic, glutamater-
gic or non-neuronal classes, even when the expression of isoforms
contained within the TSS is constant (Supplementary Table10a, c, d).
Such cases are likely to be instances where the TSS shifts between cell
types are a result of differential promoter usage—that is, the result of
a transcriptional program.
We also examined post-transcriptional programs (Supplementary
Table10b), instances where the TSSs are not differential between
classes, but where there are isoform shifts within TSSs between classes.
We find 31 isoforms from 28 TSS groups that are differential between
classes when the TSS group is not. One such example is expression
of isoforms Rtn1-201 and Rtn1-203, which share the same TSS in the
Rtn1 gene. The glutamatergic class exhibits preferential expression
of Rtn1-201, which was previously shown to be expressed in grey mat-
ter
46
, whereas the GABAergic class does not (Extended Data Fig.9).
These cases are likely to be instances where isoform shifts between
cell types are a result of differential splicing—that is, the result of a
post-transcriptional program.
Spon1
Bcl11b
C1ql3
Olfm3
Syndig1
Sema5a
Nr4a2
Lamp5
Cxcl14
Vip
Sst
Pvalb
Cldn5
MERFISH TPM
100,000
0
100,000
0
100,000
0
20,000
0
50,000
0
100,000
0
250,000
0
20,000
0
250,000
0
150,000
0
0
100,000
0
25,000
0
500,000
0
L5/6 NP (3,482)
L5 PT (6,759)
L2/3 IT (33,757)
L6b (3,487)
L6 IT (12,475)
L6 CT (23,878)
L6 IT Car3 (1,563)
Lamp5 (3,095)
Sncg (477)
Vip (2,789)
Sst (4,853)
Pvalb (7,885)
Endo (18,183)
201
201
201
201
202
201
201
202
201
201
201
201
201
Fig. 3 | Isoform atlas. Spatial isoform atlas of the MOp. The scatter plots (left)
show the locations of cells (black dots) in the indicated subclass within a single
representative slice of the mouse MOp as assayed by MERFISH. Right, each
column corresponds to a marker gene in the MERFISH datasetand each row
corresponds to a subclass (number of cells in parentheses) in which one
isoform (labelled on the diagonal) was differential in the SMART-seq dataset for
that subclass. This spatial isoform inference links isoform expression from the
SMART-seq data with the physical locations of cells expressing that isoform
from the MERFISH data. The normalized gene expression values are plotted for
each subclass–gene pair in TPM. The white circles in the violin plots represent
the mean and the white bars represent the s.d.
a
b
c
Pvalb-201 Pvalb-202
y (μm)
x (μm)
Fig. 4 | Spatial extrapolation of isoform expression. a, b, Expression of the
Pvalb-201 (a) and Pvalb-202 (b) isoforms in log1p(TPM) units for each cell
superimposed on the NCA–t-SNE plot, as assayed by SMART-seq. c, Spatial
expression of the Pvalb-201 isoform across 64 slices from the MOp, as
extrapolated from probes for the Pvalb gene assayed by MERFISH. Each cell
represented by MERFISH data is coloured by its expression of Pvalb in
normalized counts.

Nature | Vol 598 | 7 October 2021 | 199
Discussion
Our spatially resolved isoform atlas of the mouse MOp expands on
previously identified gene markers, extending the catalogue to isoform
markers for cell types characterized by the BICCN
22
. Our approach
leverages distinct strengths of different technologies, using the iso-
form resolution of SMART-seq in conjunction with the complementary
cell depth obtainable with 10x Genomics technology and the spatial
resolution produced with MERFISH to spatially place cell-type isoform
markers. This validated approach, in which we leverage technologies
that are broadly consistent (Extended Data Fig.10) yet complemen-
tary in their strengths, is important because isoform specificity could
helpinexplaining the molecular basis ofmorphological differences.
For example, Pvalb cells observed in hippocampal Pvalb interneu-
rons cannot be distinguished morphologically solely on the basis of
gene-level analysis
9,47,48
.
Spatial isoform markers also enable more targeted assays for ‘auto-
matic expression histology’and can facilitate investigation ofthe
functional significance of cell-type isoform specificity.Recently devel-
oped experimental methods for this purpose—for example, isoform
screens
49
—are a promising direction and will be key to understanding
the significance of the vast isoform diversity in the brain
50
.
Online content
Any methods, additional references, Nature Research reporting sum-
maries, source data, extended data, supplementary information,
acknowledgements, peer review information; details of author contri-
butions and competing interests; and statements of data and code avail-
ability are available at https://doi.org/10.1038/s41586-021-03969-3.
1. Picelli, S. etal. Full-length RNA-seq from single cells using Smart-seq2. Nat. Protoc. 9,
171–181 (2014).
2. Chen, K. H., Boettiger, A. N., Mofitt, J. R., Wang, S. & Zhuang, X. Spatially resolved, highly
multiplexed RNA proiling in single cells. Science 348, aaa6090 (2015).
3. Zheng, G. X. Y. etal. Massively parallel digital transcriptional proiling of single cells.
Nat. Commun. 8, 14049 (2017).
4. Weyn-Vanhentenryck, S. M. etal. Precise temporal regulation of alternative splicing
during neural development. Nat. Commun. 9, 2189 (2018).
5. Walker, R. L. etal. Genetic control of gene expression and splicing in the developing
human brain. Preprint at https://doi.org/10.1101/471193 (2018).
6. Porter, R. S., Jaamour, F. & Iwase, S. Neuron-speciic alternative splicing of transcriptional
machineries: implications for neurodevelopmental disorders. Mol. Cell. Neurosci. 87,
35–45 (2018).
7. Lukacsovich, D. etal. Single-cell RNA-seq reveals developmental origins and ontogenetic
stability of neurexin alternative splicing proiles. Cell Rep. 27, 3752-3759.e4 (2019).
8. Song, Y. etal. Single-cell alternative splicing analysis with expedition reveals splicing
dynamics during neuron differentiation. Mol. Cell 67, 148-161.e5 (2017).
9. Que, L., Winterer, J. & Földy, C. Deep survey of GABAergic interneurons: emerging
insights from gene-isoform transcriptomics. Front. Mol. Neurosci. 12, 115 (2019).
10. Huang, C.-C., Lin, Y.-S., Lee, C.-C. & Hsu, K.-S. Cell type-speciic expression of eps8 in the
mouse hippocampus. BMC Neurosci. 15, 26 (2014).
11. Zhang, Y. etal. An RNA-sequencing transcriptome and splicing database of glia, neurons,
and vascular cells of the cerebral cortex. J. Neurosci. 34, 11929–11947 (2014).
12. Sugino, K. etal. Mapping the transcriptional diversity of genetically and anatomically
deined cell populations in the mouse brain. eLife 8, e38619 (2019).
13. Bittar, P. G., Charnay, Y., Pellerin, L., Bouras, C. & Magistretti, P. J. Selective distribution of
lactate dehydrogenase isoenzymes in neurons and astrocytes of human brain. J. Cereb.
Blood Flow Metab. 16, 1079–1089 (1996).
14. Tasic, B. etal. Adult mouse cortical cell taxonomy revealed by single cell transcriptomics.
Nat. Neurosci. 19, 335–346 (2016).
15. Gandal, M. J. etal. Transcriptome-wide isoform-level dysregulation in ASD, schizophrenia,
and bipolar disorder. Science 362, eaat8127 (2018).
16. Petri, S. etal. The mRNA expression of AMPA type glutamate receptors in the primary
motor cortex of patients with amyotrophic lateral sclerosis: an insitu hybridization study.
Neurosci. Lett. 360, 170–174 (2004).
17. Macosko, E. Z. etal. Highly parallel genome-wide expression proiling of individual cells
using nanoliter droplets. Cell 161, 1202–1214 (2015).
18. Klein, A. M. etal. Droplet barcoding for single-cell transcriptomics applied to embryonic
stem cells. Cell 161, 1187–1201 (2015).
19. Ramsköld, D. etal. Full-length mRNA-seq from single-cell levels of RNA and individual
circulating tumor cells. Nat. Biotechnol. 30, 777–782 (2012).
20. Arzalluz-Luque, Á. & Conesa, A. Single-cell RNAseq for the study of isoforms-how is that
possible? Genome Biol. 19, 110 (2018).
21. Seirup, M. etal. Reproducibility across single-cell RNA-seq protocols for spatial ordering
analysis. PLoS ONE 15, e0239711 (2020).
22. Yao, Z. etal. A transcriptomic and epigenomic cell atlas of the mouse primary motor
cortex. Nature https://doi.org/10.1038/s41586-021-03500-8 (2021).
23. Wang, X., He, Y., Zhang, Q., Ren, X. & Zhang, Z. Direct comparative analyses of 10x
Genomics Chromium and Smart-seq2. Genomics Proteomics Bioinfomatics https://doi.org/
10.1016/j.gpb.2020.02.005 (2021).
24. Bray, N. L., Pimentel, H., Melsted, P. & Pachter, L. Near-optimal probabilistic RNA-seq
quantiication. Nat. Biotechnol. 34, 525–527 (2016).
25. Zhang, M. etal. Spatially resolved cell atlas of the mouse primary motor cortex by
MERFISH. Nature https://doi.org/10.1038/s41586-021-03705-x (2021).
26. Maaten, L. V. D. & Hinton, G. Visualizing data using t-SNE. J. Mach. Learn. Res. 9,
2579−2605 (2008).
27. Becht, E. etal. Dimensionality reduction for visualizing single-cell data using UMAP.
Nat. Biotechnol. 37, 38–44 (2018).
28. Oliver, P. L. etal. Oxr1 is essential for protection against oxidative stress-induced
neurodegeneration. PLoS Genet. 7, e1002338 (2011).
29. Volkert, M. R. Preventing neurodegeneration by controlling oxidative stress: the role of
OXR1. Front. Neurosci. 14, 611904 (2020).
30. Wu, Y., Davies, K. E. & Oliver, P. L. The antioxidant protein Oxr1 inluences aspects of
mitochondrial morphology. Free Radic. Biol. Med. 95, 255–267 (2016).
31. Tasic, B. etal. Shared and distinct transcriptomic cell types across neocortical areas.
Nature 563, 72–78 (2018).
32. Ntranos, V., Yi, L., Melsted, P. & Pachter, L. A discriminative learning approach to
differential expression analysis for single-cell RNA-seq. Nat. Methods 16, 163–166
(2019).
33. Sarantopoulou, D. etal. Comparative evaluation of full-length isoform quantiication from
RNA-seq. BMC Bioinformatics 22, 266 (2021).
34. Bark, I. C., Hahn, K. M., Ryabinin, A. E. & Wilson, M. C. Differential expression of SNAP-25
protein isoforms during divergent vesicle fusion events of neural development. Proc. Natl
Acad. Sci. USA 92, 1510–1514 (1995).
35. Irfan, M. etal. SNAP-25 isoforms differentially regulate synaptic transmission and
long-term synaptic plasticity at central synapses. Sci. Rep. 9, 6403 (2019).
36. Wallén-Mackenzie, Å. etal. Spatio-molecular domains identiied in the mouse
subthalamic nucleus and neighboring glutamatergic and GABAergic brain structures.
Commun. Biol. 3, 338 (2020).
37. Kim, D.-W. etal. Multimodal analysis of cell types in a hypothalamic node controlling
social behavior. Cell 179, 713-728.e17 (2019).
38. Böckers, T. M. etal. Differential expression and dendritic transcript localization of Shank
family members: identiication of a dendritic targeting element in the 3 untranslated
region of Shank1 mRNA. Mol. Cell. Neurosci. 26, 182–190 (2004).
39. Berkel, S. etal. Sex hormones regulate SHANK expression. Front. Mol. Neurosci. 11, 337
(2018).
40. Zhang, Y., Thompson, R., Zhang, H. & Xu, H. APP processing in Alzheimer’s disease.
Mol. Brain 4, 3 (2011).
41. Rash, B. G. & Grove, E. A. Area and layer patterning in the developing cerebral cortex.
Curr. Opin. Neurobiol. 16, 25–34 (2006).
42. Sansom, S. N. & Livesey, F. J. Gradients in the brain: the control of the development of form
and function in the cerebral cortex. Cold Spring Harb. Perspect. Biol. 1, a002519 (2009).
43. Trapnell, C. etal. Transcript assembly and quantiication by RNA-seq reveals unannotated
transcripts and isoform switching during cell differentiation. Nat. Biotechnol. 28, 511–515
(2010).
44. Yi, L. etal. Gene-level differential analysis at transcript-level resolution. Genome Biol. 19,
53 (2018).
45. Raj, T. etal. Integrative transcriptome analyses of the aging brain implicate altered
splicing in Alzheimer’s disease susceptibility. Nat. Genet. 50, 1584–1592 (2018).
46. Mills, J. D. etal. Unique transcriptome patterns of the white and grey matter corroborate
structural and functional heterogeneity in the human frontal lobe. PLoS ONE 8, e78480
(2013).
47. Klausberger, T. & Somogyi, P. Neuronal diversity and temporal dynamics: the unity of
hippocampal circuit operations. Science 321, 53–57 (2008).
48. Harris, K. D. etal. Classes and continua of hippocampal CA1 inhibitory neurons revealed
by single-cell transcriptomics. PLoS Biol. 16, e2006387 (2018).
49. Thomas, J. D. etal. RNA isoform screens uncover the essentiality and tumor-suppressor
activity of ultraconserved poison exons. Nat. Genet. 52, 84–94 (2020).
50. Karlsson, K. & Linnarsson, S. Single-cell mRNA isoform diversity in the mouse brain.
BMC Genomics 18, 126 (2017).
51. Beerenwinkel, N., Pachter, L. & Sturmfels, B. Epistasis and shapes of itness landscapes.
Stat. Sinica 17, 1317–1342 (2007).
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional afiliations.
Open Access This article is licensed under a Creative Commons Attribution
4.0 International License, which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license,
and indicate if changes were made. The images or other third party material in this article are
included in the article’s Creative Commons license, unless indicated otherwise in a credit line
to the material. If material is not included in the article’s Creative Commons license and your
intended use is not permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a copy of this license,
visit http://creativecommons.org/licenses/by/4.0/.
© The Author(s) 2021

Figures
Citations
More filters
Journal ArticleDOI

What is a cell type and how to define it?

Hongkui Zeng
- 01 Jul 2022 - 
TL;DR: An overview of the basic principles of cell types rooted in evolution and development and approaches to characterize and classify cell types and investigate how they contribute to the organism's function, using the mammalian brain as a primary example as mentioned in this paper .
Journal ArticleDOI

RNA velocity unraveled

TL;DR: It is argued for a more rigorous approach to RNA velocity, and a framework for Markovian analysis that points to directions for improvement and mitigation of current problems is presented.
Posted ContentDOI

Depth normalization for single-cell genomics count data

TL;DR: The results lead the authors to recommend proportional fitting prior to log transformation followed by an additional proportional fitting in single-cell genomics analysis.
Journal ArticleDOI

A single-cell guide to retinal development: Cell fate decisions of multipotent retinal progenitors in scRNA-seq

TL;DR: Recently, single-cell RNA sequencing (scRNA-seq) has been used to quantify gene expression in retinal cells across development, in multiple species, and from numerous important models of human disease as mentioned in this paper.
References
More filters
Journal Article

Visualizing Data using t-SNE

TL;DR: A new technique called t-SNE that visualizes high-dimensional data by giving each datapoint a location in a two or three-dimensional map, a variation of Stochastic Neighbor Embedding that is much easier to optimize, and produces significantly better visualizations by reducing the tendency to crowd points together in the center of the map.
Journal ArticleDOI

Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation

TL;DR: The results suggest that Cufflinks can illuminate the substantial regulatory flexibility and complexity in even this well-studied model of muscle development and that it can improve transcriptome-based genome annotation.
Journal ArticleDOI

Near-optimal probabilistic RNA-seq quantification

TL;DR: Kallisto pseudoaligns reads to a reference, producing a list of transcripts that are compatible with each read while avoiding alignment of individual bases, which removes a major computational bottleneck in RNA-seq analysis.
Journal ArticleDOI

A robust and high-throughput Cre reporting and characterization system for the whole mouse brain

TL;DR: A set of Cre reporter mice with strong, ubiquitous expression of fluorescent proteins of different spectra is generated and enables direct visualization of fine dendritic structures and axonal projections of the labeled neurons, which is useful in mapping neuronal circuitry, imaging and tracking specific cell populations in vivo.
Journal ArticleDOI

Genome-wide atlas of gene expression in the adult mouse brain

Ed S. Lein, +109 more
- 11 Jan 2007 - 
TL;DR: An anatomically comprehensive digital atlas containing the expression patterns of ∼20,000 genes in the adult mouse brain is described, providing an open, primary data resource for a wide variety of further studies concerning brain organization and function.
Related Papers (5)
Frequently Asked Questions (13)
Q1. What contributions have the authors mentioned in the paper "Isoform cell-type specificity in the mouse primary motor cortex" ?

Additionally, the authors show that isoform specificity helps to refine cell types, and that a multi-platform analysis of single-cell transcriptomic data leveraging multiple measurements provides a comprehensive atlas of transcription in the mouse primary motor cortex that improves on the possibilities offered by any single technology. 

TheSMART-seq v4 (SSv4) Ultra Low Input RNA Kit for Sequencing (Takara 634894) was used to reverse transcribe poly(A) RNA and amplify full-length cDNA. 

Highly variable isoforms and genes were identified by first computing the dispersion for each feature, and then binning all of the features into 20 bins. 

SMART-seq19 is an scRNA-seq method that produces full-length reads, enabling the quantification of individual isoforms of genes with the expectation-maximization algorithm20. 

To visualize the SMART-seq data with predefined cluster labels produced via a joint analysis with many other data types the authors performed NCA59 on the full scaled log(TPM + 1) matrix using the subcluster labels, to ten components. 

Unique transcriptome patterns of the white and grey matter corroborate structural and functional heterogeneity in the human frontal lobe. 

While MERFISH probes only measure abundance of Pvalb at the gene level (Fig. 4c), extrapolation from the SMART-seq quantifications can be used to refine the MERFISH result to reveal the spatial expression pattern of the Pvalb-201 isoform. 

The authors analysed 6,160 mouse MOp cells assayed with SMART-seq, 280,327 cells assayed with MERFISH, and 94,162 cells assayed with 10x Genomics Chromium v3. 

These cases are likely to be instances where isoform shifts between cell types are a result of differential splicing—that is, the result of a post-transcriptional program. 

The GTF and the GRCm38 genome fasta file (https://github.com/ pachterlab/BYVSTZP_2020/releases/tag/biorxiv_v1), provided by the BICCN consortium, were used to create a transcriptome fasta file, transcripts-to-genes map, and kallisto index using kb ref -i index.idx, -g t2g.txt -f1 transcriptome.fa genome.fa genes.gtf. 

Without being able to rule out that the low correlation for L5 IT cells across the technologies was due to confounding between batch and sex in the dataset, the authors decided to excluded the subclass from their analyses. 

detection of such genes and their associated isoforms requires meaningful cell-type assignments and accurate isoform quantifications. 

The authors identified 1,971 isoforms from 128 groups of TSSs where the TSSs are preferentially expressed in either GABAergic, glutamatergic or non-neuronal classes, even when the expression of isoforms contained within the TSS is constant (Supplementary Table 10a, c, d).