2022 08 27 505530v1 Full

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 61

bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022.

The copyright holder for this preprint


(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Cell Atlas at Single-Nuclei Resolution of the Adult Human Adrenal Gland and

Adrenocortical Adenomas

Barbara Altieri1,8, A. Kerim Secener2,3,4,8, Somesh Sai2,3,4, Cornelius Fischer2,3, Silviu Sbiera1,

Panagiota Arampatzi5, Sabine Herterich6, Laura-Sophie Landwehr1, Sarah N. Vitcetz2,3, Caroline

Braeuning2, Martin Fassnacht1,6,9*, Cristina L. Ronchi1,7,9*, Sascha Sauer2,3,5,9*

1
Division of Endocrinology and Diabetes, Department of Internal Medicine I, University

Hospital, University of Würzburg, Würzburg (Germany).


2
Max Delbrück Center for Molecular Medicine, Berlin (Germany).
3
Berlin Institute of Health, Berlin (Germany).
4
Institute of Biochemistry, Department of Biology, Chemistry and Pharmacy, Free University

Berlin (Germany).
5
Core Unit SysMed, University of Würzburg, Würzburg (Germany).
6
Central Laboratory University Hospital Würzburg, Würzburg (Germany).
7
Institute of Metabolism and System Research, University of Birmingham, Birmingham (United

Kingdom).
8
These authors contributed equally: Barbara Altieri, Ali Kerim Secener.
9
These authors jointly supervised this work: Martin Fassnacht, Cristina L Ronchi, Sascha Sauer.

*Correspondence to:

Prof. Martin Fassnacht


Division of Endocrinology and Diabetes, Department of Internal Medicine I
University Hospital of Würzburg
Oberduerrbacher Str 6
97080 Würzburg (Germany)
Tel: +49-931-201-39700, Email: fassnacht_M@ukw.de

1
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

ORCID ID: 0000-0001-6170-6398

Dr. Cristina L. Ronchi


Institute of Metabolism and System Research
University of Birmingham
B15 2TT Edgbaston
Birmingham (United Kingdom)
Phone: +44-121-4146857, Email: C.L.Ronchi@bham.ac.uk
ORCID ID: 0000-0001-5020-2071

Dr. Sascha Sauer


Max Delbrück Center for Molecular Medicine
Robert-Rössle-Straße 10
13092 Berlin
Germany
Phone: +49-30-9406-2995, Email: sascha_sauer_2020@gmx.de
ORCID ID: 0000-0002-4794-3693

2
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Abstract

The human adrenal gland is a complex endocrine tissue. Developmental studies on this tissue

have been limited to animal models or human foetus. Here, we present a cell atlas analysis of the

adult human normal adrenal gland, combining single-nuclei RNA sequencing and spatial

transcriptome data to reconstruct adrenal gland development and tumourigenesis. We identified

two populations of potential progenitor cells resident within the adrenal cortex: adrenocortical

progenitors NR2F2+-ID1+ cells, located within and underneath the capsule, and medullary

progenitors SYT1+-CHGA- cells, located in islets in the subcapsular region. Using pseudotime

analyses, we provided evidence of the centripetal nature of adrenocortical cell development and

of the essential role played by the Wnt/β-catenin pathway in the adrenocortical self-renewal. By

comparing transcriptional profiles of cells of normal adrenal glands and adrenocortical adenomas

we revealed a high heterogeneity with six adenoma-specific clusters. Overall, our results give

insights into adrenal plasticity and mechanisms underlying adrenocortical tumourigenesis.

Key words: Adrenal cortex, adrenal medulla, stem cells, progenitor cells, adrenocortical

tumours, tumourigenesis, heterogeneity, cortisol secretion, CTNNB1.

3
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Introduction

The human adrenal gland is a complex endocrine tissue that maintains homeostasis by responding

to various physiological stimuli by secreting steroid hormones and catecholamines. The adult

adrenal gland consists of two functionally distinct parts, the cortex and the medulla, which

develop from different embryological origins1. According to textbooks, the adrenal cortex is

organized in three zones bearing diverse morphological and functional characteristics. From the

outer to innermost layer, histological analyses identify the zonae glomerulosa (ZG), fasciculata

(ZF) and reticularis (ZR) involved in mineralocorticoid, glucocorticoid, and androgen synthesis

and secretion, respectively. Beyond the cortex, the centre of the adrenal gland is occupied by

catecholamine-secreting medulla, originating from neural crest cells.

Several animal model studies have been performed to reveal the architecture of adrenal glands

and their development. According to these studies, the adrenocortical zonation is driven by

several signalling pathways, including the Wnt/β-catenin, the hedgehog (HH), the cAMP/protein

kinase A (PKA), and the insulin-like growth (IGF) pathways2. Among these, both Wnt/β-catenin

and HH play an important role in the adrenal subcapsular area, where they presumably contribute

to the centripetal renovation of the adrenal cortex3. The dysregulation of these signalling

pathways is likely involved in human adrenal tumourigenesis4. However, having recent studies

mostly been performed in mice, these findings cannot be directly transferred to humans.

Benign tumours deriving from the adrenal cortex (adrenocortical adenomas, ACA) are frequent in

the general population (up to 10% in age > 70 years). They are usually incidentally discovered

and mostly endocrine inactive adenomas (EIA)5,6. However, benign tumours may provoke

autonomous steroid secretion associated with clinically relevant disorders, such as adrenal

Cushing’s syndrome in cortisol-producing adenomas (CPA) or Conn's syndrome in aldosterone-

producing adenomas. On the other side, malignant adrenocortical carcinomas (ACC) are rare

aggressive cancers with overall poor prognosis. In the last years, comprehensive genomics studies

of benign and malignant tumours have identified alterations in signalling pathways involved in

4
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

adrenocortical tumourigenesis (i.e.Wnt/β-catenin, Rb/p53 signalling and/or IGF system) and

autonomous cortisol excess (i.e. cAMP/PKA pathway)5-11. However, the molecular mechanisms

underlying the pathogenesis of a vast percentage of adrenocortical tumours remained largely

unexplained. Therefore, a deeper understanding of the normal adrenal growth, differentiation and

self-maintenance of adrenal cells in the adrenal tumourigenesis is urgently required.

Recently, single-cell/single-nuclei RNA-sequencing (sc/snRNA-seq) and spatial transcriptomics

have provided unprecedented insights into cellular heterogeneity allowing the identification of in

part unexpected cellular identities12-14. Particularly, cell atlases of normal human tissues based on

scRNA-seq data may provide the basis for cellular analysis of pathological human samples - such

as tumours - to gain a deeper understanding of tumourigenesis15-17. Whereas some recent scRNA-

seq based studies focused on the neural compartment of the adrenal gland18, to our knowledge so

far only one mouse study investigated on the single-cell transcriptome level the adrenal cortex by

analysing the effect of stress on the hypothalamic-pituitary-adrenal axis 19. However, adrenal

physiology significantly differs between rodents and humans.

Therefore, we aimed to provide a first comprehensive cell atlas of the cortex of the adult human

normal adrenal gland (NAG) at single nuclei resolution including spatial information. This new

approach allowed us to directly find new cell-specific human biomarkers and gain insights in

molecular processes of adrenal self-renewal and homeostasis in humans. Furthermore, we

integrated our atlas with snRNA-seq datasets from a clinical ACA cohort - comprised of EIA and

CPA patient samples - to reveal human tumour-specific cell subpopulations and potential

treatment targets.

Results

Single-nuclei transcriptome sequencing of NAGs

We sequenced 11,931 nuclei from six NAGs (Table 1 and Supplementary Fig. 1), obtaining an

average depth of 25 million reads per sample. In an unsupervised cluster analysis, we identified

5
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

six main cell clusters with distinct gene expression signatures. The identity of each cluster was

assigned by cross-referencing upregulated transcripts with canonical markers from the literature.

The dominant cluster was populated by classical adrenocortical cell types (Fig. 1A), including

subclusters of ZG, ZF and ZR, as well as a new subcluster named “adreno-medullary progenitors”

(AMP) (Fig. 1B). Highly expressed genes typical for these cortex subclusters are shown in Fig.

1C. The remaining “satellite” clusters were annotated as “Medulla”, “Fibroblasts & Connective

tissue” (FC), “Stem/Progenitor and Vascular Endothelial cells” (Stem/P&VEC), “Myeloid cells”

and “Lymphoid cells” (Fig. 1D). The top 100 differentially expressed genes (DEGs) defining the

individual clusters of the NAG are listed in Supplementary Table 1.

Zonation of the adrenal cortex

Within the adrenal cortex cluster, different cell subpopulations showed high expression of genes

encoding for key steroidogenic enzymes representing the three cortex zones (Fig. 1B,

Supplementary Fig. 2A-C, Supplementary Table 1). Cells from the ZG subcluster showed high

expression of CYP11B2, as well as DACH1 and ANO4, all previously described to be highly

selective for this zone20-23. The ZF was characterized by high expression of CYP17A1, CYP11B1,

and HSD3B2, whereas we observed elevated expression of CYB5A, SULT2A1, and GSTA1 in the

ZR24,25, as expected. Further gene enrichment analyses with DEGs from the three adrenocortical

zones showed a significant overlap with regard to our annotation. In fact, aldosterone synthesis

(fold enrichment (FE) >8), aldosterone-regulated sodium reabsorption (FE >6), endocrine and

other factor-regulated calcium reabsorption (FE >4), and Wnt signalling (FE >2) were

upregulated in the ZG subcluster, while cortisol synthesis and cholesterol metabolism were

clearly overrepresented in the ZF (FE >23 and >13, respectively (Supplementary Fig. 3A-B).

Steroid biosynthesis (FE >27), metabolism of xenobiotics by cytochrome P450 (FE >13), and

estrogen signalling (FE >3) were higly represented in the ZR subcluster (Supplementary Fig. 3C).

6
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

To further analyse the transcriptomic zonation of the adrenal cortex, we performed

immunohistochemistry (IHC) to detect protein expression of selected mRNA-based markers in

the different zones of the adrenal glands (represented via the kernel density26 in the “Uniform

Manifold Approximation and Projection”, UMAP; Fig. 2A). Immuno-staining of DACH1 was

uniform in the subcapsular region while CYP11B2 appeared in specific niches (Fig. 2C: DACH1,

CYP11B2). These findings are consistent with the expression of the corresponding mRNAs in our

single nuclei dataset: CYP11B2 featured a low expression density in ZG but DACH1 covered a

substantial portion of this cluster (Fig. 2A: ZG). Moreover, CYP17A1 protein expression was

detected in the entire cortex region (Fig. 2C: CYP17A1), corresponding to a higher RNA

expression density in ZF with lower expression in the adjacent ZG and ZR clusters (Fig. 2A: ZF).

In contrast, CYB5A protein staining clearly distinguished between ZF and ZR (Fig. 2C: CYB5A,

SULT2A1). Although in the snRNA dataset both CYB5A and SULT2A1 densities highlighted

similar regions of the UMAP (Fig. 2: ZR), SULT2A1 was not selectively expressed in ZR as

CYB5a, but was also expressed (at a lower intensity) in the ZG.

Adrenal progenitor and stem cells

Of note, within the main adrenal cortex cluster we identified a subcluster of sympathoadrenal

lineage cells. These cells were characterised by transcriptional similarities to both neuronal cells

and cells of the “Medulla” cluster (for comprehensive initial snRNA-seq analyses we included

not only the cortex but also the medulla) (Fig. 1B, Supplementary Fig. 2F and Supplementary

Table 1). As in the medulla, these cells specifically expressed genes involved in

neurotransmission, such as SYT1, NRG1, NRXN3, GRIK1, and CADPS. Gene enrichment analysis

further corroborated the chromaffin-like nature of these cells, revealing sets of genes involved in

cholinergic and glutamatergic synapse (FE >7), glutamatergic synapse (FE >5), and axon

guidance (FE >4) (Supplementary Fig. 3F). However, cells within this subcluster did not express

markers of the mature medulla such as CHGA and CHGB, TH, PNMT, and DBH. Due to the

7
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

expression of neuronal markers and the absence of hints to the mature medulla, we hypothesize

that this subcluster represents early progenitors of the adrenal medulla, henceforth named as the

“Adreno-Medullary Progenitors” (AMP) subcluster. We validated this hypothesis by using IHC

to investigate the protein expression of SYT1 (synaptotagmin 1) and CHGA (chromogranin) (Fig.

3A-B). We observed low SYT1 expression in the entire cortex, whereas the medulla showed

moderate expression. Interestingly, we found groups of cells with stronger SYT1 staining in the

subcapsular region, which were organized as niches (Fig. 3B.1). In line with snRNA-seq data,

these cells did not show CHGA expression in the double staining SYT1-CHGA (Fig. 3B.2),

confirming that these niches may truly contain progenitor cells of the adrenal medulla.

Interestingly, recent studies in murine models showed that nestin+ medullary progenitor cells

were located not only in the medulla but also in niches of the mouse adrenal capsule, to contribute

to the plasticity of the adrenal cortex. The here found presence of two different progenitor

populations within the adult adrenal cortex might suggest a higher degree of flexibility of tissue

renewal capacity than expected27 to ensure the maintenance of the adult adrenal cortex28-32.

Another cell cluster that includes a significant number of adrenal stem cell markers is the

Stem/P&VEC cluster. Here, classical endothelial genes, such as FLT1 (also known as VEGFR1),

KDR (also known as VEGFR2), and TEK, as well as several genes associated with cell

proliferation and differentiation, including ETS1, ETS2, INSR, and DNASE1L3 were found

(Supplementary Fig. 2E, Supplementary Table 1). Moreover, NOTCH1, a known marker of

adrenal cortical progenitors33 (involved in cell fate specification, differentiation, proliferation, and

survival) and its target gene HES1 (responsible for the regulation of genome-wide glucocorticoid

signalling34) were also both expressed at a high level (Supplementary Fig. 2E). Simultaneously,

low gene expression levels of genes associated to steroidogenesis, i. e. NR5A1 (also known as

SF1) and STAR were found within this cluster. One more interesting finding was the distinct

expression of NR2F2 (also known as COUP-TFII, Fig. 1D), a potential marker of adrenocortical

stem cells in mouse models35,36 and humans37. NR2F2 expression was also observed in the

8
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

"Fibroblasts and connective tissue” cluster (Fig. 3C) that shared a similar transcriptome profile to

the adrenal capsule, expressing COL1A2 and RSPO3 (Supplementary Table 1). Recently, it was

reported that NR2F2 expression is most likely related to RSPO3+ capsular cells and that RSPO3

plays a direct role in the mouse adrenal homeostasis28. In addition to NR2F2, we found ID1, a

known marker of mesenchymal stem cells, exclusively expressed in the same cluster. To gain

more insight in the “stem” identity of this cluster, we performed IHC and microscopy-based

tissue analysis with the undifferentiated protein markers NR2F235,36 and ID1. NR2F2 staining was

found in nuclei of rare, sparsely distributed cells. These cells were mostly localized at the

subcapsular level, whereas few additional NR2F2+ cells were found within the inner cortex,

following the trend of centripetal migration of adrenocortical progenitor cells from ZG to ZF to

ZR (Fig. 3D.1). ID1 nuclei staining showed similar distribution when compared to NR2F2+ cells

(Fig. 3D.1). Via double immunostaining, we found sparse NR2F2+-ID1+ cells mostly localized at

capsular/subcapsular levels that appeared to migrate centripetally within the cortex, suggesting

that ID1 may represent a new marker of stem/progenitor cells (Fig. 3D.2). Interestingly, a

previous study also highlighted the role of ID1 in the tumourigenesis of adrenocortical cancer in

dogs38. Pathway analysis corroborated our hypothesis, revealing significant enrichment of

adherens junction (FE >5) and different signalling pathways, such as mitogen-activated protein

kinase (MAPK, FE >3), Wnt pathway (FE > 2) and signalling regulating the pluripotency of stem

cells (FE >2) (Supplementary Fig. 3E).

Characterisation of additional adrenal cell clusters

As shown in Fig. 1, we detected four other cell types in the adrenal glands. The “Medulla” cluster

showed an expression pattern typical for chromaffin cells with a predominance of CHGA, CHGB,

PNMT, DBH, and TH (Fig. 1D, Supplementary Fig. 2F, Supplementary Table 1). Within the

“Myeloid cells” cluster, we observed a large population of macrophages, mostly polarised into

alternatively activated (M2) macrophages with high expression of MRC1 (also known as CD206)

9
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

and CD163 together with MSR1 (also known as CD204) (Fig. 1D, Supplementary Fig. 2G,

Supplementary Table 1). Other highly expressed genes were SRGN, a proteoglycan that is

typically expressed in myeloid and lymphoid cells, CSF1R, CD86, and CD14 (Supplementary

Fig. 2G). Furthermore, we observed a second immune cluster of lymphoid cells, which showed

typical T cell markers, such as IL7R and CD96 (Fig. 1D, Supplementary Table 1), as well as

CD247 (part of the T-cell receptor-CD3 complex), RUNX3 and ITK (involved in T cell

differentiation), PRF1 and GZMA (indicators of granule mediated cytotoxic activity of natural

killer and cytotoxic T cells) (Supplementary Table 1, Supplementary Fig. 2H). Of note, only few

markers for B cells were found, e. g. IKZF3 (involved in the regulation of B lymphocyte

proliferation and differentiation) (Supplementary Fig. 2H and Supplementary Table 1).

The “Fibroblasts and connective tissue” cluster was characterised by cells that highly expressed

collagen markers, such as COL1A2 and COL4A1, and markers of the extracellular matrix, such as

MGP, FBLN1, LAMA2 and CCN1 (Fig. 1D, Supplementary Fig. 2I and Supplementary Table 1).

Using IHC, we confirmed protein expression of COL1A2 in the adrenal capsule and MGP in

extracellular matrix and fibroblasts (Supplementary Fig. 4A).

Pathway analyses validated the satellite cluster annotations, revealing enrichment of cluster-

specific gene sets, such as glutamatergic and cholinergic synapse (FE >8 and >5, respectively) for

“Medulla”, phagosome and NF-kappa B signalling (FE >4 for both pathways) for “Myeloid

cells”, Th1 and Th2 cell differentiation (FE >5) and B and T cell receptor signalling pathway (FE

>4) for “Lymphoid cells” and focal adhesion (FE >5) and extracellular matrix-receptor

interaction (FE >4) for “Fibroblasts and connective tissue” cluster (Supplementary Fig. 3F-I).

Adrenal cortex zonation by spatial transcriptomics in NAG

We validated the cellular and tissue assigments of genes based on snRNA-seq data by spatial

transcriptomics using the Visium assay in two consecutive adrenal gland sections deriving from

the exemplary patient sample #NAG-7. As we further focused in this study on the adrenal cortex,

10
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

the medulla was not part of the analysed tissue sections. To identify each zone, we transferred the

cell type annotation from the microdroplet-based snRNA-seq into the Visium data set using a

cell-to-spot correlation framework (Fig. 4). Mapping of the adrenal cortex zone markers

confirmed the snRNA-seq dataset. ZG appeared as a thin line of cells towards the outer cortex,

while the distribution of spots with high ZF score was rather ubiquitous. Moreover, spots with

higher ZR scores seemed to locate closer towards the center of the tissue section, although a

substantial overlap between ZF and ZR was observed (Fig. 4 and Supplementary Fig. 5).

Regarding the satellite clusters, the label transfer predicted similar mapping for “Fibroblasts and

connective tissue” and “Lymphoid cells” to the region surrounding the cortex, although the latter

displayed a low overall mapping score, probably due to the low abundance of lymphoid cells

(Fig. 4). Of note, “Fibroblasts and connective tissue” markers such as COL1A2 and RSPO3 were

exclusively expressed in this region, while “Lymphoid cells” markers could not be found

(Supplementary Fig. 5). The mapping of the “Myeloid cells” cluster showed a specific expression

of immune markers mostly around the small blood vessels within the adrenal cortex (Fig. 4).

Finally, the Stem/P&VEC mapped with a high prediction score in the outer cortex, specifically

localized in a subcapsular niche, while a medium to low prediction score was observed in the

inner cortex (Fig. 4).

Pseudotime analysis confirms centripetal development of the adrenal cortex

Next, we aimed to understand the developmental relationship between the observed cell clusters

within the adrenal cortex. To predict a differentiation trajectory, we performed pseudotime

analysis39 using the spatial transcriptomics data. The integration of the two NAG replicates

resulted in 5 distinct Louvain clusters (0 =Capsule, 1 =ZG, 2 =ZF-ZR, 3 =blood vessel, 4

=endothelial) that we annotated based on the results of the label transfer (Fig. 5A). As root node

for pseudotime computation we selected cells expressing RSPO3, a well-established marker of

capsule adrenocortical progenitors of mouse adrenal gland23,28 (within the capsule cluster, RSPO3

11
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

was specifically expressed by a subset of cells (Fig. 5B)). The resulting unsupervised lineage

trajectory of the adrenal cortex started with capsular cells, proceeded through the ZG cluster, and

ended in the ZF-ZR cells, indicating a centripetal development of the adrenal gland.

To further investigate potential changes in global gene expression dynamics along the trajectory,

we applied pseudotime analysis in DEGs of the three main clusters (capsule, ZG, and ZF-ZR)

(Fig. 5C). In total, we found 829 genes that significantly varied across pseudotime. Among them,

RSPO3 was comparatively strongly expressed in capsular, early pseudo-time cells, and weakly

expressed within the inner-zones, late cell types (Fig. 5D). A similar pattern of cellular

development was observed for ID1. INSR gene expression was specifically switched on in ZG

(early time point in development) and its expression diminished throughout the inner cortex (later

time points) (Fig. 5D). The selective activation of these genes at the early stage confirmed the

stem/progenitor nature of the subpopulation of cells within the “Stem/P&VEC cluster” found in

snRNA-seq analysis. Considering the steroidogenic-related genes, CYP11B2 was first activated in

the early ZG cells and significantly decreased throughout the ZG-ZR formation (Fig. 5D). We

further observed an increased gene expression in the late ZG cells - and throughout the

populations of the inner cortex - for CXCR4 and PRKAR1A (cAMP/PKA pathway) and ADGRV1

(cholesterol synthesis pathway) (Supplementary Fig. 6). Of note, the expression of DLK1, which

is a proposed marker of a subcapsular cell population, was observed at two stages of the

trajectory: a first peak appeared around the early development capsular region (Supplementary

Fig 7A-B, indicated by a green box) and a second one at the ZG to ZF-ZR transition zone

(Supplementary Fig. 7A-B, indicated by a blue box). Interestingly, DLK1 expression visualized

on the adrenal sections revealed a similar pattern described by Hadjidemetriou et al.40, i.e as

“clusters of cells” (Supplementary Fig. 7B, regions inside blue boxes).

Moreover, the pseudotime analysis revealed interesting expression of Wnt/β-catenin pathway-

related genes (Supplementary Fig. 6). SFRP1, SFRP2 and SFRP4 were found to be expressed in

early or late capsular cells, but their expression decreased throughout the remaining cells

12
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

subpopulations. WNT4 was expressed in ZG only. Other genes such as CTNNA1 or CTNNA1L

were specifically switched on in the late ZG subpopulation and continued to be mildly expressed

in the ZF-ZR zones (Supplementary Fig. 6).

Integration of NAG and ACA snRNAseq data reveals adenomas-specific cells

Next, we used our human reference atlas to analyse cell type heterogeneity and identify possible

tumourigenesis processes in our ACA patient samples (Table 2). For this, we sequenced 21,794

single-nuclei transcriptomes from 12 snap-frozen ACA samples (7 CPAs and 5 EIAs) and

integrated this data with 5154 highest scoring NAG nuclei.

The differential expression analyses indicated that several genes involved in cell signalling, tissue

remodelling, cell replication and RNA transcription were significantly upregulated in ACAs

compared with NAGs (Fig. 6A.1-2). Particularly genes like IGF2R (average fold-change for gene

expression, avg-log2FC >4.2, p<1e-6 in both EIA and CPA), SLBP (avg-log2FC >3.4, p<1e-6 in

both EIA and CPA), GAS2 (avg-log2FC >3.4, p=1.04e-272 in EIA and avg-log2FC >2.0,

p=8.85e-153 in CPA), and SP100 (avg-log2FC >1.8, p=2.45e-208 in EIA and avg-log2FC >1.6,

p=9.33e-184 in CPA) were strongly expressed in both EIA and CPA (Fig. 6A.3). Genes like

MMP26 (avg-log2FC >3.6, p<1e-6) and ADGRG2 (avg-log2FC > 1.5, p=2.27e-138) were

strongly expressed in EIAs, whereas RBFOX1 (avg-log2FC > 3.1, p<1e-6) and HMGCS1 (avg-

log2FC >1.6, p=4.51e-134) were predominantly expressed in CPAs (Fig. 6A.4).

In line with the UMAP results of NAGs, in ACAs we found a central main cluster and 5 satellite

clusters comprised of myeloid, lymphoid, endothelial, stromal (mainly composed of fibroblasts

and connective tissue) and medullary cells (Fig. 6B). The central major cluster showed a complex

organization including the three cortex-specific ZG, ZF and ZR and seven other subclusters (Fig.

6B-C). The “Cholesterol-and Steroid-Enriched Metabolism” (CSEM) subcluster was also found

in a very small percentage in NAGs (~1%). The six remaining subclusters were specifically

expressed in ACAs and were named as “Adenoma MicroEnvironment” (AME), “ZG-like”, and

13
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

“adenoma-specific clusters” (AC1-4). These subclusters were mostly distributed in the periphery

of the central major cluster, except for AME that occupied a large area in the central region (Fig.

6B). Top 100 DEGs defining the individual clusters of the integration of NAG and ACA are

listed in Supplementary Table 2. In short, RNA markers specific for the ZF and ZR clusters of the

NAG were largely expressed within the central cluster, and RNA markers such as CYP17A1 and

CYP11B1 were rather ubiquitously expressed at low to medium levels across the tumour clusters

(Fig. 6C), confirming the adrenocortical nature of these cells.

ACA-specific clusters distribution according to subtype and mutational status

We further explored the cluster/cell type composition across the 18 samples by grouping our

samples according to tissue entity (CPA, EIA or NAG) and mutational status – available from

previous next-generation DNA or Sanger sequencing of bulk tumour tissues (no driver mutation,

CTNNB1 mutations, and PRKACA/GNAS mutations) (Fig. 6D).

The ZG-like and the AME clusters showed a homogeneous distribution, without significant

differences among the samples (Fig. 6D). Within the “ZG-like” cluster, common ZG markers,

such as DACH1, ANO4, and NR4A2 (also known as NURR1) were dominantly expressed (Fig.

7A), indicating the presence of normal adrenocortical cells in ACAs. However, unlike ZG, the

ZG-like cluster was characterised by lower to null expression of CYP11B2 (avg-log2FC >-2,

p=8.69e-36), CALN1 (avg-log2FC >-2.3, p=9.60e-82) and MC2R (avg-log2FC >-1.9, p=1.17e-

31). Moreover, ZG-like cluster showed also significant expression of non-ZG specific genes, such

as CYP11B1 (avg-log2FC >1.2, p=2.05e-16), IGF2R (avg-log2FC >4.4, p=3.79e-78) and GAS2

(avg-log2FC >2.4, p=.82e-12). Pathway analyses revealed enrichment of several pathways

involved in cell signalling, such as protein phosphorylation (FE >6), as well as pathways involved

in calcium signalling (FE >3) and aldosterone secretion (FE >2) (Fig. 7B and Supplementary Fig.

8).

14
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Cells of the AME overexpressed genes like APOE, ADIPOR2 and RARRES2 involved in the

lipoprotein’s metabolism, as well as IGF1R and NDRG2 (Fig. 6C and 7A). IGF1R is a

transmembrane tyrosine kinase receptor of the IGF family that activates a mitogenic signalling

pathway, stimulating cell proliferation, division, and translation, and inhibiting apoptosis41.

NDRG2 regulates the Wnt signalling pathway by modulating CTNNB1-target genes42. Pathway

analyses revealed enrichment of several pathways involved in cell signalling, such as AMPK (FE

>4) and insulin signalling pathways (FE >3), cholesterol trafficking (FE >4), and pathways

associated with different tight, focal and adherens junctions (FE >2) (Fig. 7B and Supplementary

Fig. 8).

The CSEM cluster was significantly over-represented in the CPA cohort compared to EIAs

(p=0.03; Fig.8A), with high expression of genes related to cholesterol metabolism, including

HMGCS1, SQLE, and INSIG1, but also KCND2and POLE2 (Fig. 7A), which are involved in

DNA repair and replication. The enrichment pathway analyses confirmed this finding, showing

several pathways involved in cholesterol metabolism and steroidogenesis, among which a very

high enrichment of cholesterol and steroid biosynthesis (FE >9), pathways involved in

ubiquitination processes, including protein ubiquitination (FE >5) and cellular response to

oxidative stress (FE >2.5) (Fig. 7B and Supplementary Fig. 8).

The remaining clusters presented a slightly different distribution within the tumour subtypes and

mutational status. Specifically, AC1 was mostly found in two CTNNB1-mutated samples, one

EIA (EIA-1, ~48%) and one CPA (CPA-4, ~46%) (Fig 6D and Fig. 8B), with a high expression

of the β-catenin target genes LEF1, AFF3, and ISM1 (Fig 7A, Supplementary Table 2). Both

AFF3 and ISM1 are known to be overexpressed in benign and malignant adrenocortical tumours

with CTNNB1 mutation43-45. Also, we found a high expression of FTO, a N6-methyladenosine

that promotes cells proliferation, migration and chemo-radiotherapy resistance by targeting

CTNNB1 in different cancer types46,47 (Fig. 6C and Fig. 7A). In addition, other genes involved in

cell proliferation, including GAS2 and TERB1, and long non-coding RNAs, such as CASC16 and

15
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

C9orf84, were found to be highly expressed within this cluster (Fig. 7A). The enrichment analysis

revealed a very high expression of different pathways involved in spliceosome (FE >10), cell

proliferation and tissue development (FE >3), and extracellular cell-matrix adhesion (FE >5) (Fig.

7B and Supplementary Fig.8).

AC2 cluster was more abundant in two EIA samples (EIA-2 and EIA-5) (Fig. 6D) without

correlation with the mutational status (Fig. 8B). The AC2 cluster was characterised by an

overexpression of genes like MMP26, ZNF98 and SP100 (Fig. 6C and 7A) associated with

tumour promotion. Enrichment analysis showed also for AC2 a high expression of spliceosome

pathway (FE >6) and ubiquitin-related pathways, including SUMOylation (FE >5) and cell cycle

(FE >2) (Fig 7B and Supplementary Fig. 8).

In contrast, AC3 was significantly over-represented in CPAs with PRKACA/GNAS mutation

compared to those with CTNNB1-mutated (p=0.02) (Fig. 6D and 8B). Particularly, it was very

abundant in the single CPA with GNAS mutation (CPA-7, ~23%). Of note, we observed a high

expression of PDE8A within this cluster (Fig. 7A). Also, a high expression of SAE1, associated to

SUMOylation, GRB14 and FBLN1, associated with cell growth and extracellular matrix,

respectively, was found (Fig. 6C and 7A). Enrichment analyses confirmed a high enrichment in

genes of the SUMOylation pathway (FE >4), cell adhesion molecular binding (FE >3), but also

spliceosome (FE >2) (Fig 7B and Supplementary Fig. 8). Interestingly, AC1-3 showed a high

enrichment of elements associated to spliceosome, sometimes differently expressed among the

three. As such, splicing pathways were observed at very high levels in AC1 and decreased

through AC2 and AC3, where they were found at lower levels.

Looking into the AC4 cluster, its distribution was similar among all tumour samples except for

one CPA with CTNNB1 mutation (CPA-3), where it represented ~36% of the cells (Fig. 8B).

Within this cluster we found a high expression of the proto-oncogene MYBL2, as well as of

CGGBP1, which regulates cell cycle in cancer cells (Fig. 6C and 7A). Pathway analyses showed

significant enrichment of several pathways involved in extracellular matrix, cell-cell

16
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

communication (FE >2) or different types of cells junctions (FE >2) and oxidative stress induced

senescence (FE >3) (Fig. 7B and Supplementary Fig. 8).

Immune clusters

We observed a decrease in quantity of both lymphoid and myeloid cells from NAGs to ACAs.

Particularly, the lymphoid cluster was predominantly enriched in NAGs compared to EIAs

(p=0.02) and by trend compared to CPAs (p=0.09) (Fig 8C). No difference was observed among

the ACAs according to the mutational status. The expression of UTY significantly decreased from

NAGs to both EIAs (avg-log2FC >-2, p=1.92e-08) and CPAs (avg-log2FC > 1.5, p=5.32e-09)

(Supplementary Fig. 9A). Other genes like HLA-A, HLA-B, HLA-C, and CCL5 followed a trend

to decreased expression in CPA compared to NAG (avg-log2FC >-1.3, p=2.84e-08 for HLA-A,

avg-log2FC >-1.2, p=6.55e-07 for HLA-B, avg-log2FC >-1.7, p=5.58e-07 for HLA-C, avg-

log2FC >-1.5, p=1.05e-06 for CCL5) (Supplementary Fig. 9A).

The myeloid cluster was similarly represented in NAGs when compared to EIAs, but less

represented in CPA samples (p=0.07) (Fig. 8D). Interestingly, the distribution of the myeloid

cluster among the ACA samples changed according to the mutational status. Particularly, ACA

samples with CTNNB1 mutation revealed a higher abundance of myeloid cells compared to

samples with no driver (p=0.03) and PRKACA/GNAS mutations (p=0.08) (Fig. 8D). Among

differentially expressed genes (Supplementary Fig. 9B), MSR1 was significantly downregulated

in EIAs (avg-log2FC >-1.2, p=2.5e-19) and CPAs (avg-log2FC =-0.9, p=4.49e-12) compared to

NAGs. Similarly, SRGN decreased in EIA (avg-log2FC >-1.2, p=5.24e-27) and CPA (avg-

log2FC >-0.6, p=3.57e-11) compared to NAGs. Genes like CSF1R and CD74 were less expressed

in CPAs compared to both NAGs (avg-log2FC >-0.9, p=2.07e-07 for CSF1R and avg-log2FC -

0.9, p=1.87e-10 for CD74) and EIA (avg-log2FC >-1.0, p=3.04e-10 for CSF1R and avg-log2FC

>-1.3, p=4.38e-24 for CD74) (Supplementary Fig. 9B).

17
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Discussion

In this study, we provide a high-resolution, comprehensive transcriptional cell atlas of the normal

adult human adrenal gland including spatial RNA and protein expression information. We could

distinguish characteristic zones of the adrenal cortex (ZG, ZF, ZR), as well as the adrenal medulla

and capsule. In addition, we found different satellite clusters, including lymphoid, myeloid and

vascular cells. Moreover, by comparing our single-cell transcriptome atlas of the adult human

normal adrenal glands to a cohort of adrenocortical adenomas, we revealed seven adenoma-

specific cell types that were involved in adrenocortical tumourigenesis.

One of the key findings of our study was the identification of two different populations of

potential progenitor cells within the adult human NAG: the adrenocortical progenitor-like

NR2F2+-ID1+ cells, located within and underneath the capsule, and the medullary progenitor-like

SYT1+-CHGA- cells, gathering in niches in the subcapsular region of the adrenal cortex. To our

knowledge, we here show for the first time the presence of NR2F2+ progenitor cells in adult

human adrenal glands. Moreover, we also found the co-expression of NR2F2 with ID1 (a

mesenchymal stem cell marker) in few cells that were predominantly located under the capsule.

As observed for progenitor cells in animal models, also in human adrenal glands these

subcapsular cells showed a centripetal migratory pattern. The high differentiation potential of

ID1+ cells is supported by high expression within the same “Stem/P&VEC cluster” of genes

associated with cell proliferation and differentiation (ETS1, INSR, HES1, DNASE1L3,

NOTCH1)33. Moreover, the trajectory analysis of spatial transcriptomic data enabled us to

observe the selective activation of some genes at the early stages of pseudotime, confirming the

progenitor nature of this subpopulation. In addition, the microenviroment of this cluster is rich in

adherens junctions and vascular cells, which are essential in cell-cell communication, cell renewal

and differentiation48,49. All these findings suggest that these ID1+ cells might represent a new

progenitor population in the adult adrenal cortex.

18
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

In addition, we observed a small subcluster of cells, which expressed neuronal-like gene markers,

including SYT1, NRG1, NRXN3, GRIK1 and CADPS as found in chromaffin cells of the adrenal

medulla. The subcluster was void of markers of mature medullary cells such as CHGA. Using

IHC we validated the presence of SYT1+- CHGA- cells that form clusters in the subcapsular

region of the cortex, as had been observed for the nestin+ medullary progenitor cells in mouse.

However, we could not identify nestin itself, potentially due to the differences in gene expression

between mice and humans. Nevertheless, we cannot exclude that the rarity of these progenitors in

adult adrenal glands or their presence only in early-stage adrenal development is responsible for

this fact. The neuronal-like nature of the expressed markers within this subcluster raised the

question if SYT1+- CHGA- cells could represent a progenitor population of the adrenal medulla

that is located in the subcapsular region of the adult human adrenal cortex, as had been observed

in animal models50. Moreover, in line with the undifferentiated nature of NR2F2+-ID1+ and

SYT1+-CHGA- cells, both potential progenitor clusters showed little to no expression (< 5% in

both clusters) of mature cells markers at transcriptional levels, including NR5A1 and STAR in the

Stem/P & VEC cluster, and CHGA, CHGB, TH, PNMT, and DBH in the AMP cluster.

Interestingly, by combining pseudotime and spatial transcriptomic analysis, we corroborated the

centripetal nature of adrenocortical cell development. However, it still remains to be

demonstrated ex vivo whether NR2F2+-ID1+ and SYT1+-CHGA- cells represent populations

capable to undergo steroidogenic or chromaffin differentiation. Future lineage tracing

experiments will enable us to functionally characterise these progenitor subpopulations, this being

beyond the scope of this pilot study.

The trajectory and pseudotime analyses allowed us to look more deeply into the transition zone

between the capsular and sub-capsular region, where the progenitor cells are mostly located. Here

we showed, for the first time, a selective activation of members of the Wnt/β-catenin pathway

(i.e. SFRP2 and SFRP4), which play an important role in adrenocortical renewal and

differentiation51, as well as in adrenocortical tumourigenesis11,52,53. Particularly, both SFRP2 and

19
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

SFRP4 were expressed in human mesenchymal stem cells isolated from bone marrow54, and a

decreased expression of SFRP2 was associated with aldosterone-producing adenoma

development55. These findings corroborated an essential role of the Wnt/β-catenin pathway for

progenitor cell formation and function throughout the adrenal development. Intriguingly, DLK1

expression was observed in cells organised as clusters in the subcapsular region, as previously

described40. Pseudotime analyses revealed that DLK1 expression exhibits two peaks (first during

the capsule-ZG transition and second during ZG to ZF-ZR transition), strengthening its potential

role in adrenocortical remodelling.

Recently, it was proposed that adrenal renewal is sexually dimorphic36,56. In our study, we were

not able to find any significant transcriptomic differences between the two sexes, potentially due

to the small sample size (tissue from two females versus four males) and the limited number of

(few hundreds of) detected expressed genes per nucleus. Similarly, age-specific differences were

not observed, as all samples of this cohort belonged to adult subjects (ranging from 49 to 81 years

of age).

Our transcriptome atlases of healthy adult human adrenal glands will be an important tool as

reference to investigate adrenal-related disease. In a first step, we interrogated benign

adrenocortical tumours (adenomas), which were further classified according to glucocorticoid

secretion and mutational status. Using this approach, a high intratumoural heterogeneity was

observed, confirming for the first time at single-nuclei level previous morphological and genetic

findings in bulk tumours57-59.

The comparative analysis of the transcriptional profiles of NAGs and ACAs revealed the

presence of seven adenoma-specific clusters. First, our results revealed an adenoma

microenvironment (AME) cluster, which was quite homogeneously distributed among the

samples. This represents a lipid-enriched milieu, similar to the tumour microenvironment

described in other solid tumours60. It has been demonstrated that such an enrichment in adipocyte

and lipid metabolism related genes plays a key role in aberrant cell growth and cancer

20
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

progression, as well as in immune escape and treatment-resistance60. Furthermore, as reported for

the tumour microenvironment in colon cancer61, a high expression of RARRES2, a key modulator

of tumour growth and inflammation, was detected within this cluster. This finding is in line with a

previous study reporting a higher RARRES2 expression in benign compared to malignant

adrenocortical tumours62. Within the AME, also a high expression of IGF1R, which plays an

important role in the adrenocortical tumourigenesis, was observed41. Interestingly, an

overepression of IGF1R within the tumour microenvironment was reported to promote tumour

initiation and progression in lung cancer and medulloblastoma63,64.

The “cholesterol- and steroid-enriched metabolism” cluster, which showed a high expression of

cholesterol pathway genes (i. e. HMGCR, MSMO1, INSIG1, HMGCS1), was, as expected, largely

represented in CPAs65. Furthermore, in the single CPA with GNAS mutation (CPA-7), we

observed a high prevalence of AC3, characterised by an overexpression of PDE8A. These results

confirm previous reports that highlighted the potential role of the phosphodiesterase 8 family

members in the pathogenesis of Cushing’s syndrome11,66. On the contrary, the AC2 cluster was

significantly abundant in two out five EIA samples (EIA-2 and EIA-5), where a high enrichment

of post transcriptional modifications (SUMOylation, deubiquitination) was detected.

A high morphological heterogeneity was particularly observed in CTNNB1-mutated adenomas, as

recently reported in hepatocellular carcinomas harbouring mutations in the β-catenin gene67,68.

Based on the transcriptome profile, we observed three types of CTNNB1-mutated ACA: one with

considerable abundance of AC1 (more than 45% of cells, including one EIA and one CPA); one

CPA with abundance of AC4 (more than 35% of cells); and the remaining three EIAs without

preference for a specific adenoma cluster. Of note, in the CPA with abundance of AC1 we have

observed a high expression of AFF3, ISM1 and LEF1, all known CTNNB1 target genes in benign

and malignant adrenocortical tumours43-45. Particularly, it has been demonstrated that AFF3

mediates the oncogenic effects of β-catenin in ACC cell line NCI-H295R by acting on

transcription and RNA splicing44. Similarly, in the AC4 abundant CPA, a high expression of

21
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

MYBL2 was found. It has been shown that MYBL2 is a Wnt/β‐catenin target gene in different

types of cancer-xenograft models69. Furthermore, MYBL2 regulates cell proliferation, survival and

differentiation and its overexpression is associated with poor outcome in different cancers types70,

including ACC71. These findings might support the previously reported hypothesis of a potential

adenoma-carcinoma sequence at least in a subgroup of CTNNB1-mutated adrenocortical

tumours11.

Another notable finding emerging from the transcriptional profiles of adenomas was the

enrichment of elements associated with spliceosome observed within three ACs (AC1-3),

suggesting a potential role of spliceosome pathway in adrenocortical tumourigenesis, as

demonstrated in other benign and malignant tumours72-74.

Interestingly, a different distribution of immune clusters emerged among tissue entities. Both

lymphoid and myeloid clusters were better represented in normal adrenals than in adenomas.

Although this trend was more evident for lymphoid cells, the median number of myeloid cells

showed a tendency to decrease in CPA compared to EIA and NAG. This finding is in line with

the known interplay between glucocorticoids and tumour-immune infiltration observed in

malignant adrenocortical tumours75, suggesting the inhibition of tumour-infiltrating immune cells

also in CPAs. In addition, we demonstrated for the first time a high prevalence of M2-like

macrophages within the myeloid cluster in normal adrenal. M2-like macrophages have anti-

inflammatory properties and promote cell homeostasis, confirming the potential trophic function

of adrenal macrophages76. This shifted macrophage polarisation towards M2-like phenotype and,

in particular, towards M2c subgroup, characterised by a high expression of the scavenger

receptors MRC1 and CD16377, could be related to the autocrine action of glucocorticoids secreted

by the adrenal gland. To note, in adenomas, the mutational status significantly affected the

myeloid cluster, which was highly enriched in CTNNB1-mutated adenomas. As observed in

NAG, adenoma-associated macrophages presented a prominent M2-like polarisation,

corroborating previous studies reporting that Wnt/β-catenin signalling promotes M2-like

22
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

macrophage in tumours78 and that the M2 signature is highly prevalent in CTNNB1-mutated

samples79.

Short summary and outlook

In conclusion, using snRNA-seq and spatial RNA and protein expression analyses of adult human

adrenal glands, we provide a first comprehensive cell atlas to identify complex mechanisms of

homeostasis and self-renewal of the normal adrenal gland. Moreover, the present reference atlas

allowed us already to analyse molecular processes of adrenocortical tumourigenesis and

autonomous cortisol secretion in adenomas. In the future, further improved snRNA-seq

technologies that will generate significantly more sequences per nucleus and provide data from

more nuclei per analysis will help to ameliorate transcriptional data analysis, for example to

deeper characterise developmental processes of the two newly identified progenitor cell types

(SYT1+-CHGA-- and NR2F2+-ID1+-cells) in NAGs. Furthermore, ex vivo experiments will also

be needed to analyse in more detail the influence of markers such as DLK1 and ID1 on cellular

differentiation and specification.

Materials and Methods

Tissue sample selection

For single-nuclei RNA-seq (snRNA-seq), nuclei were isolated from six snap-frozen samples

collected from NAGs deriving from the tissue surrounding EIA or from adrenalectomies

performed during surgery for renal cell carcinoma (Supplementary Fig. 1). Demographic data and

details of detected cellular transcriptomes are summarized in Table 1. Additionally, single nuclei

were isolated from 12 fresh-frozen ACA samples, including 7 CPAs and 5 EIAs. Among these

samples, 4 tissues were included in two previous studies on whole-exome sequencing10 or bulk

RNA-sequencing11. Clinical characteristics, hormonal secretion, and detected cellular

transcriptomes are summarized in Table 2 and Supplementary Table 3. For validation by

23
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

immunohistochemistry (see below), 16 formalin-fixed paraffin-embedded (FFPE) tissue sections

of NAGs were analysed (Supplementary Table 4). FFPE tissues included 3 out of 6 NAGs

evaluated by snRNA-seq and one sample used for the Visium analysis.

Tissue samples were selected from the Würzburg Adrenal Biomaterial Archive, part of the

BMBF-funded Interdisciplinary Bank of Biomaterials and Data Würzburg (IBDW) applying the

highest standards for biobanking80 (details in Supplementary Material).

The study was approved by the ethics committee of the University of Würzburg (No. 93/02 and

88/11) and written informed consent was obtained from all subjects.

Sequencing of driver mutations in adenoma tissues

Known driver hot-spot mutations in CTNNB1 (exon 3), PRKACA (p.Leu206) and GNAS

(p.Arg201 and p.Gln227)10,11 were evaluated by Sanger sequencing in bulk tumour tissues (n=8).

Four samples were analysed by Whole Exome Sequencing or RNA-seq as part of previous studies

from our group10,11. DNA was isolated from snap-frozen ACA tissues with the Maxwell® 16

Tissue DNA Purification Kit (#AS1030, Promega, Madison, WI, USA) according to the

manufacturer’s instructions. The mutations were genotyped by Polymerase Chain Reaction

(PCR), as previously reported81 (details in Supplementary Material).

Clinical data collection

For ACAs, the endocrine workup was conducted according to the adrenal incidentalomas

guidelines issued by the European Society of Endocrinology/European Network for the Study of

Adrenal Tumors5 and the Endocrine Society guidelines for the diagnosis of Cushing’s

syndrome82. Hormone levels were measured using commercially available analytical procedures

(details in Supplementary Material), as previously reported83. CPA derived from patients with

overt Cushing's syndrome and unequivocal laboratory results (for details see Supplementary

Material).

24
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Single-nuclei RNA sequencing (snRNA-Seq)

Single-nuclei isolation: We applied a previously published protocol for nuclei isolation from

snap-frozen tissue84 (Supplementary Fig. 1 and Supplementary Material). The nuclei integrity and

purity were checked under a microscope and the yield was quantified using a Neubauer chamber.

Single-nuclei RNA sequencing: We used the inDropTM method consisting of three major steps:

Silanization, Reverse Transcription and Library preparation (1CellBioTM, USA) following

manufacturer’s protocols. First, chips were silanized, primed with the inDrop hydrogel beads,

droplet making oil, RT/lysis mix and the nuclei. After completing the run protocol, cell

encapsulations were collected and exposed to UV light to release photocleavable primers from

hydrogel beads. Subsequently, reverse transcription was performed, and the beads filtered out.

From there on, the library preparation step followed the CEL-Seq 2 protocol85: the RT product

was digested by ExoI and HinFI and purified using AMpure XP beads (Beckman CoulterTM,

USA). Then, the second strand cDNA synthesis and in vitro transcription (IVT) steps were

carried out. The resulting RNA was subjected to fragmentation and then reverse-transcribed using

random hexamers. The cDNA was ultimately purified using AMpure XP beads and quantified by

qPCR assay86 (Roche Light Cycler 480 Instrument II, Switzerland).

After quantification, the 6 normal adrenal gland (NAG), 2 endocrine inactive adenoma (EIA-3,

EIA-4) and 2 cortisol producing adenoma (CPA-1, CPA-2) libraries were pooled together and

sequenced using an Illumina HiSeq 2000 instrument with 60 bases for read (R1), 6 for the

Illumina index and 50 for the read 2 (R2). The 6 samples were integrated using the standard

Seurat (version 3.0) integration pipeline87 (freely available software): The resulting data is shown

in Supplementary Fig. 1C. Cells from the two subtypes showed a similar clustering pattern,

suggesting that the data were well integrated. Of note, data derived from either NAG-EIA or

NAG-RCC patients showed correlation with a high Pearson correlation score of 0.99

(Supplementary Fig. 1D), indicating that NAGs could be merged in one single group.

25
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Moreover, the remaining 8 adrenocortical adenoma (ACA) libraries were pooled together and

sequenced on two NovaSeq S1 flow cell at 2 x 100 configuration.

Generation of count matrices and core processing: The resulting base calls (BCL) files were

converted to FASTQ files using bcl2fastq tool (Illumina). These raw sequencing reads were

processed with the zUMI88 (version 2.5 – default parameters) pipeline. The count matrices from

reads spanning both introns and exons (inex) were used for subsequent analysis. Prior to pre-

processing, the genes/features were converted from Ensembl (ENSG) IDs to gene symbols using

a custom script. The same cut-off was used to filter nuclei and genes likewise from individual

samples (min.features= 200 and min.cells =3, CreateSeuratObject). The fraction of reads

mapping to the mitochondrial chromosome (MT-) was calculated (PercentageFeatureSet).

Data normalization and integration using Seurat: The filtered data was log-normalized

(NormalizeData) and 2000 highly variable genes (HVG) were selected. The six samples were

then integrated (1. FindIntegrationAnchors, 2. IntegratedData)89. The integrated data was then

scaled (ScaleData) and the effect of mitochondrial genes was regressed out.

Dimensionality Reduction and Clustering: Thirty principal components (PCs) were computed

(RunPCA) and first 15 were used to generate the UMAP reduction (RunUMAP). Subsequently,

the k-nearest neighbours (knn) were determined (FindNeighbors) and clustering was performed

(resolution=0.01, FindClusters). The resulting clusters were annotated using known marker

genes and differential gene expression (DGE – FindAllMarkers) analysis.

Module Scoring: A list of top DEGs and prior knowledge was curated for each cluster, and

combined gene expression scores were calculated (AddModuleScore). The Seurat object was

further filtered using these scores, with the following thresholds: ZG >0.3, ZF >0.8, ZR >0.5, MC

>0, LC >0, FC >0, Stem/P&VEC >0, AM >0, AMP >0. This resulted in a count matrix of 5154

nuclei that was then integrated with the ACA dataset. A Shiny implementation (R) was set up to

easily navigate within this single-nuclei atlas.

26
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Gene enrichment analysis: We used pathfindR90 to perform the gene enrichment analysis using

the FindAllMarkers output table from Seurat. The analysis was performed using the function

run_pathfindR for both NAG and ACA samples. For the NAGs, KEGG pathways was used as

reference. To better cover tumour heterogeneity in ACAs, three different databases (KEGG,

Reactome and GO-All) were used as reference. For each ACA cluster, similar terms across the

three databases were grouped together under broader categories (Splicing, Extracellular matrix,

Cell proliferation and tissue development, Ubiquitin related, Senescence, Cell Cycle, Receptor

tyrosine kinases, Cell signalling, Cholesterol trafficking, ABC transporters, Calcium signalling,

Aldosterone synthesis, Steroidogenesis, Cholesterol metabolism, and Other) based on shared

genes.

Visium (10x Genomics) spatial gene expression assay

Frozen adrenal gland samples (NAGs) were embedded in OCT (TissueTek) and cryo-sectioned

(Thermo Scientific, CryoStar NX50) with a 10 µm thickness. The sections were then placed on

tissue optimization and gene expression slides. 20 minutes was determined as the optimal

permeabilisation temperature. Bright field H&E images were taken with a 10X objective (Leica

DMi8 – A). After recovery of cDNA from the slide, the libraries were prepared following the

gene expression user guide (10x Genomics – CG000239 Rev A). Subsequent to pooling, libraries

were loaded at 200 pM and sequenced on a HiSeq-4000 (Illumina).

Generation of Visium count matrices. The base call (BCL) files generated from the raw

sequencing data were converted to FASTQ reads using bcl2fastq tool (Illumina). The FASTQ

reads were mapped to human reference dataset GRCh38 (build 2020-A; refdata-gex-GRCh38-

2020-A) using Space Ranger v1.1.0 count pipeline. On average the two slides captured each 1269

spots under the tissue, and both yielded 1973 median genes and 225,346 mean reads per spot.

Seurat Processing of Visium count data. The two Visium Seurat objects were created using

default parameters (Load10X_Spatial), and were subsequently normalised (SCTransform). Thirty

27
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

PCs were computed and first 20 were used to generate the UMAP reduction. Knn of each spot

was then determined (FindNeighbors) and clustering was performed (resolution =0.2,

FindClusters), yielding 6 spatial clusters.

Label transfer from snRNA-seq data onto Visium. Label transferring was performed according

to the label transfer workflow in Seurat. Anchors, or cell-spot pairwise comparisons were

computed (FindTransferAnchors), and subsequently a prediction assay was generated

(TransferData), containing the prediction scores for each cell type within each spot. Spatially

variable features list were generated and included in both objects (assay=“SCT”,

selection.method=“markvariogram”,FindSpatiallyVariableFeatures).

Trajectory analysis

Both Visium objects were integrated with the SCTransform integration framework using default

parameters. The Louvain clustering was performed (resolution=0.2) and those were annotated

with prior knowledge from the label transfer analysis. The SCT adjusted count matrix (“counts”

slot in SCT assay) was extracted from the integrated object, and a “cell data set” object was

generated (new_cell_data_set) within Monocle 339. Subsequently, the cds object was normalized

(preprocess_cds) and seurat computed UMAP coordinates were added into reducedDims slot.

The trajectory was then fit (close_loop =F, use_partition =F, learn_graph) and pseudotime was

computed (order_cells) by setting the root to RSPO3+ cells. After getting the genes that vary as a

function of pseudotime (neighbor_graph="principal_graph", graph_test), they were grouped into

modules (resolution = 0.02, find_gene_modules). Genes from top scoring modules (0 for

Capsule, 1 for ZG and 2 for ZF_ZR) were combined to construct the top gene list that

differentially change across pseudotime.

28
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

ACA-NAG data integration

Count matrix merging and normalization. Count matrices from 12 ACA (37,037 nuclei: 7 CPA

and 5 EIA) and highest scoring 5154 NAG nuclei (See Module Score section above) were

merged. The Seurat object was then generated (min.features=200, min.cells=3,

CreateSeuratObject), cell cycle (S.Score, G2M.Score) and mitochondrial genes were scored

(PercentageFeatureSet) and regressed out during the normalisation step with SCTransform91.

Integration using Harmony: Post normalisation, sample integration was performed within the

Harmony framework (1. RunPCA, 2. RunHarmony with 40 PCs)92. Using the Harmony

dimensions, knns were computed (FindNeighbors) and clustering was done (resolution=0.2,

FindClusters). Finally, a UMAP reduction was performed on the Harmony dimensions

(RunUMAP). Clusters with less than 5 differentially expressed features were removed.

Analysis of DEG among the ACA clusters: Logistic regression test was used within the

FindMarkers framework implemented in Seurat for computing DEGs across subtype (NAG, EIA,

CPA) and clusters. The results were visualised as volcano plots (EnhancedVolcano v1.6.0) and

heatmaps (ComplexHeatmap v2.4.3)

Cluster distribution among tissue samples: The distribution of the clusters among the samples

were evaluated considering the type of the tissue (NAG vs EIA vs CPA) and the mutational status

(no driver vs CTNNB1-mutation vs PRKACA/GNAS-mutation) for clusters selectively expressed

only in ACA. To this aim, the percentage of the number of cells within the cluster per sample was

considered. Data were analysed using the non-parametric Student´s T-tests and Mann-Whitney U

test, as appropriate. A p-value <0.05 was considered statistically significant. The statistical

analysis was performed with GraphPad Prism version 9 (GraphPad Software, San Diego, CA,

USA).

29
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Immunohistochemistry

IHC was used to validate the expression and spatial context of the proteins of selected genes from

the transcriptome analysis. This assay was performed with FFPE material from 16 NAGs, as

previously described93,94. Primary antibodies are summarised in Supplementary Table 5.

Moreover, double immunostaining was applied to validate new markers of adrenocortical

stem/progenitors and adrenomedullary progenitors. The combination of NR2F2/ID1 and

CHGA/SYT1 was used to test FFPE consecutive section from 9 NAGs. Details of the IHC assay

including image acquisition is reported in the Supplementary material.

Data availability

Raw gene expression counts can be found under “Supplementary Data” a the website

accompanying this paper (https://.....). Sequencing data was submitted to the ArrayExpress

database of the European Bioinformatics Institute (EBI); accession number: E-MTAB-12129.

Visium data can be retrieved from here: https://doi.org/10.5281/zenodo.5289292.

Acknowledgements

The authors are grateful to Sonja Steinhauer, Antonia Lorey, and Martina Zink for excellent

technical support. We are also thankful to Celso E. Gomez-Sanchez who provided selective

antibodies detecting human CYP11B2. We thank Thomas Conrad for administrative and

technical support.

This work has been supported by the Deutsche Forschungsgemeinschaft (DFG) (project FA-

466/8-1, RO-5435/3-1 and 405560224 (M.F., C.L.R., and S.S.)) and within the CRC/Transregio

(project number: 314061271 – TRR 205), and the Deutsche Krebshilfe (70113526 M.F. and

C.L.R.). This work has been carried out with the help of the Interdisciplinary Bank of

Biomaterials and Data of the University Hospital of Würzburg and the Julius Maximilian

30
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

University of Würzburg (IBDW), supported by the Federal Ministry for Education and Research

(Grant number FKZ: 01EY1102).

Author contributions

C.L.R., M.F. and S.S. designed and coordinated the study and provided funding. B.A., L.S.L.,

and C.L.R. collected the tissues samples. A.K.S., P.A., S.N.V. and C.B. prepared the samples for

the sequencing and generated sequencing data. A.K.S., So.Sa. and C.F. performed bioinformatics

analysis. A.K.S. performed Visium experiments. B.A. and Si.Sb. performed

immunohistochemistry experiments. S.H. evaluated the mutational status of the included

adenomas. B.A. and A.K.S. performed statistical analysis and generated figures. Data

interrogation and interpretation were carried out by B.A., A.K.S., So.Sa., C.F., Si.Sb., M.F.,

C.L.R., and S.S. B.A., A.K.S., C.L.R. and S.S. wrote the manuscript and Si.Sb., L.S.L. and M.F.

provided critical feedback. All other authors read and approved the manuscript.

31
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

References
1 Xing, Y., Lerario, A. M., Rainey, W. & Hammer, G. D. Development of adrenal
cortex zonation. Endocrinol Metab Clin North Am 44, 243-274 (2015).
https://doi.org:10.1016/j.ecl.2015.02.001
2 Lyraki, R. & Schedl, A. Adrenal cortex renewal in health and disease. Nat Rev
Endocrinol 17, 421-434 (2021). https://doi.org:10.1038/s41574-021-00491-4
3 King, P., Paul, A. & Laufer, E. Shh signaling regulates adrenocortical
development and identifies progenitors of steroidogenic lineages. Proc Natl Acad
Sci U S A 106, 21185-21190 (2009). https://doi.org:10.1073/pnas.0909471106
4 Penny, M. K., Finco, I. & Hammer, G. D. Cell signaling pathways in the adrenal
cortex: Links to stem/progenitor biology and neoplasia. Mol Cell Endocrinol 445,
42-54 (2017). https://doi.org:10.1016/j.mce.2016.12.005
5 Fassnacht, M. et al. Management of adrenal incidentalomas: European Society of
Endocrinology Clinical Practice Guideline in collaboration with the European
Network for the Study of Adrenal Tumors. Eur J Endocrinol 175, G1-G34
(2016). https://doi.org:10.1530/EJE-16-0467
6 Sherlock, M. et al. Adrenal Incidentaloma. Endocr Rev 41 (2020).
https://doi.org:10.1210/endrev/bnaa008
7 Prete, A. et al. Cardiometabolic Disease Burden and Steroid Excretion in Benign
Adrenal Tumors : A Cross-Sectional Multicenter Study. Ann Intern Med 175,
325-334 (2022). https://doi.org:10.7326/m21-1737
8 Deutschbein, T. et al. Age-dependent and sex-dependent disparity in mortality in
patients with adrenal incidentalomas and autonomous cortisol secretion: an
international, retrospective, cohort study. The Lancet Diabetes & Endocrinology
10, 499-508 (2022). https://doi.org:https://doi.org/10.1016/S2213-8587(22)00100-
0
9 Beuschlein, F. et al. Constitutive activation of PKA catalytic subunit in adrenal
Cushing's syndrome. N Engl J Med 370, 1019-1028 (2014).
https://doi.org:10.1056/NEJMoa1310359
10 Ronchi, C. L. et al. Genetic Landscape of Sporadic Unilateral Adrenocortical
Adenomas Without PRKACA p.Leu206Arg Mutation. J Clin Endocrinol Metab
101, 3526-3538 (2016). https://doi.org:10.1210/jc.2016-1586
11 Di Dalmazi, G. et al. RNA Sequencing and Somatic Mutation Status of
Adrenocortical Tumors: Novel Pathogenetic Insights. J Clin Endocrinol Metab
105 (2020). https://doi.org:10.1210/clinem/dgaa616
12 Praktiknjo, S. D. et al. Tracing tumorigenesis in a solid tumor model at single-cell
resolution. Nature Communications 11, 991 (2020).
https://doi.org:10.1038/s41467-020-14777-0
13 Baccin, C. et al. Combined single-cell and spatial transcriptomics reveal the
molecular, cellular and spatial bone marrow niche organization. Nature Cell
Biology 22, 38-48 (2020). https://doi.org:10.1038/s41556-019-0439-6
14 Jones, R. C. et al. The Tabula Sapiens: A multiple-organ, single-cell
transcriptomic atlas of humans. Science 376, eabl4896 (2022).
https://doi.org:doi:10.1126/science.abl4896
15 Regev, A. et al. The Human Cell Atlas. Elife 6 (2017).
https://doi.org:10.7554/eLife.27041

32
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

16 Aran, D. et al. Comprehensive analysis of normal adjacent to tumor


transcriptomes. Nature Communications 8, 1077 (2017).
https://doi.org:10.1038/s41467-017-01027-z
17 Young, M. D. et al. Single-cell transcriptomes from human kidneys reveal the
cellular identity of renal tumors. Science 361, 594-599 (2018).
https://doi.org:doi:10.1126/science.aat1699
18 Bedoya-Reina, O. C. et al. Single-nuclei transcriptomes from human adrenal
gland reveal distinct cellular identities of low and high-risk neuroblastoma
tumors. Nature Communications 12, 5309 (2021). https://doi.org:10.1038/s41467-
021-24870-7
19 Lopez, J. P. et al. Single-cell molecular profiling of all three components of the
HPA axis reveals adrenal ABCB1 as a regulator of stress adaptation. Science
Advances 7, eabe4497 (2021). https://doi.org:doi:10.1126/sciadv.abe4497
20 Zhou, J. et al. DACH1, a zona glomerulosa selective gene in the human adrenal,
activates transforming growth factor-β signaling and suppresses aldosterone
secretion. Hypertension 65, 1103-1110 (2015).
https://doi.org:10.1161/hyp.0000000000000025
21 Maniero, C. et al. ANO4 (Anoctamin 4) Is a Novel Marker of Zona Glomerulosa
That Regulates Stimulated Aldosterone Secretion. Hypertension 74, 1152-1159
(2019). https://doi.org:10.1161/hypertensionaha.119.13287
22 Kobuke, K. et al. Calneuron 1 Increased Ca(2+) in the Endoplasmic Reticulum
and Aldosterone Production in Aldosterone-Producing Adenoma. Hypertension
71, 125-133 (2018). https://doi.org:10.1161/hypertensionaha.117.10205
23 Shaikh, L. H. et al. LGR5 Activates Noncanonical Wnt Signaling and Inhibits
Aldosterone Production in the Human Adrenal. J Clin Endocrinol Metab 100,
E836-844 (2015). https://doi.org:10.1210/jc.2015-1734
24 Rege, J. et al. Transcriptome profiling reveals differentially expressed transcripts
between the human adrenal zona fasciculata and zona reticularis. J Clin
Endocrinol Metab 99, E518-527 (2014). https://doi.org:10.1210/jc.2013-3198
25 Nakamura, Y. et al. 3βHSD and CYB5A double positive adrenocortical cells
during adrenal development/aging. Endocr Res 40, 8-13 (2015).
https://doi.org:10.3109/07435800.2014.895377
26 Alquicira-Hernandez, J. & Powell, J. E. Nebulosa recovers single cell gene
expression signals by kernel density estimation. Bioinformatics (2021).
https://doi.org:10.1093/bioinformatics/btab003
27 Visvader, J. E. & Clevers, H. Tissue-specific designs of stem cell hierarchies. Nat
Cell Biol 18, 349-355 (2016). https://doi.org:10.1038/ncb3332
28 Vidal, V. et al. The adrenal capsule is a signaling center controlling cell renewal
and zonation through Rspo3. Genes Dev 30, 1389-1394 (2016).
https://doi.org:10.1101/gad.277756.116
29 Finco, I., Lerario, A. M. & Hammer, G. D. Sonic Hedgehog and WNT Signaling
Promote Adrenal Gland Regeneration in Male Mice. Endocrinology 159, 579-596
(2018). https://doi.org:10.1210/en.2017-03061
30 Freedman, B. D. et al. Adrenocortical zonation results from lineage conversion of
differentiated zona glomerulosa cells. Dev Cell 26, 666-673 (2013).
https://doi.org:10.1016/j.devcel.2013.07.016

33
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

31 Bandiera, R. et al. WT1 maintains adrenal-gonadal primordium identity and


marks a population of AGP-like progenitors within the adrenal gland. Dev Cell
27, 5-18 (2013). https://doi.org:10.1016/j.devcel.2013.09.003
32 Dörner, J. et al. GLI1(+) progenitor cells in the adrenal capsule of the adult mouse
give rise to heterotopic gonadal-like tissue. Mol Cell Endocrinol 441, 164-175
(2017). https://doi.org:10.1016/j.mce.2016.08.043
33 Poli, G. et al. Human fetal adrenal cells retain age-related stem- and endocrine-
differentiation potential in culture. Faseb j 33, 2263-2277 (2019).
https://doi.org:10.1096/fj.201801028RR
34 Revollo, J. R. et al. HES1 is a master regulator of glucocorticoid receptor-
dependent gene expression. Sci Signal 6, ra103 (2013).
https://doi.org:10.1126/scisignal.2004389
35 Wood, M. A. et al. Fetal adrenal capsular cells serve as progenitor cells for
steroidogenic and stromal adrenocortical cell lineages in M. musculus.
Development 140, 4522-4532 (2013). https://doi.org:10.1242/dev.092775
36 Grabek, A. et al. The Adult Adrenal Cortex Undergoes Rapid Tissue Renewal in a
Sex-Specific Manner. Cell Stem Cell 25, 290-296.e292 (2019).
https://doi.org:10.1016/j.stem.2019.04.012
37 Suzuki, T. et al. Chicken ovalbumin upstream promoter transcription factor II in
the human adrenal cortex and its disorders. J Clin Endocrinol Metab 85, 2752-
2757 (2000). https://doi.org:10.1210/jcem.85.8.6730
38 Kool, M. M. et al. Insulin-like growth factor--phosphatidylinositol 3 kinase
signaling in canine cortisol-secreting adrenocortical tumors. J Vet Intern Med 29,
214-224 (2015). https://doi.org:10.1111/jvim.12528
39 Trapnell, C. et al. The dynamics and regulators of cell fate decisions are revealed
by pseudotemporal ordering of single cells. Nat Biotechnol 32, 381-386 (2014).
https://doi.org:10.1038/nbt.2859
40 Hadjidemetriou, I. et al. DLK1/PREF1 marks a novel cell population in the
human adrenal cortex. J Steroid Biochem Mol Biol 193, 105422 (2019).
https://doi.org:10.1016/j.jsbmb.2019.105422
41 Altieri, B., Colao, A. & Faggiano, A. The role of insulin-like growth factor
system in the adrenocortical tumors. Minerva Endocrinol 44, 43-57 (2019).
https://doi.org:10.23736/s0391-1977.18.02882-1
42 Kim, J. T. et al. NDRG2 and PRA1 interact and synergistically inhibit T-cell
factor/β-catenin signaling. FEBS Lett 586, 3962-3968 (2012).
https://doi.org:10.1016/j.febslet.2012.09.045
43 Backman, S. et al. RNA Sequencing Provides Novel Insights into the
Transcriptome of Aldosterone Producing Adenomas. Sci Rep 9, 6269 (2019).
https://doi.org:10.1038/s41598-019-41525-2
44 Lefèvre, L. et al. Combined transcriptome studies identify AFF3 as a mediator of
the oncogenic effects of β-catenin in adrenocortical carcinoma. Oncogenesis 4,
e161-e161 (2015). https://doi.org:10.1038/oncsis.2015.20
45 Durand, J., Lampron, A., Mazzuco, T. L., Chapman, A. & Bourdeau, I.
Characterization of differential gene expression in adrenocortical tumors
harboring beta-catenin (CTNNB1) mutations. J Clin Endocrinol Metab 96,
E1206-1211 (2011). https://doi.org:10.1210/jc.2010-2143

34
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

46 Zhang, Y. et al. The RNA N6-Methyladenosine Demethylase FTO Promotes


Head and Neck Squamous Cell Carcinoma Proliferation and Migration by
Increasing CTNNB1. Int J Gen Med 14, 8785-8795 (2021).
https://doi.org:10.2147/ijgm.S339095
47 Zhou, S. et al. FTO regulates the chemo-radiotherapy resistance of cervical
squamous cell carcinoma (CSCC) by targeting β-catenin through mRNA
demethylation. Mol Carcinog 57, 590-597 (2018).
https://doi.org:10.1002/mc.22782
48 Fernandez-Gonzalez, R. & Peifer, M. Powering morphogenesis: multiscale
challenges at the interface of cell adhesion and the cytoskeleton. Mol Biol Cell 33
(2022). https://doi.org:10.1091/mbc.E21-09-0452
49 Goto, N. et al. Lymphatics and fibroblasts support intestinal stem cells in
homeostasis and injury. Cell Stem Cell 29, 1246-1261 e1246 (2022).
https://doi.org:10.1016/j.stem.2022.06.013
50 Steenblock, C. et al. Isolation and characterization of adrenocortical progenitors
involved in the adaptation to stress. Proceedings of the National Academy of
Sciences 115, 12997-13002 (2018). https://doi.org:doi:10.1073/pnas.1814072115
51 Walczak, E. M. & Hammer, G. D. Regulation of the adrenocortical stem cell
niche: implications for disease. Nat Rev Endocrinol 11, 14-28 (2015).
https://doi.org:10.1038/nrendo.2014.166
52 Assié, G. et al. Integrated genomic characterization of adrenocortical carcinoma.
Nat Genet 46, 607-612 (2014). https://doi.org:10.1038/ng.2953
53 Lippert, J. et al. Targeted Molecular Analysis in Adrenocortical Carcinomas: A
Strategy Toward Improved Personalized Prognostication. J Clin Endocrinol
Metab 103, 4511-4523 (2018). https://doi.org:10.1210/jc.2018-01348
54 Etheridge, S. L., Spencer, G. J., Heath, D. J. & Genever, P. G. Expression
profiling and functional analysis of wnt signaling mechanisms in mesenchymal
stem cells. Stem Cells 22, 849-860 (2004). https://doi.org:10.1634/stemcells.22-5-
849
55 Berthon, A. et al. WNT/β-catenin signalling is activated in aldosterone-producing
adenomas and controls aldosterone production. Hum Mol Genet 23, 889-905
(2014). https://doi.org:10.1093/hmg/ddt484
56 Ruggiero, C. et al. Integrative genomic analysis reveals a conserved role for
prolactin signalling in the regulation of adrenal function. Clin Transl Med 11,
e630 (2021). https://doi.org:10.1002/ctm2.630
57 Gao, X. et al. Intratumoral heterogeneity of the tumor cells based on in situ
cortisol excess in cortisol-producing adenomas; ∼An association among
morphometry, genotype and cellular senescence∼. J Steroid Biochem Mol Biol
204, 105764 (2020). https://doi.org:10.1016/j.jsbmb.2020.105764
58 Rege, J. et al. Targeted Mutational Analysis of Cortisol-Producing Adenomas. J
Clin Endocrinol Metab 107, e594-e603 (2022).
https://doi.org:10.1210/clinem/dgab682
59 De Sousa, K. et al. Genetic, Cellular, and Molecular Heterogeneity in Adrenals
With Aldosterone-Producing Adenoma. Hypertension 75, 1034-1044 (2020).
https://doi.org:10.1161/hypertensionaha.119.14177

35
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

60 Corn, K. C., Windham, M. A. & Rafat, M. Lipids in the tumor microenvironment:


From cancer progression to treatment. Prog Lipid Res 80, 101055 (2020).
https://doi.org:10.1016/j.plipres.2020.101055
61 Qi, J. et al. Single-cell and spatial analysis reveal interaction of FAP+ fibroblasts
and SPP1+ macrophages in colorectal cancer. Nature Communications 13, 1742
(2022). https://doi.org:10.1038/s41467-022-29366-6
62 Liu-Chittenden, Y. et al. RARRES2 functions as a tumor suppressor by
promoting β-catenin phosphorylation/degradation and inhibiting p38
phosphorylation in adrenocortical carcinoma. Oncogene 36, 3541-3552 (2017).
https://doi.org:10.1038/onc.2016.497
63 Alfaro-Arnedo, E. et al. IGF1R acts as a cancer-promoting factor in the tumor
microenvironment facilitating lung metastasis implantation and progression.
Oncogene 41, 3625-3639 (2022). https://doi.org:10.1038/s41388-022-02376-w
64 Yao, M. et al. Astrocytic trans-Differentiation Completes a Multicellular
Paracrine Feedback Loop Required for Medulloblastoma Tumor Growth. Cell
180, 502-520.e519 (2020). https://doi.org:10.1016/j.cell.2019.12.024
65 London, E. et al. Cholesterol Biosynthesis and Trafficking in Cortisol-Producing
Lesions of the Adrenal Cortex. J Clin Endocrinol Metab 100, 3660-3667 (2015).
https://doi.org:10.1210/jc.2015-2212
66 Horvath, A. et al. A cAMP-specific phosphodiesterase (PDE8B) that is mutated in
adrenal hyperplasia is expressed widely in human and mouse tissues: a novel
PDE8B isoform in human adrenal cortex. Eur J Hum Genet 16, 1245-1253
(2008). https://doi.org:10.1038/ejhg.2008.85
67 Torbenson, M. et al. Morphological heterogeneity in beta-catenin-mutated
hepatocellular carcinomas: implications for tumor molecular classification. Hum
Pathol 119, 15-27 (2022). https://doi.org:10.1016/j.humpath.2021.09.009
68 Friemel, J. et al. Intratumor heterogeneity in hepatocellular carcinoma. Clin
Cancer Res 21, 1951-1961 (2015). https://doi.org:10.1158/1078-0432.Ccr-14-
0122
69 Kaur, A. et al. WNT inhibition creates a BRCA-like state in Wnt-addicted cancer.
EMBO Mol Med 13, e13349 (2021). https://doi.org:10.15252/emmm.202013349
70 Musa, J., Aynaud, M. M., Mirabeau, O., Delattre, O. & Grünewald, T. G. MYBL2
(B-Myb): a central regulator of cell proliferation, cell survival and differentiation
involved in tumorigenesis. Cell Death Dis 8, e2895 (2017).
https://doi.org:10.1038/cddis.2017.244
71 Zhao, J., Liu, B. & Li, X. A transcription factor signature predicts the survival of
patients with adrenocortical carcinoma. PeerJ 9, e12433 (2021).
https://doi.org:10.7717/peerj.12433
72 Taniguchi-Ponciano, K. et al. Proteomic and Transcriptomic Analysis Identify
Spliceosome as a Significant Component of the Molecular Machinery in the
Pituitary Tumors Derived from POU1F1- and NR5A1-Cell Lineages. Genes
(Basel) 11 (2020). https://doi.org:10.3390/genes11121422
73 Armenia, J. et al. The long tail of oncogenic drivers in prostate cancer. Nat Genet
50, 645-651 (2018). https://doi.org:10.1038/s41588-018-0078-z
74 Gahete, M. D., Herman-Sanchez, N., Fuentes-Fayos, A. C., Lopez-Canovas, J. L.
& Luque, R. M. Dysregulation of splicing variants and spliceosome components

36
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

in breast cancer. Endocr Relat Cancer (2022). https://doi.org:10.1530/erc-22-


0019
75 Landwehr, L. S. et al. Interplay between glucocorticoids and tumor-infiltrating
lymphocytes on the prognosis of adrenocortical carcinoma. J Immunother Cancer
8 (2020). https://doi.org:10.1136/jitc-2019-000469
76 Mitani, F., Mukai, K., Miyamoto, H., Suematsu, M. & Ishimura, Y. Development
of functional zonation in the rat adrenal cortex. Endocrinology 140, 3342-3353
(1999). https://doi.org:10.1210/endo.140.7.6859
77 Murray, P. J. et al. Macrophage activation and polarization: nomenclature and
experimental guidelines. Immunity 41, 14-20 (2014).
https://doi.org:10.1016/j.immuni.2014.06.008
78 Yang, Y. et al. Crosstalk between hepatic tumor cells and macrophages via
Wnt/β-catenin signaling promotes M2-like macrophage polarization and
reinforces tumor malignant behaviors. Cell Death & Disease 9, 793 (2018).
https://doi.org:10.1038/s41419-018-0818-0
79 Thorsson, V. et al. The Immune Landscape of Cancer. Immunity 48, 812-
830.e814 (2018). https://doi.org:10.1016/j.immuni.2018.03.023
80 Geiger, J. et al. Hospital-integrated Biobanking as a Service – The
Interdisciplinary Bank of Biomaterials and Data Wuerzburg (ibdw). Open Journal
of Bioresources 5 (2018). https://doi.org:10.5334/ojb.38
81 Altieri, B. et al. Effects of Germline CYP2W1*6 and CYP2B6*6 Single
Nucleotide Polymorphisms on Mitotane Treatment in Adrenocortical Carcinoma:
A Multicenter ENSAT Study. Cancers (Basel) 12 (2020).
https://doi.org:10.3390/cancers12020359
82 Nieman, L. K. et al. The Diagnosis of Cushing's Syndrome: An Endocrine
Society Clinical Practice Guideline. The Journal of Clinical Endocrinology &
Metabolism 93, 1526-1540 (2008). https://doi.org:10.1210/jc.2008-0125
83 Detomas, M. et al. Case Report: Consecutive Adrenal Cushing's Syndrome and
Cushing's Disease in a Patient With Somatic CTNNB1, USP8, and NR3C1
Mutations. Front Endocrinol (Lausanne) 12, 731579 (2021).
https://doi.org:10.3389/fendo.2021.731579
84 Krishnaswami, S. R. et al. Using single nuclei for RNA-seq to capture the
transcriptome of postmortem neurons. Nature Protocols 11, 499-524 (2016).
https://doi.org:10.1038/nprot.2016.015
85 Hashimshony, T. et al. CEL-Seq2: sensitive highly-multiplexed single-cell RNA-
Seq. Genome Biology 17, 77 (2016). https://doi.org:10.1186/s13059-016-0938-8
86 Klein, A. M. et al. Droplet barcoding for single-cell transcriptomics applied to
embryonic stem cells. Cell 161, 1187-1201 (2015).
https://doi.org:10.1016/j.cell.2015.04.044
87 Stuart, T. et al. Comprehensive Integration of Single-Cell Data. Cell 177, 1888-
1902.e1821 (2019). https://doi.org:10.1016/j.cell.2019.05.031
88 Parekh, S., Ziegenhain, C., Vieth, B., Enard, W. & Hellmann, I. zUMIs - A fast
and flexible pipeline to process RNA sequencing data with UMIs. GigaScience 7
(2018). https://doi.org:10.1093/gigascience/giy059

37
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

89 Butler, A., Hoffman, P., Smibert, P., Papalexi, E. & Satija, R. Integrating single-
cell transcriptomic data across different conditions, technologies, and species.
Nature Biotechnology 36, 411-420 (2018). https://doi.org:10.1038/nbt.4096
90 Ulgen, E., Ozisik, O. & Sezerman, O. U. pathfindR: An R Package for
Comprehensive Identification of Enriched Pathways in Omics Data Through
Active Subnetworks. Front Genet 10, 858 (2019).
https://doi.org:10.3389/fgene.2019.00858
91 Hafemeister, C. & Satija, R. Normalization and variance stabilization of single-
cell RNA-seq data using regularized negative binomial regression. Genome
Biology 20, 296 (2019). https://doi.org:10.1186/s13059-019-1874-1
92 Korsunsky, I. et al. Fast, sensitive and accurate integration of single-cell data with
Harmony. Nature Methods 16, 1289-1296 (2019). https://doi.org:10.1038/s41592-
019-0619-0
93 Altieri, B. et al. Livin/BIRC7 expression as malignancy marker in adrenocortical
tumors. Oncotarget 8, 9323-9338 (2017).
https://doi.org:10.18632/oncotarget.14067
94 Laufs, V. et al. ERCC1 as predictive biomarker to platinum-based chemotherapy
in adrenocortical carcinomas. Eur J Endocrinol 178, 181-188 (2018).
https://doi.org:10.1530/eje-17-0788

38
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Figure legends

Fig 1. A comprehensive adult human normal adrenal gland cell atlas at single-nuclei

resolution.

A. Left: Transverse section depicting the three major adrenal zones (capsule, adrenal cortex, and

medulla). Right: UMAP (Uniform Manifold Approximation and Projection) representation of the

six integrated single-nuclei transcriptomic datasets (clusters were annotated by using known

marker genes and DEG analysis). B. Adrenal cortex clusters: left: UMAP representation of the

adrenal cortex clusters with complete zonation inferred by module scoring (see Fig. 2); right:

feature expression plots representing genes specific for zonae reticularis, fasciculata, glomerulosa

and adreno-medullary progenitors (scale represents log normalized expression values). C.

Anatomical sketch of the adult human normal adrenal gland: the displayed markers for each

identified cell type are based on DGE analysis in single-cell transcriptome data, available

literature (see text for references) and in-situ validations by immunohistochemistry. D. Satellite

clusters. Feature expression plots representing genes used in annotation of following clusters:

“Medulla”, “Myeloid cells”, “Lymphoid cells”, “Stem/P&VEC“ (Stem/Progenitor & Vascular

Endothelial cells) and “Fibroblasts & Connective tissue” (scale represents log normalized

expression values).

Fig 2. Correlation of single-nuclei transcriptome clusters with the zonation of the adrenal

cortex

A. Kernel density estimation of selected features (in frames) for each of the three adrenocortical

zones: ZG (CYP11B2, DACH1), ZF (CYP17A1) and ZR (CYB5A, SULT2A1). B. Example of

hematoxylin and eosin (H&E) staining that served to distinguish the different zones of the normal

adrenal gland (at 10x enlargement all the three zones of the cortex (glomerulosa, ZG; fasciculata,

ZF; reticularis, ZR) together with the capsule (C) and the adrenal medulla (M) were observed). C.

39
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Consecutive adrenal sections were stained by immunohistochemistry (IHC) for the markers that

are indicated in the boxes in A. CYP11B2 staining was confirmed to be expressed in the ZG as

sporadic and scattered cells underneath the adrenal capsule or as clusters of cells (previously

termed as aldosterone-producing cell clusters). DACH1 was expressed in the nuclei of the ZG

cells. CYP17A1 was strongly expressed in the cytoplasm of adrenocortical cells of ZF and ZR,

but not in the ZG. CYB5A1 was selectively stained in the cytoplasm of cells of the ZR, whereas

SULT2A1 was stained in both ZF and ZR, although in the ZR a stronger staining was observed.

All images were acquired by Leica Aperio Versa brightfield scanning microscope (Leica,

Germany). 1x pictures with an enlargement of 10x of the same area of consecutive slides were

used to obtain the IHC pictures.

Fig. 3. Adreno-medullary and adrenocortical progenitor cells in adult human normal

adrenal glands.

A. RNA expression of SYT1 and CHGA in the AMP cluster. B.1. SYT1 and CHGA

immunostaining. A low SYT1 protein expression was observed in cells of the entire cortex,

whereas cells of the medulla showed moderate expression. Interestingly, groups of cells localized

in the subcapsular region (organized as niches) showed a strong SYT1 staining. CHGA staining

was found to be pronounced in medullary cells but was very low in cells of the inner part of the

adrenal cortex and no staining was observed in the outer part of the adrenal cortex and in the

adrenal capsule. B.2. CHGA/SYT1 double staining. At 2x enlargement, cells of the medulla

showed a strong CHGA staining (fuchsin-red color). At 10x enlargement, a SYT1 strong positive

niche (starker green-blue color) was marked with a yellow line. All images were acquired by

Leica Aperio Versa brightfield scanning microscope (Leica, Germany). A 1x picture with an

enlargement of 10x and 20x of the same area of consecutive slides were used to obtain the IHC

pictures. C. Expression of NR2F2 and ID1 genes in the Stem/P&VEC cluster. D.1. NR2F2 and

ID1 immunostaining. IHC analysis revealed sparse, rare cells mostly localized at the subcapsular

40
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

levels that showed a high nucleus staining of NR2F2 or ID1 (marked with yellow circles). Those

cells located under the capsule seemed to centripetally migrate from ZG to ZF and ZR (orange

arrows represented the centripetal migration in adrenal glands). D.2. NR2F2/ID1 double staining.

Nuclei of cells positive for both NR2F2 and ID1 were stained in dark blue (because of the sum of

fuchsin-red + blue-green colours) and were mostly located under the capsule. Double positive

NR2F2/ID1 nuclei were indicated with red circles at 10x enlargement and red arrows at 20x

enlargement. Orange arrows at 10x enlargement represent the centripetal migration in adrenal

glands.

Fig. 4. Validation of adrenal zonation by spatial transcriptomics in adult human normal

adrenal glands.

A pairwise (cell-to-spot) score is calculated within the label transfer framework in Seurat (V3)

and projected onto sections. Two sections are represented for each mapped cluster. Medulla and

AMP clusters were rendered transparent as the label transfer did not detect the presence of these

cell types in the Visium assay.

Fig. 5: Trajectory analysis of spatial transcriptomics data in adult human normal adrenal

glands.

A. Top: Integrated UMAP (Uniform Manifold Approximation and Projection) from both tissue

sections analysed using the Visium assay (10x Genomics). Trajectory was visualised as a black

line. Bottom: clusters identified in the integrated object transferred to the tissue (Visium)

sections. B. Pseudotime estimation both for the UMAP and the tissue sections: white circles

represent the root node (chosen based on RSPO3 expression, indicated with a black arrow (right

panel)). C. Heatmap of 829 differentially expressed genes across pseudotime: cells were ordered

by pseudotime (columns), genes were ordered by expression and pseudotime (rows). Top genes

are highlighted. D. Expression variation of RSPO3, ID1, INSR and CYP11B2 over pseudotime:

41
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

smoothed lines were generated based on the scatter profile of each gene (95% confidence interval

displayed in grey around the line). Vertical lines represent the transition zones from 0-Capsule

(red) to 1-ZG (green) and 1-ZG to 2-ZF-ZR (ochre).

Figure 6. Identification of adenoma-specific clusters in adrenocortical adenoma (ACA)

A. 1. Volcano plot depicting differentially expressed genes between NAG and EIA. Genes

characteristic of EIA subtype are written in bold. The x axis represents fold change (Log2FC) and

y axis represents P values (-Log10P). 2. Volcano plot depicting differentially expressed genes

between NAG and CPA. Genes characteristic of CPA subtype are written in bold. 3. Violin plots

showing log normalized expression values of common signatures across the three subtypes

(NAG, EIA, CPA). 4. Left: Violin plots showing log normalized EIA specific gene expression

values. Right: Violin plots showing log normalized CPA specific gene expression values.

B. UMAP (Uniform Manifold Approximation and Projection) representation of the 18 integrated

samples (normal adrenal glands (NAG, n=6); cortisol-producing adenomas (CPA, n=7);

endocrine inactive adenomas (EIA, n=5) coloured by cluster identity. The UMAP could be

divided into two groups: the “central cluster” containing previously identified cortex cell types

(ZG, ZF and ZR) as well as ZG-like, Adenoma specific (AC1, AC2, AC3, and AC4),

Cholesterol-and steroid-enriched metabolism (CSEM) and adenoma microenvironment (AME)

clusters, and the “satellite clusters”, which are Medulla, Myeloid cells, Lymphoid cells ,

Endothelial and Stromal. NAG cell type distribution in the UMAP is represented within the top

right box. C. Scaled average expression of selected markers for each cluster (dot size represents

the percentage of cells in each cluster expressing the marker). D. Subtype and sample specific

composition coloured by cluster identity: samples (x axis) are ordered by mutational signatures.

Figure 7. Transcriptomic profile of the newly identified ACA clusters

A. Heatmap representing top identified features for the 7 newly identified adenoma clusters.

42
bioRxiv preprint doi: https://doi.org/10.1101/2022.08.27.505530; this version posted August 28, 2022. The copyright holder for this preprint
(which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

Colour scale represents Z-score computed from log normalized expression matrix.

B. Top 5 significant pathways for each cluster combining results from 3 separate databases

(KEGG, Reactome and GO-All (GO-MF, GO-BP, GO-CC))

Figure 8. Cluster distribution in ACA samples across subtype and mutational status

Distribution of A. Cholesterol-and steroid-enriched metabolism (CSEM) cluster in cortisol

producing adenomas (CPA) versus endocrine inactive adenomas (EIA; upper panel) and across

mutational status (lower panel); B. Adenoma clusters (AC1-4) across mutational status; C.

Lymphoid cluster across subtypes (upper panel) and mutational status (lower panel); D. Myeloid

cluster across subtypes (upper panel) and mutational status (lower panel). Significant cluster

enrichments are encircled in red.

43
Fig. 1
Fig. 2
A Markers UMAP C IHC staining

CYP11B2 CYP11B2 CYP11B2 DACH1


LGR5
DACH1

ZG CACNB2
DACH1
CALN1
MCOLN3
B
H&E staining
MC2R
ANO4

CYP17A1 CYP17A1

HSD3B2 CYP17A1
FDX1
ZF CYP11B1
POR
NOV M
ZR
LEF1
ZF
CYB5A C
CYB5A SULT2A1
ZG
SULT2A1
CYB5A
TSPAN12
ZR SULT2A1
PAPSS2
CYP17A1
HSD3B2

Low High
Density
C
Fig. 3

ID1
SYT1

NR2F2
CHGA
ZG ZG ZG ZG
ZF ZF ZF ZF
ZR ZR ZR ZR
AM AM AM AM
AM AM AM AM
P P P P
M M M M
St C St C St C St C
em em em em
/P L /P L /P L /P LC
& C & C & C &
VE VE VE VE
C C C C
FC FC FC FC
B

D
1

ID1
SYT1

NR2F2
CHGA

20x

20x
2

2
NR2F2 + ID1
CHGA + SYT1

20x
10x
2x

10x
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Supplementary figures
Supplementary Fig. 1
Supplementary Fig. 2
Supplementary Fig. 3
Supplementary Fig. 4

A Fibroblast & connective tissue

H&E

COL1A2 MGP

M Stem/P&VEC Medulla
ZR

ZF
C
ZG

ENTPD1 CHGA

B C
Supplementary Fig. 5
Supplementary Fig. 6

Miscellaneous

Wnt/β-cat
Supplementary Fig. 7
Supplementary Fig. 8
Supplementary Fig. 9
NAG EIA CPA
A

Lymphoid cells
Expression
1

0.5

-0.5
B -1

Myeloid cells

You might also like