Skip to main content
eLife logoLink to eLife
. 2019 Mar 26;8:e43882. doi: 10.7554/eLife.43882

Single-cell expression profiling reveals dynamic flux of cardiac stromal, vascular and immune cells in health and injury

Nona Farbehi 1,2,3,4,, Ralph Patrick 1,2,5,, Aude Dorison 1,2, Munira Xaymardan 1,2,6, Vaibhao Janbandhu 1,2,5, Katharina Wystub-Lis 1, Joshua WK Ho 1,5, Robert E Nordon 2,4,, Richard P Harvey 1,2,7,
Editors: Edward Morrisey8, Harry C Dietz9
PMCID: PMC6459677  PMID: 30912746

Abstract

Besides cardiomyocytes (CM), the heart contains numerous interstitial cell types which play key roles in heart repair, regeneration and disease, including fibroblast, vascular and immune cells. However, a comprehensive understanding of this interactive cell community is lacking. We performed single-cell RNA-sequencing of the total non-CM fraction and enriched (Pdgfra-GFP+) fibroblast lineage cells from murine hearts at days 3 and 7 post-sham or myocardial infarction (MI) surgery. Clustering of >30,000 single cells identified >30 populations representing nine cell lineages, including a previously undescribed fibroblast lineage trajectory present in both sham and MI hearts leading to a uniquely activated cell state defined in part by a strong anti-WNT transcriptome signature. We also uncovered novel myofibroblast subtypes expressing either pro-fibrotic or anti-fibrotic signatures. Our data highlight non-linear dynamics in myeloid and fibroblast lineages after cardiac injury, and provide an entry point for deeper analysis of cardiac homeostasis, inflammation, fibrosis, repair and regeneration.

Research organism: Mouse

eLife digest

In our bodies, heart attacks lead to cell death and inflammation. This is then followed by a healing phase where the organ repairs itself. There are many types of heart cells, from muscle and pacemaker cells that help to create the beating motion, to so-called fibroblasts that act as a supporting network. Yet, it is still unclear how individual cells participate in the heart's response to injury.

All cells possess the same genetic information, but they turn on or off different genes depending on the specific tasks that they need to perform. Spotting which genes are activated in individual cells can therefore provide clues about their exact roles in the body. Until recently, technological limitations meant that this information was difficult to access, because it was only possible to capture the global response of a group of cells in a sample.

A new method called single-cell RNA sequencing is now allowing researchers to study the activities of many genes in thousands of individual cells at the same time. Here, Farbehi, Patrick et al. performed single-cell RNA sequencing on over 30,000 individual cells from healthy and injured mouse hearts. Computational approaches were then used to cluster cells into groups according to the activities of their genes.

The experiments identified over 30 distinct sub-types of cell, including several that were previously unknown. For example, a group of fibroblasts that express a gene called Wif1 was discovered. Previous genetic studies have shown that Wif1 is essential for the heart's response to injury. Further experiments by Farbehi, Patrick et al. indicated that this new sub-type of cells may control the timing of the different aspects of heart repair after damage.

Tens of millions of people around the world suffer from heart attacks and other heart diseases. Knowing how different types of heart cells participate in repair mechanisms may help to find new targets for drugs and other treatments.

Introduction

Cardiovascular disease including myocardial infarction (MI) remains a leading cause of morbidity and mortality in the Western and developing worlds. After acute MI, millions of cardiomyocytes (CM) are lost by necrosis and apoptosis, and an initially adaptive collagen-rich scar is laid down to preserve chamber geometry and prevent rupture. The mammalian heart is regarded as being poorly regenerative as the long-term sequelae in virtually all etiologies of heart disease involve increased wall stiffness, reduced heart function and progression to heart failure. However, some inbred strains of mice show surprising cardiac reparative abilities (Patterson et al., 2017), and CM renewal and heart regeneration can be stimulated experimentally (D'Uva et al., 2015; Mohamed et al., 2018; Srivastava and DeWitt, 2016; Wang et al., 2018), garnering optimism that heart regeneration can be achieved in humans.

Cardiac chamber walls are composed of a complex, interdependent community of interstitial cells, including vascular, fibroblast, immune and neuronal cells, although how they interact in cardiac homeostasis, injury and repair, is relatively unexplored. In regenerative systems, connective tissues play key roles in defining positional information, and organizing tissue architecture and niche environments (Nacu et al., 2013; Chan et al., 2013; Greicius et al., 2018). Cardiac fibroblasts represent ~10% of all cardiac cells (Pinto et al., 2016) and are distributed throughout the cardiac interstitial, perivascular and sub-epicardial spaces, where they are proposed to have sentinel, paracrine, mechanical, extracellular matrix (ECM) and electrical functions (Shinde and Frangogiannis, 2014; Tallquist and Molkentin, 2017). After injury, inflammation is principally executed by poly-functional monocytes (Mo) and macrophages (MΦ), and is necessary to protect against pathogens and autoimmunity, and to coordinate healing. Fibroblasts also participate in inflammation and phagocytosis and are the principal drivers of fibrotic repair (Shinde and Frangogiannis, 2014; Gourdie et al., 2016). In heart repair, timely resolution of inflammation is necessary for limiting fibrosis and enabling tissue replacement, while uncontrolled inflammation leads to increased fibrosis and chamber wall stiffening, poor electro-mechanical coupling, continued loss of CMs and worsening outcomes (Mescher, 2017; Lai et al., 2017; Williams et al., 2018).

The general principles of inflammation and fibrosis have been mapped in different organs, and the implementation of specific lineage-tracing tools has provided significant new insights into cardiac leukocyte and fibroblast origins and fate (Tallquist and Molkentin, 2017; Williams et al., 2018; Fu et al., 2018; Ivey et al., 2018; Kanisicak et al., 2016; Moore-Morris et al., 2014; Ensan et al., 2016; Heidt et al., 2014; Molawi et al., 2014; Epelman et al., 2014; Plein et al., 2018). However, controversies persist around nomenclature, defining markers, origins, heterogeneity and plasticity (Tallquist and Molkentin, 2017; Epelman et al., 2015; Swirski and Nahrendorf, 2018). Even the question of whether the transitions from quiescent to activated fibroblast, then to myofibroblast, should be seen as differentiation in the classical sense, degrees of a scalable and reversible continuum governed by the injury environment, or a branched dynamic network, is unresolved (Tallquist and Molkentin, 2017; Ivey and Tallquist, 2016; Travers et al., 2016).

One approach to a deeper understanding of cardiac population biology is through single-cell genomics, including single-cell RNA sequencing (scRNA-seq). Single-cell methods have the power to overcome the limitations of bulk cell analyses, where insights into complex cell system dynamics are lost (Tanay and Regev, 2017). The rich data generated by single-cell methods allow new computational frameworks for inferring cell dynamics and causality, unencumbered by strict a priori notions of cell identity, hierarchy, trajectory and markers.

Here, we present the first comprehensive analysis of cellular lineage heterogeneity, dynamics and intercellular communication among immune and stromal (non-CM) cells in healthy and injured adult mouse hearts using scRNA-seq. Clustering analysis of >30,000 cells identified over 30 cell populations across the total non-CM fraction and enriched (Pdgfra-GFP+) fibroblast lineage cells. These populations comprised most of the known cell types and their dynamics after injury, as well as novel cell types and their intermediates. We describe a novel population of activated fibroblasts present in both sham and injured hearts expressing a strong anti-Wingless-related integration site (WNT) transcriptome signature, a putative pre-proliferative state, and three novel myofibroblast subtypes expressing pro-fibrotic or anti-fibrotic (including anti-WNT) signatures. We were also able to distinguish the major tissue-resident and infiltrating Mo/MΦ, and numerous minor populations. Overall, our data reveal dynamic, multi-dimensional lineage trajectories in the injured heart. This deep resource will provide novel insights into the inflammatory and fibrotic cascades in the injured mouse heart that may suggest novel molecular or cellular targets for enhancing heart repair and regeneration in man.

Results

Single-cell RNA-seq of total cardiac interstitial cell population

We performed single-cell expression profiling on the total cardiac interstitial cell population (TIP) using the 10x Genomics Chromium platform, from hearts of 8 weeks old male PdgfraGFP/+ mice at days 3 and 7 post-sham or MI surgery. To enrich for cells relevant to cardiac ischemic injury and repair, we isolated TIP cells from dissected ventricles and interventricular septum, excluding cells of the atria, annulus fibrosus and atrioventricular valves (Figure 1—figure supplement 1A).

Transcriptional profiles of 13,331 cells were captured after quality control filtering (sham: 5,723; MI-day 3: 3,875; MI-day 7: 3,733). To identify cells with distinct lineage identities and transcriptional states, we performed unbiased clustering on an aggregate of cells using the Seurat R package (Butler et al., 2018), with cell populations visualized in t-SNE dimensionality reduction plots (Materials and methods). For initial analyses, populations expressing markers of multiple lineages (hybrids) were removed; however, select examples are discussed in more detail below.

TIP cells were represented by a total of 24 populations and nine distinct cell lineages (Figure 1A–D; Figure 1—figure supplement 1B–F). Major cell types comprised fibroblasts/myofibroblasts (Col1a1+Pdgfra+GFP+), endothelial cells (ECs; Kdr+Pecam1+), mural cells (Cspg4+Pdgfrb+), Mo and MΦ (Cd68+Itgam+), dendritic-like (DC) cells (Cd209a+Itgam+), glial cells (Plp1+Kcna1+), B-cells (Cd79a+Ms4a1+), T-cells (Cd3e+Cd3d+Lef1+) and natural killer cells (NKCs; Klrk1+Ccl5+) (Figure 1A–E; Figure 1—figure supplement 1D; Supplementary file 1). New lineage markers were identified; for example, Vtn, encoding Vitronectin, was specifically expressed in mural cells, whereas Kcna1, encoding the potassium voltage-gated channel subfamily A member 1, was highly specific to glial cells (Figure 1D and Figure 1—figure supplement 1D).

Figure 1. TIP scRNA-seq.

(A) t-SNE plots showing detected lineages and sub-populations in TIP across conditions. (B) t-SNE plot of aggregate TIP cells with identified sub-populations. (C) Dendrogram of sub-populations according to average RNA expression. (D) Expression of select marker genes across TIP cells as visualized on t-SNE plots. (E) Cell population percentages across conditions determined to be significantly modulated according to Differential Proporation Analysis (DPA) (p<0.01).

Figure 1—source data 1. Source data for FACS quantifications summarized in Figure 1—figure supplement 6D,E and Figure 1—figure supplement 7B,C.
DOI: 10.7554/eLife.43882.012

Figure 1.

Figure 1—figure supplement 1. Experimental procedures, population proportions and gene expression characterstics of sub-populations within TIP scRNA-seq.

Figure 1—figure supplement 1.

(A) Diagram of experimental procedures. (B) Percentage of total cells in each cell-population according to experimental condition. (C) Percentage of total cells in clusters as grouped together according to high-level cell type and as according to experimental condition. (D) Dot plot showing expression of top upregulated genes across TIP populations. (E, F) Box plot showing number of UMIs (E) and genes (F) detected across TIP sub-populations after quality control filtering.
Figure 1—figure supplement 2. Gene expression data for TIP sub-populations.

Figure 1—figure supplement 2.

(A) Distribution of Pdgfra and Pdgfra-GFP expression across cell populations in TIP. (B) Expression of selected endothelial markers as visualized in box and t-SNE plots. (C) Expression of selected cell cycle marker genes on t-SNE plots. (D) Dot plot visualization according to condition of top marker genes for lymphoid lineage populations.
Figure 1—figure supplement 3. Differential proportion analysis (DPA) procedure and evaluation.

Figure 1—figure supplement 3.

(A,B) Procedure for performing DPA. (C) Percentage of cell populations between sham conditions from GFP+ cohort. (D) Comparisons of p-values from Fisher’s exact test and DPA for evaluating proportion changes between sham experiments. (E) Comparisons of specificity for Fisher’s exact test and DPA on control replicate simulation experiment. (F–H) Sensitivity, specificity and precision for simulated control vs condition experiments with comparisons between Fisher’s exact test and DPA.
Figure 1—figure supplement 4. Clustering of TIP scRNA-seq prior to removal of minor hybrid populations.

Figure 1—figure supplement 4.

(A) t-SNE plot of aggregate TIP cells with all identified 29 sub-populations prior to filtering of minor hybrid populations. (B) Dendrogram of all 29 sub-populations according to average RNA expression. (C) Expression of example marker genes that illustrate hybrid populations (F-EC, M2MΦ-EC, EC-L1, EC-L2 and BC-TC) as visualized in box and t-SNE plots. (E) Percentage of total cells in each sub-population according to experimental condition.
Figure 1—figure supplement 5. A typical workflow of sequential gating strategy for doublet exclusion is shown.

Figure 1—figure supplement 5.

(A) Scatter gate, FSC-A vs SSC-A, was plotted to exclude cell debris; the three subsequent ‘doublet discrimination’ gating strategies to exclude doublets are: (B) FSC-H vs FSC-A plot, (C) FSC-H vs FSC-W and (D) SSC-H vs SSC-W. Blue and red gates depict ‘single cell’ and ‘doublet’ events, respectively. Only the cells in the blue gates were sorted and/or analyzed. The scatter profiles of putative doublets (red dots) are also illustrated in panel (C) and (D).
Figure 1—figure supplement 6. FACS data supporting F-EC population.

Figure 1—figure supplement 6.

FACS plots showing expression of GFP or CD31 (axis label) in the (A) wild type sample stained with isotype control antibody; (B) wild type sample stained with anti-CD31 antibody; (C) PdfgraGFP/+ sample stained with isotype control antibody; (D, E) PdfgraGFP/+ sample stained with anti-CD31 antibody 7 days after Sham (D) or MI (E) surgery. Numbers indicate percentage of GFP+CD31+ double positive cells in live single cells in respective plot (mean numbers indicated).
Figure 1—figure supplement 7. FACS data supporting the M2 MΦ -EC population.

Figure 1—figure supplement 7.

FACS plots showing expression of indicated antigen (axis label) in the (A) wild type sample stained with isotype control antibodies and PdfgraGFP/+ sample stained with indicated antibodies 7 days after Sham (B) or MI (C) surgery.
Numbers indicate percentage of gated cells in live single cells in respective analyses (mean numbers indicated).
Figure 1—figure supplement 8. Atrial Fibrilation (AF) associated genes in TIP.

Figure 1—figure supplement 8.

Overlap of genes associated with AF from Roselli et al. (2018) and genes upregulated in TIP sub-populations with log2 fold-change >1. Displayed are genes upregulated in EC (top) or mural (bottom) cell sub-populations with expression visualized in box and t-SNE plots. Key indicates the location of sub-populations on the t-SNE plot.

Within the fibroblast lineage, we identified several sub-populations, each marked by expression of Pdgfra, Pdgfra-GFP, Col1a1 and other canonical fiboblast markers (Tallquist and Molkentin, 2017; Ivey and Tallquist, 2016) (Figure 1A–D; Figure 1—figure supplement 2A). We describe these in more detail after enrichment below.

There were three major sub-populations of ECs (EC1, EC2, EC3), comprising vascular and lymphatic lineages (Pecam1+Kdr+Ly6a+) (Figure 1D; Figure 1—figure supplement 2B). The majority EC1 population expressed Ly6a (encoding SCA1) as well as the vascular transcription factor (TF) Sox17, and likely represents microvascular ECs. EC2 expressed canonical arterial endothelial markers such as Bmx, Sema3g and Efnb2, as well as TF genes Sox17 and Hey1 (Figure 1—figure supplement 2B), the latter acting downstream of NOTCH which is required for arterial EC fate. EC3 almost uniquely expressed venous EC marker Nr2f2 (encoding COUPTFII) and Von Willebrand factor (Vwf), and a minority (~3%) expressed Prox1 and Lyve1 (Figure 1—figure supplement 2B), consistent with a lymphatic identity. A small number of ECs were cycling (Cyc; Figure 1A–C; Figure 1—figure supplement 2C). Our EC data are broadly consistent with recently published single-cell data (Zhao et al., 2018).

Among lymphocytes, a single B-cell population (BC) expressed Cd79a, Ms4a1 and Ly6d (Figure 1D; Figure 1—figure supplement 2D). T-cell sub-populations TC1-Cd8 (Cd8a+) and TC2-Cd4 (Cd4+Lef1+) likely represent cytotoxic and helper T-cells, respectively. NKCs exclusively expressed Klrk1 and upregulated Lck, Ccl5 and Ctsw (Figure 1—figure supplement 2D).

Differential proportion analysis for detecting cell population dynamics

We observed major injury-induced cellular responses and flux after MI, including expansion of Mo/MΦ populations at MI-day 3, as well as myofibroblasts and an additional MΦ population at MI-day 7 (Figure 1A; Figure 1—figure supplement 1B,C). To analyze whether changes in the proportion of populations were greater than expected by chance, we developed a novel permutation-based statistical test (differential proportion analysis; DPA) that considered sources of variation which could arise from experimental procedures (such as differing cell numbers and cell-type capture bias) or in silico analysis (cluster assignment accuracy) (Materials and methods; Figure 1—figure supplement 3A–H). DPA identified 12 populations showing significant (p<0.01) flux between conditions (Figure 1E; Supplementary file 2); for example, the fibroblast sub-populations F-SL and F-SH (see below) decreased sharply in proportion at MI-day 3, while M1 and M2 MΦ populations expanded at days 3 and 7 after MI, respectively.

Monocyte/macrophage cell identity and dynamics

Cardiac tissue-resident MΦ originate from CX3CR1+ progenitors in the yolk sac and Mo from fetal liver and post-natal bone marrow (Ensan et al., 2016), and have roles in immunity, coronary artery and pacemaker development, and heart regeneration (Lavine et al., 2014; Hulsmans et al., 2017). Resident MΦ are long lived and self-renewing (Epelman et al., 2014; Bajpai et al., 2018), although some are supplanted by blood-derived Mo with age or injury (Heidt et al., 2014; Molawi et al., 2014; Dick et al., 2019). MI triggers a biphasic cascade of inflammation and remodeling, with the acute phase involving early influx of neutrophils and CCR2+LY6C2high pro-inflammatory M1 Mo/MΦ, which phagocytose debris and secrete pro-inflammatory factors IL-1β, IL-6 and TNFα to amplify the inflammatory response (Swirski and Nahrendorf, 2018). The repair phase begins around MI-day 3 when non-classical LY6C2-F4/80high M2 MΦ accumulate and secrete anti-inflammatory cytokines such as Il-10 and TGF-β, and stimulate angiogenesis (Epelman et al., 2015; Swirski and Nahrendorf, 2018).

In sham hearts, we identified cardiac tissue-resident MΦ with the signature Cx3cr1highAdgre1(F4/80)highH2-Aa(MHC-II)+Itgam(CD11b)lowLy6c2lowCcr2- (MAC-TR; Figure 1A,B; Figure 2A–D) (Ensan et al., 2016; Epelman et al., 2014; Swirski and Nahrendorf, 2018; Lavine et al., 2014; Lavine et al., 2018). Recent work using flow cytometry and scRNA-seq has delineated several subsets of cardiac tissue-resident MΦ, including a pro-regenerative population with the signature TIMD4+LYVE1+MHC-IIlowCCR2-, that self-renew and are not replaced by blood monocytes even after injury (Dick et al., 2019). We could discern this same population at the scRNA-seq level as a subset within MAC-TR, which persisted after injury (Figure 2—figure supplement 1A; Figure 2B). The additional major subset of CCR2- tissue-resident MΦ (Dick et al., 2019) could also be recognised at the scRNA-seq level as the Timd4-Lyve1-H2-Aa(MHC-II)highCcr2- subset of MAC-TR – this population has been shown to have a low monocyte dependence during homeostasis but is almost fully replaced by monocytes after MI (Dick et al., 2019).

Figure 2. Cardiac Mo/MΦ populations.

(A) t-SNE plot showing extracted Mo/MΦ populations. (B) Expression of select immune cell markers as visualized in box-plots and t-SNE plots. Arrows indicate Ccr2low sub-population of M2MΦ. (C) Dot-plot of top upregulated genes for each Mo/MΦ population where color indicates experimental conditions. (D) Heatmap of differentially expressed genes between Mo/MΦ populations with representative significant GO Biological Process terms.

Figure 2.

Figure 2—figure supplement 1. Mo and MΦ marker genes and Diffusion Map analysis.

Figure 2—figure supplement 1.

(A) Expression of selected genes marking Mo and MΦ sub-populations visualized using box and t-SNE plots. (B) Expression of H2-Aa (MHC-II) and Igf1 between sub-populations of M2MΦ defined according to Ccr2 expression (Ccr2high and Ccr2low) as visualized in box-plots. (C) Major Mo/MΦ populations visualized on diffusion components 1–3.

Among other minor resident Mo/MΦ populations detected, the most abundant (pale green cells in Figure 1A,B) clustered with the M2 MΦ present at MI-day 7. In fact, all minor Mo/MΦ populations in sham hearts aligned with adult monocyte-derived Mo/MΦ populations which influx after MI (Figure 1A,B), consistent with recent findings (Dick et al., 2019). A prominent B-cell, and minor DC, T- and NK cell populations were also present in sham hearts. These populations may represent a mixture of resident cells and those involved in homeostatic immunosurveillance (Lavine et al., 2018), although we cannot exclude a response to sham operation.

At MI-day 3, a major influx population was identified as classical blood-derived M1 Mo, based on the signature Adgre1(F4/80)+Itgam(CD11b)+Fcgr1(CD64)+Ly6c2highCcr2highH2-Aa(MHC-II)low (M1Mo; Figure 2A–C) (Epelman et al., 2015; Swirski and Nahrendorf, 2018). Differentially expressed genes showed Gene Ontology (GO) term over-representation for cell migration, inflammation and T cell activation (Figure 2D). In FACS, Mo are distinguished from MΦ by having lower size and granularity, and lower levels of MΦ markers including Adgre(F4/80), Itgam(CD11b) and H2-Aa(MHC-II) (Bajpai et al., 2018; Hilgendorf et al., 2014). M1Mo identified at MI-day 3 were also low or negative for other MΦ markers Siglec1, Mrc1, Maf, Trem2 and Mertk, the latter involved in phagocytosis (Figure 2—figure supplement 1A) (Bajpai et al., 2018), and C1 complement genes C1qa, b and c (Figure 2C), which are involved (in addition to complement fixation) in recruitment of new inflammatory cells and protection against autoimmunity (Emmens et al., 2017; Thielens et al., 2017).

The more abundant population at MI-day 3 was identified as classical Mo-derived M1 MΦ based on the signature Ccr2highAdgre1(F4/80)+Ly6c2+H2-Aa(MHC-II)+ (M1MΦ; Figure 1A,B; Figure 2A–D). This assignment was supported by expression of the additional MΦ markers cited above, including Mertk and C1q, and hierarchical clustering, which showed M1MΦ to be most closely related to M1Mo (Figure 1C), as for human cognates (Bajpai et al., 2018). Differentially expressed genes showed GO term over-representation for leukocyte migration and responses to interleukin-1 (Figure 2D).

The most prominent population at MI-day 7 was identified as non-classical M2 MΦ involved in inflammation resolution and repair, with the signature Ccr2highAdgre1(F4/80)+H2-Aa(MHC-II)highLy6c2- (M2MΦ; Figure 1A,B; Figure 2A–D) (Epelman et al., 2015; Swirski and Nahrendorf, 2018). Differentially expressed genes showed GO term over-representation for antigen presentation via MHC class II (Figure 2D). As expected, the non-classical M2 MΦ population increased late during injury repair from <2% of TIP in sham and MI-day 3 hearts, to 16% at MI-day 7. Interestingly, the population dendrogram showed that M2MΦ were most closely related to MAC-TR (Figure 1C), and similarities between resident and subsets of infiltrating Mo/MΦ have been discerned recently using single-cell methods (Dick et al., 2019). Both MAC-TR and M2MΦ expressed Cx3cr1, often used to define tissue-resident MΦ (Figure 2B), and both upregulated pro-regenerative genes Igf1 (Figure 2B) and Pdgfb/c (Figure 2—figure supplement 1A). The majority of M2MΦ were Ccr2high (important for migration); however, a minor sub-populaion was Ccr2low (arrows, Figure 2B) and these expressed the highest levels of Igf1 and lower levels of MHC-II (Figure 2—figure supplement 1B). In this sense they are similar to the CCR2-MHC-IIlow subset of tissue-resident MΦ which appear to be yolk sac-derived (Dick et al., 2019; Leid et al., 2016), and which play a role through expression of IGF1 and IGF2 in remodeling of the fetal coronary vascular plexus (Leid et al., 2016). However, whether in the adult post-MI heart they represent persisting resident cells or an infiltrating population that has matured into a MAC-TR-like MΦ state will require lineage mapping approaches.

Il10, associated with the anti-inflammatory functions of M2MΦ, was expressed in only few cells in our dataset and may be at the limit of detection (Figure 2—figure supplement 1A). Expression of mouse genes encoding cognates of human CD14 and CD16/FCGR3, previously used to define classical, non-classical and intermediate Mo in human blood (Villani et al., 2017), did not help to discrimate the above Mo/MΦ populations, nor did new markers recently highlighted from CyTOF analysis (Williams et al., 2018). Moreover, the M2MΦ marker Arg1 (encoding Arginase 1) was more lowly expressed in M2MΦ described here than in M1MΦ, consistent with findings that ARG1 does not always mark M2 cells (Jablonski et al., 2015). Neither the M2 MΦ, nor any other myeloid population, expressed Col1a genes, likely precluding the presence of myeloid-derived fibroblasts (Duerrschmid et al., 2015).

Diffusion Map (Angerer et al., 2016) analysis applied to model possible temporal (pseudotime) changes in major Mo/MΦ populations (Figure 2—figure supplement 1C) revealed a continuum of states resolved into a trajectory from early infiltrating M1Mo (left) and inflammatory M1MΦ (centre), to the late peaking M2MΦ (right), similar to a recent scRNA-seq study (Dick et al., 2019) and consistent with the current model in which M1 Mo differentiate into M2 cells in situ (Lavine et al., 2018; Hilgendorf et al., 2014). The Diffusion Map also demonstrated the convergence of M2MΦ present at MI-day 7 with tissue-resident MΦ (MAC-TR) in sham hearts, a relationship reflected in the population dendrogram (Figure 1C; see Discussion).

The minor myeloid populations also showed different expression profiles and dynamics (Figure 1A–C,E; Figure 2A–D). For example, MAC6 showed upregulation of granulocyte markers including S100a9 and Csf3r (Supplementary file 3), with sub-populations expressing markers of neutrophils (Ly6g) and eosinophils (Siglecf) (Figure 2—figure supplement 1A). MAC-IFNIC cells showed strong upregulation of interferon (IFN)-induced genes including Ifit3, Ifit1 and Cxcl10 (Figure 2C), consistent with GO term analysis implicating responses to IFN α, β, and γ (Figure 2D). These cells appear to arise from Ccr2+ MΦ as opposed to monocytes (Dick et al., 2019), and likely correspond to the recently described inflammatory MΦ subtype that has negative effects on heart repair after MI through promotion of inflammatory cell types, and cytokine and chemokine expression (King et al., 2017).

Cell-cell communication analysis in TIP

We constructed cell-cell communication networks with weighted edges reflecting expression fold-changes of ligands and receptors in source and target populations, respectively (Materials and methods). Ligand-receptor interactions were derived from a curated map of human ligand-receptor pairs (Ramilowski et al., 2015) with mouse-specific weights added after reference to the STRING database (Szklarczyk et al., 2017). Based on permutation testing of randomized network connections, 91 cell-cell relationships with weighted paths higher than expected by chance (Padj <0.01) were identified (Figure 3A–B). Myofibroblasts (MYO) and MΦ populations M1MΦ and MAC-8 exhibited the largest number of outbound connections, with MYO having the highest weight. ECs by far showed the largest number and weighting of significant inbound connections. Strikingly, fibroblast populations (F-SH, F-SL, F-Act and F-WntX) appeared to communicate exclusively with vascular (ECs and mural) and glial cells. In line with this result, immunofluorescence analysis of sham and MI-day 3 hearts showed that Pdgfra-GFP+ fibroblasts were observed in close spatial relationships or direct contact with CD31+ endothelial cells (Figure 3—figure supplement 1A–B).

Figure 3. Cell-cell ligand-receptor network analysis.

(A) Hierarchical network diagram of significant cell-cell interaction pathways. Arrows and edge color indicates direction (ligand:receptor) and edge thickness indicates the sum of weighted paths between populations. (B) Comparison of total incoming path weights vs total outgoing path weights across populations. (C) Summed ligand weights across souce ligand and receptor target paths for top ligands in MYO. (D) Tree plot showing outgoing connections from the Glial cells. Top node refers to source population, second layer to ligands, third layer to receptors and leaf nodes represent target populations.

Figure 3.

Figure 3—figure supplement 1. Pdgfra-GFP+ cells localization in healthy and diseased hearts.

Figure 3—figure supplement 1.

(A–B) Representative image of GFP (green) and CD31 (red) co-immunostainings in sham (A) and MI-day 3 (B) hearts. Arrowheads show GFP+ cells in close proximity/contact with CD31+ cells. Scale bar - 20 μm.

Top-weighted interactions involving MYO were driven mostly by ECM-associated ligands including COL1a1, COL1a2, Fibronectin 1, Pleiotrophin, COL3a1, COL5a2, Biglycan, Metalloprotease inhibitor one and COL5a1, that can engage with receptors expressed in numerous populations (Figure 3C). The minority glial cell population expressed canonical neuronal glia markers, including Plp1, Prnp and Gfra3 (Figure 1—figure supplement 1D), and likely support cardiac sympathetic nerves essential for cardiac regeneration in neonates (Mahmoud et al., 2015). Glial cells also appeared to communicate with the three EC populations and mural cells (Figure 3A), consistent with the phenomenon of neurovascular congruence in the cardiac sympathetic plexus (Stubbs et al., 2009). In support of this, we detected eight ligands highly specific to glial cells (expressed in <5% of other TIP cells) including Dhh (Desert Hedgehog) and Semaphorin genes Sema3b and Sema4f (Figure 3D), involved in both neural and angiogenic development (Gamboa et al., 2017). Thus, these maps suggest the extent, directionality and complexity of interactions between cardiac cell types in homeostasis and injury.

Hybrid populations in TIP

We detected five minor populations expressing markers of two lineages (Figure 1—figure supplement 4A–D). Such ‘hybrid’ cells may betray trans-differentiation events or doublets in proximity that are resistant to the conditions of dissociation. Microdroplet microfluidics platforms are also known to generate a significant number of doublets (Zheng et al., 2017a); thus, the provenance of hybrid cells requires independent validation.

ECs are highly plastic and endothelial-to-mesenchyme transition (EndMT) has been reported to generate fibroblasts after cardiac injury (Kovacic et al., 2019). Conversely, cardiac fibroblasts have been observed to transdifferentiate to ECs (Ubil et al., 2014), albeit that this has been disputed (He et al., 2017). The F-EC hybrid population co-expressed markers of fibroblasts and ECs, and segregated with other fibroblast populations (Figure 1—figure supplement 4A–C). To explore this population, we isolated interstitial cells from dissected ventricles of PdgfraGFP/+ mice after sham or MI surgery, and asked whether we could detect GFP+CD31+ cells using flow cytometry and a stringent gating strategy that excluded cell aggregates (Materials and methods; Figure 1—figure supplement 5A–D). We detected 2.4 ± 0.28% of GFP+CD31+ cells in sham hearts, 1.51 ± 0.26% in MI-day 7 hearts, and none in controls (Figure 1—figure supplement 6A–E) - thus, while double positive cells were found, they did not appear responsive to injury.

An ability of Mo/MΦ to transdifferentiate into endothelial-like cells in different settings has been documented in vitro and in vivo, and has therapeutic implications (Das et al., 2015), although a natural plasticity in Mo toward an endothelial cell fate in vivo does not have strong support (Basile and Yoder, 2014). The M2MΦ-EC hybrid population co-expressed markers of ECs (Kdr+Pecam1+Sox17+ Efnb+Mcam+) and M2MΦ (Ccr2highAdgre1[F4/80]+H2-Aa[MHC-II]highCx3cr1+Mrc1[CD206]+Ly6 c2-), and segregated with M2MΦ (Figure 1—figure supplement 4A,B). Flow revealed 0.56 ± 0.02% single live CD31+CD45+ cells in sham-day 7 hearts, increasing to 4.04 ± 1.03% in MI-day 7 hearts, demonstrating an increase in injury (Figure 1—figure supplement 7A–C). Among these, 35.67 ± 3.01% were F4/80+CD206+ (a signature of M1 and M2 MΦ) in sham hearts, increasing to 60.03 ± 4.60% in MI-day 7 hearts. It is well known that the expression of EC markers on the surface of bone-marrow-derived cells is insufficient to define them as ECs, although they can be angiogenesis promoting cells (Basile and Yoder, 2014). While these data do not exclude the possibility of doublet formation in our scRNA-seq experiments, they support the existence of distinct F-EC and M2MΦ-EC populations with hybrid qualities and different responses to injury. These warrant further investigation.

Single-cell RNA-seq of the Pdgfra-GFP+ cardiac fibroblast lineage

A major subset of fibroblasts in the uninjured adult murine heart express the cell surface stem/progenitor cell markers SCA1 and/or PDGFRα (Kanisicak et al., 2016; Asli et al., 2017; Chong et al., 2011; Noseda et al., 2015). However, when fibroblasts differentiate into MYO, they reduce these markers and express fibrogenic (e.g. Periostin; POSTN) and/or contractile (e.g. αSmooth Muscle Actin; αSMA) proteins (Fu et al., 2018). To circumvent the dominance of immune cells in TIP following MI, which dilute out other cell populations, and to focus on fibroblast sub-populations (Figure 1A), we performed single-cell expression profiling on PDGFRα+CD31- cardiac interstitial cells at days 3 and 7 post-sham or MI. GFP fluorescence from PdgfraGFP/+ mice was used as a surrogate lineage tracing tool and enabled us to capture both GFPhigh fibroblasts as well as their derivatives in MI mice, including MYO (Asli et al., 2017). We sorted for GFP+CD31- cells (Figure 4—figure supplement 1A), although did not use SCA1 as an index marker so as to capture the substantial Pdgfra-GFP+ fibroblast population that is negative or low for SCA1 expression (Figure 4—figure supplement 1B).

We performed unbiased clustering on an aggregate of the 16,787 cells (Materials and methods), identifying 11 sub-populations (Figure 4A–D; Figure 4—figure supplement 1C–F). The two sham conditions showed high concordance (Figure 4—figure supplement 1E,F) and are displayed merged (Figure 4A,B) unless indicated. All populations showed expression of canonical fibroblast markers Pdgfra, GFP, Ddr2 and Col1a1, albeit at varying proportions and levels (Figure 4E), and major changes in cell proportions were seen between conditions (Figure 4D). Here, we refer to ‘activated fibroblasts’ and myofibroblasts (MYO) as distinct cell entities, without prejudice about their stability, origin, fate or contractile nature.

Figure 4. Pdgfa-GFP+ scRNA-seq.

(A) t-SNE plot of GFP+ cells separated according to experimental condition (sham, MI-day 3, MI-day 7). (B) t-SNE plot showing aggregate of GFP+ cells across conditions. (C) Dendrogram of populations determined by average RNA expression in populations. (D) Percentages of cells in each population according to experimental condition. Stars indicate significant change across conditions according to DPA (p<0.01). (E) Expression of select genes in different populations as visualized in box and t-SNE plots. (F) Dot-plot of top five upregulated genes for each population where color indicates strength of expression and size of dot represents percentage of cells expressing the gene.

Figure 4—source data 1. Source data for quantification of colony counts summarized in Figure 4—figure supplement 2E.
DOI: 10.7554/eLife.43882.023

Figure 4.

Figure 4—figure supplement 1. FACS profiles and scRNA-seq analysis for GFP+/CD31- cells.

Figure 4—figure supplement 1.

(A) FACS plots and gating strategies for sorting GFP+/CD31- population at 7 days post-sham or MI surgery. (B) FACS plots showing SCA1 expression profiles in GFP+/CD31- cells at 7 days post-sham or MI surgery. (C) Boxplot showing number of UMIs detected per-cell across experimental conditions in GFP+/CD31- fraction after quality control filtering. (D) Boxplot showing number of genes detected per-cell across experimental conditions in GFP+/CD31- fraction after quality control filtering. (E) t-SNE visualization of scRNA-seq clusters for all sham and MI time-points in GFP+/CD31- cells. (F) Cell population proportions across all experimental conditions.
Figure 4—figure supplement 2. Comparison of GFP+ populations with FACS-sorted S+P+ cells.

Figure 4—figure supplement 2.

(A) ROC curves showing prediction accuracy of iRF classifier for predicting populations in sham cells. ROC calculated across a 10-fold cross-validation test. (B) Counts of iRF predicted cell identities across two scRNA-seq (Fluidigm) experiments on S+P+ cells. (C) Thy1 expression across conditions. (D) Distribution of Thy1.2 (CD90.2) protein expression by FACS in S+Pdgfra-GFP+cells. (E) Colony counts comparing S+Pdgfra-GFP+ with Thy1.2 fractions (low vs high) isolated by FACS (n = 3). Indicated p-values are derived from Student’s t-test.
Figure 4—figure supplement 3. Expression of selected marker genes across GFP+ populations visualized in box and t-SNE plots.

Figure 4—figure supplement 3.

Circles indicate relevant population for the displayed marker.
Figure 4—figure supplement 4. Population-specific expression of transcription factors marking fibroblast and myofibroblast sub-populations.

Figure 4—figure supplement 4.

Figure 4—figure supplement 5. Comparisons with Skelly et al. (2018) and Gladka et al. (2018) scRNA-seq data-sets.

Figure 4—figure supplement 5.

(A–C) Reanalysis of Skelly et al. scRNA-seq data. (A) t-SNE plot showing main fibroblast populations detected from clustering analysis and identities according to those described in GFP+. (B) iRF predictions of GFP+ sham populations on fibroblasts from Skelly et al. data-set. (C) Expression of representative marker genes according to box plots and as visualized on t-SNE plots. Shows fibroblast markers (Pdgfra, Ddr2, Col1a1) and sub-population-specific markers. (D) Expression of Ckap4 gene, which was identified by Gladka et al. as a novel marker of activated fibroblasts, as visualized in box and t-SNE plots across TIP and GFP+ sub populations.

In sham conditions, two major fibroblast populations could be distinguished on the basis of scRNA-seq. We termed these Fibroblast-Sca1-high (F-SH) and Fibroblast-Sca1-low (F-SL), as the highest upregulated gene in F-SH was Ly6a(Sca1) (Figure 4E; Supplementary file 4). F-SH contained the highest frequency of Pdgfra and Ly6a(Sca1)-expressing cells and likely corresponds to the PDGFRα+SCA1+ (S+P+) population previously defined by FACS (Pinto et al., 2016; Chong et al., 2011) (see also Figure 4—figure supplement 1A) and enriched in cardiac colony-forming mesenchymal stromal cell (MSC)-like cells (cCFU-F), which show multi-lineage differentiation and self-renewal in vitro (Chong et al., 2011; Noseda et al., 2015). In order to confirm the relationship between F-SH and S+P+, we performed deeper scRNA-seq on 103 FACS-purified S+P+ cells from uninjured wild-type mice using the Fluidigm platform and predicted cell identity using an iterative Random Forest (iRF) classifier (Basu et al., 2018) trained on populations defined in our GFP+ experiments in sham conditions using the Chromium platform (Materials and methods; Figure 4—figure supplement 2A). Approximately 60% of single S+P+ cells analyzed by Fluidigm were predicted to correspond to the F-SH population (Figure 4—figure supplement 2B), compared to <30% among total sham GFP+ cells (Figure 4D), indicating that S+P+ cells are significantly over-represented in F-SH cells (Fisher’s exact test, p=8.13e-11). We previously showed that cCFU-F are enriched in the S+P+ population (Chong et al., 2011; Noseda et al., 2015). THY1/CD90 is a recognised MSC marker, and Thy1 was upregulated in F-SH with high significance (p=4.48e-176; Figure 4—figure supplement 2C). Furthermore, FACS-isolated S+Pdgfra-GFP+CD90.2high cells isolated from healthy hearts showed a ~ 6 fold enrichment in cCFU-F compared to S+Pdgfra-GFP+CD90.2low cells (Figure 4—figure supplement 2D,E). Together, these results show that the F-SH population contains a subset of cells expressing Pdgfra, Ly6a(Sca1) and Thy1(Cd90) that is enriched in cCFU-F, highlighting the distinct expression signatures and functional properties of F-SH and F-SL.

We calculated differentially expressed (DE) genes between F-SL and F-SH in sham conditions (Supplementary file 5). F-SH was characterized by over-representation of genes involved in the biological process (BP) cell adhesion, which included cell surface receptor genes Ackr3(Cxcr-7), Thy1(Cd90), Axl and Cd34. In contrast, F-SL was characterized by GO BP terms signaling and signal transduction (Supplementary file 6). Within the signal transduction category, protein localization prediction with LocTree3 (Goldberg et al., 2014) indicated an over-represented majority (19/28) of secreted proteins (Fisher’s exact test, p=0.03), with 10/19 identified as ligands, including APOE, BMP4 and ADM. Thus, F-SL, a major sub-division of fibroblasts, has a unique secretory phenotype distinct from that in F-SH, which is enriched in MSC-like colony forming cells.

Novel Pdgfra-GFP+ fibroblast populations

We identified two previously unstudied GFP+ fibroblast populations termed Fibroblast - Wnt expressing (F-WntX) and Fibroblast - transitory (F-Trans). These were present in both sham and MI hearts, although had diminished significantly by MI-day 7 (DPA; p<0.01; Figure 4D). In F-WntX, differential gene expression analysis showed that the top upregulated gene was Wif1, encoding a secreted canonical WNT pathway inhibitor essential for cardiac repair after MI (Meyer et al., 2017). WIF1 can also antagonize Connective Tissue Growth Factor (CTGF) signaling (Surmann-Schmitt et al., 2012), which plays a supportive role in cardiac fibrosis (Travers et al., 2016). Wif1 was almost uniquely expressed in F-WntX in all conditions (Figure 4E). A single cell in the Fluidigm data corresponded to the F-Wntx population (Figure 4—figure supplement 2B). Multiple other WNT pathway-related genes were upregulated in F-WntX encoding WNT ligands (WNT5a, WNT16), soluble WNT antagonists (DKK3, SFRP2), membrane-bound WNT receptor (FRZB) and AXIN2, a component of the β-catenin destruction complex (see schematic in Figure 5A; Figure 4—figure supplement 3). F-WntX also showed upregulated Fmod, which inhibits fibrillogenesis and sequesters pro-fibrotic factor TGF-β within ECM (Zheng et al., 2014; Zheng et al., 2017b) (Supplementary file 4). Overall, this signature suggests an anti-WNT, anti-CTGF and anti-TGF-β extracellular and intracellular signaling milieu for F-WntX cells. F-WntX cells expressed Postn(Periostin), Acta2(αSMA), Tagln(Transgelin) and Scx(Scleraxis), in both sham and MI conditions (Figure 5B; Figure 5—figure supplement 1A), suggesting an activated state even in the absence of injury (Tallquist and Molkentin, 2017). The adjacent cluster, F-Trans, did not express activation markers, nor the WNT signature identified in F-WntX, or other uniquely identifying markers; however, cell trajectory analysis, described below, allowed us to assign F-Trans as a transitionary population between F-WntX and F-SL fibroblasts.

Figure 5. Features of the F-WntX population.

(A) Differentially expressed genes in F-WntX overlaid on Wnt pathway maps. Wnt node includes genes Wnt5a and Wnt16. (B) Detection rate, representing the percentage of cells expressing a gene, across all conditions for cells in F-WntX or all GFP+ cells combined. (C) Example GO BP terms over-represented (FDR < 0.05) in genes upregulated in F-WntX compared to F-SL/F-SH populations. (D) Tree plot showing ligand-receptor connections from F-WntX to EC sub-populations as calculated in TIP. Top node refers to source population, second layer to ligands, third layer to receptors and leaf nodes represent target populations. (E) Examples of F-WntX:EC ligand and corresponding receptor expression as visualised in box and t-SNE plots. For each ligand the corresponding receptor is immediately below.

Figure 5.

Figure 5—figure supplement 1. Activation and paracrine signature of F-WntX cells.

Figure 5—figure supplement 1.

(A) Expression of markers of activation visualized on condition-specific t-SNE plots. (B) GO Molecular Function terms over-represented (FDR < 0.05) in genes upregulated in F-WntX compared to F-SL/F-SH. (C, D) Venn diagram of overlap between F-WntX marker genes in ‘Regulation of cell proliferation’ category and MF categories ‘Signaling receptor binding’ (C) and ‘Growth factor binding’ (D). (E) Examples of ligands upregulated in F-WntX and corresponding receptors as expressed in TIP and visualized in box and t-SNE plots. For each ligand the corresponding receptor is immediately below. Circle indicates the location of F-WntX on the t-SNE plot.

We examined DE and GO BP terms in F-WntX compared with F-SL and F-SH combined. Notably, negative regulation of Wnt signaling pathway was over-represented in DE genes for F-WntX (Figure 5C), driven by WNT pathway antagonist genes Wif1, Dkk3, Frzb and Sfrp2, discussed above, as well as Apoe, Nkd1 and Wwtr1, also known to interact with the WNT pathway. BP terms related to negative regulation of development/differentiation, extracellular matrix (ECM) organization and signaling, were also significant. ECM organization terms were over-represented in the adjacent F-Trans (Supplementary file 6), involving genes also upregulated in F-WntX (e.g. Eln[Elastin], Vit[Vitrin] and Mfap4), and others upregulated in F-Trans but not F-WntX, including collagen genes (Col1a1, Col3a1 and Col14a1) and Fbln1(Fibrilin-1).

The top GO BP term in F-WntX was regulation of cell proliferation (Figure 5C); however, F-WntX did not express cell cycle markers under any condition. Analysis of Molecular Function (MF) terms revealed over-representation of signaling receptor binding and growth factor binding (Figure 5—figure supplement 1B), overlapping significantly with regulation of cell proliferation (Figure 5—figure supplement 1C,D; Fisher’s exact test, p<1e-03). In this latter term, there were several cytokine and chemokine genes, including Pdgfa, Tgfb3, Ptn, Ccl19 and Cxcl12, some of which bind receptors that were down-regulated in F-WntX, strongly suggesting paracrine functions related to their expression in F-WntX (Figure 5—figure supplement 1E). A paracrine function for F-WntX was supported by our ligand-receptor analysis, which indicated that F-WntX cells communicate most significantly with ECs (Figure 3A). Analysis of top upregulated ligands in F-WntX connecting to receptors in ECs (Figure 5D) identified several factors such as Ptn(pleiotrophin), Myoc(myocilin) and Timp3(TIMP metallopeptidase inhibitor 3). Here again, the corresponding receptor was expressed in ECs but downregulated in F-WntX (Figure 5E).

Localization and composition of WIF1+ cells

To explore the location of F-WntX cells and their behaviour after injury, we examined the expression of WIF1 protein in Pdgfra-GFP sham and MI hearts by immunofluorescence (IF), after first confirming that our chosen antibody detected known sites of Wif1 expression in E14.5 embryos (Figure 6—figure supplement 1). Interestingly, we detected WIF1 protein only in the infarct border zone at MI-day 3, but not in sham hearts or at MI-days 1 or 7 (Figure 6A–C, Figure 6—figure supplement 2). In cardiac cells (and some embryonic cells) WIF1 staining was perinuclear, and we demonstrated co-expression of WIF1 and the golgi marker GM130 (Figure 6D), consistent with WIF1 being a secreted protein. We found WIF1 in ~4% of total nuclei of the infarct border zone at MI-day 3, with a fraction of these (~5%) being GFP+ by IF, indicating a fibroblast identity (Figure 6C,I), and overall ~17% were positive for Ki67 (Figure 6E,I). WIF1+ cells were negative for CD31, and negative or very low for αSMA, with rare exceptions (Figure 6F,F’,H). However, WIF1 was also expressed in ~4% of total CD45+ cells in the infarct border zone (~15% of WIF1+ cells were CD45+) (Figure 6G,I). We observed frequent close proximity or contact between WIF1+ cells and CD31+ ECs in tissue sections (Figure 6H), in line with the predicted cell-cell ligand-receptor connection between F-WntX cells and ECs (Figure 3A). Such proximity was less obvious for α-SMA+ or CD45+ cells (Figure 6F,G). Overall, our data suggest that WIF1 expression is post-transcriptionally regulated and injury-dependent, appearing in the infarct border-zone at MI-day 3 in a subset of fibroblasts and immune cells. The temporal window of WIF1 expression overlaps with fibroblast activation and expansion, and the beginning of EC renewal and MYO differentiation.

Figure 6. WIF1 localization and co-expression in injured and uninjured hearts of Pdgfra-GFP+ mice.

(A) Representative image of WIF1 (red), GFP (green) and Wheat Germ Agglutinin (WGA, grey) co-immunostaining showing the border zone at MI-day 3. Arrowheads show WIF1+ cells. Scale bar - 50 μm. (B–C) Representative images of WIF1 (red) and GFP (green) co-immunostainings showing left ventricle (sham, (B) or infarcted border zone at MI-day 3 (C). Arrowheads show WIF1+ cells, Asterix shows WIF1+GFP+ cells. Scale bars - 20 μm. (D–G) Representative images of co-immunostainings for WIF1 (red), GFP (green) and markers (gray) for golgi (GM130, D, Arrowheads show WIF1+GM130+ cells), proliferation (Ki67, E, Arrowheads show WIF1+Ki67+ cells), smooth muscle cells and myofibroblasts (α-SMA, F, (F’) showing an example of a WIF1+GFP+α-SMA+ cell from another section), and leukocytes (CD45, G, Arrowheads show WIF1+CD45+ cells, Asterix shows GFP+CD45+ cells). Scale bar - 20 μm. (H) Representative image of co-immunostaining for WIF1 (green), WGA (gray) and endothelial cell marker CD31 (green). Arrowheads show WIF1+ cells in close proximity/contact with CD31+ cells. Scale bar - 20 μm. (I) Quantification of marker-positive cells in the infarcted border zone of MI-day 3 hearts. n = 4.

Figure 6—source data 1. Source data for quantification of marker-positive cells summarized in Figure 6I.
DOI: 10.7554/eLife.43882.029

Figure 6.

Figure 6—figure supplement 1. WIF1 expression pattern in E14.5 embryos.

Figure 6—figure supplement 1.

Representatives images of sagittal sections of E14.5 embryos immunostained with WIF1 (red) antibody. Scale bar is 50 μm.
Figure 6—figure supplement 2. WIF1 protein expression after sham or MI.

Figure 6—figure supplement 2.

(A–D) Representative images of WIF1 (red) and GFP (green) co-immunostainings in the left ventricle (sham, A) or in the infarcted border zone of hearts, 1 (B), 3 (C) or 7 (D) days after surgery. Arrowheads show WIF1+ cells, Asterix shows WIF1+GFP+ cell. Scale bar - 20 μm.

Flux of fibroblasts and myofibroblasts after MI

MI is associated with appearance of activated fibroblasts and myofibroblasts. MI-day 7 was characterized by the appearance of a large population of myofibroblasts (MYO), representing 52% of GFP+ cells at MI-day 7 in our data (Figure 4A–D). MYO showed strong upregulation of numerous collagen genes (e.g. Col1a1, Col3a1, Col5a2), as well as Postn (99.5%) and Acta2 (61%) at high levels, indicative of an activated state and suggestive of a contractile phenotype for a subset of cells (Tallquist and Molkentin, 2017; Travers et al., 2016) (Figure 4E; Figure 4—figure supplement 3). Upregulated genes involved in wound healing and cell migration included Fn(Fibronectin) and Cthrc1(Collagen Triple Helix Repeat Containing I), the latter representing a highly specific marker for MYO (Figure 4F). MYO showed decreased expression of Pdgfra, Pdgfra-GFP, Ly6a(Sca1), Thy1(Cd90) and Cd34 (Figure 4E; Figure 4—figure supplement 2C; Supplementary file 4), indicating loss of stem/progenitor cell markers.

Earlier in the injury process, there was a distinct increase in a population with a signature consistent with activated fibroblasts (Fibroblast: activated; F-Act). These represented 48% of GFP+ cells at MI-day 3, before diminishing to 12% at MI-day 7 (Figure 4A–D). F-Act expressed Postn at high levels in ~80% of cells (Figure 4E) consistent with an activated state (Tallquist and Molkentin, 2017). They expressed Acta2 in 28% and 35% of cells at MI-day 3 and MI-day 7, respectively, although at much lower levels compared to MYO, suggesting an emerging contractile phenotype in some cells. On the population dendrogram, F-Act was most closely related to fibroblast populations (F-SH, F-SL, F-Trans) and was more distant from MYO (Figure 4C). Whereas F-Act expressed few genes that could be considered highly specific, the top upregulated gene was Cilp (Figure 4F), encoding a matricellular protein and inhibitor of TGF-β1 signaling, consistent with F-Act being a pre-MYO population in which fibrosis is constrained. The expansion of F-Act at MI-day 3 correlated with a decrease in the proportion of F-SH and F-SL, whereas the diminishment of the F-Act population at MI-day 7 coincided with the appearance of MYO and an apparent partial restoration of F-SH and F-SL cells (Figure 1A,E; Figure 4A,D; see Discussion).

A distinct GFP+ population contained fibroblasts undergoing proliferation (Fibroblast - cycling; F-Cyc), comprising 15% of GFP+ cells at MI-day 3 and 3% at MI-day 7 (Figure 4D), consistent with studies showing peak fibroblast proliferation at MI-days 2–4 (Fu et al., 2018; Ivey et al., 2018). F-Cyc uniquely expressed a strong cell cycle gene signature, including Ccnb2(CyclinB), Cdk1(Cyclin dependent kinase 1) and Mki67(Ki67) (Figure 4E; Figure 4—figure supplement 3), and expressed both Postn (88%) and Acta2 (76%) at high levels (Figure 4E; see below).

Cell trajectory analysis of Pdgfra-GFP+ cells

To look at potential relationships between the major GFP+ populations, we analyzed cell trajectories using Diffusion Maps (Materials and methods). MYO, F-WntX and F-Cyc were represented as three different trajectories along diffusion components 1, 2 and 3, respectively (Figure 7A), with the root containing the two large unactivated fibroblast populations F-SH and F-SL, which were most prominent in sham hearts. F-Trans was an intermediary population along the trajectory to F-WntX, and F-Act was an intermediary population for both F-Cyc and MYO branches. F-Cyc, characterized by expression of a strong cell cycle gene signature, was represented most strongly at MI-day 3, whereas MYO was exclusively associated with MI-day 7 (Figure 7B,C). These data suggest that F-Act expands by proliferation up to MI-day 3 (F-Cyc trajectory) and differentiates to MYO during the transition from MI-day 3 to MI-day 7. The presence of some F-Cyc-like cells between the F-Cyc and MYO trajectories at MI-day 7 raises the possibility that a small fraction of F-Act cells differentiate rapidly into MYO after or during division (see below).

Figure 7. Diffusion Map analysis of GFP+ cells.

Figure 7.

(A) 3D Diffusion Map of main fibroblast/myofibroblast populations with cells colored according to population. (B) 2D Diffusion Map facetted according to experimental condition. (C) Expression of marker genes on main trajectories of diffusion components across conditions. (D) Heatmap of differentially expressed genes with representative GO Biological Process terms.

We examined DE and GO BP terms in F-Act, F-Cyc and MYO compared with F-SL and F-SH combined, across all conditions (Figure 7D; Supplementary file 5; Supplementary file 6). DE genes for F-Act were over-represented in terms for collagen fibril organization (including several collagen genes) and regulation of wound healing. Many of these genes were also expressed in F-Cyc and MYO; however, there was an additional large gene signature strongly upregulated in MYO compared to F-Act (Figure 7D). DE genes for MYO demonstrated GO BP terms for collagen fibril organization and cell adhesion, containing collagen genes Col3a1, Col5a1, Col11a1 and Col14a1, and others involved in cell:cell and cell:matrix adhesion including Thbs1 (encoding Thrombospondin 1) and Fbn1. Other terms included angiogenesis and heart development as well as negative regulation of canonical Wnt signaling pathway, containing many genes previously identified in F-WntX.

Minor Pdgfra-GFP+ populations

Minor GFP+ populations included epicardial cells (EPI), observed only at MI-day 7. These expressed Wt1 (Figure 4—figure supplement 3) and overall were related transcriptionally to dissected adult mouse epicardium (Bochmann et al., 2010) (Spearman’s correlation test, p=0.014, r = 0.26). These are likely to be epicardial-derived fibroblasts that arise after MI as the epicardium reactivates its developmental program (including Pdgfra expression) (Zhou et al., 2011). Consistent with a previous cardiac scRNA-seq analysis on uninjured hearts (Skelly et al., 2018), we did not detect epicardial cells in TIP data, suggesting that these cells are under-sampled in our experiments - this is likely technical as epicardial cells have been detected readily in single-nucleus RNA-seq (Hu et al., 2018).

The minor population F-IFNS (Fibroblast: Interferon stimulated), found in all conditions, was negative for Cd45 and expressed high levels of Col1a1 and other fibroblast markers (Figure 4E; Figure 4—figure supplement 3), demonstrating a fibroblast identity, and interferon-responsive genes (Figure 4F; Supplementary file 4) (Zhou et al., 2013). Other minor GFP+ populations were EC and MAC (Figure 4A–C), which had EC and MΦ identities, respectively (Materials and methods).

Transcription factors expressed in Pdgfra-GFP+ cells

Several TF genes expressed in GFP+ cells may drive differentiation or responses to environmental stimuli (Figure 4—figure supplement 4). Scleraxis, already mentioned, was expressed in F-WntX cells and MYO. The basic helix-loop-helix factor gene, Tcf21, an accepted marker of cardiac fibroblasts (Tallquist and Molkentin, 2017), was expressed in most GFP+ populations with the exception of F-WntX cells. T-box factor gene Tbx20, another fibroblast marker, was expressed across all GFP+ cells and upregulated in F-WntX. The homeodomain TF gene, Meox1, which is part of the cardiac fetal gene expression signature reactivated in injured hearts (Lu et al., 2018), showed upregulation in subsets of GFP+ cells, most prominently in activated populations (F-Act, F-Wntx, F-Cyc and MYO), and may drive the activated state. Csrp2a, Zfp385a and Hmgb2 expression was also restricted among populations, with Csrp2a and Zfp385a showing strikingly complementary patterns. Interestingly, the homeodomain TF gene, Prrx1, which in BM is expressed in a subset of mesenchymal cells with CFU-F and multi-lineage differentiation potential (Kfoury and Scadden, 2015), was expressed across all GFP+ populations in all conditions.

Activated fibroblast and myofibroblast sub-populations

We re-clustered the sham and MI datasets at days 3 and 7 individually (Figure 8A,B), and repeated Diffusion Map analysis of GFP+ fibroblast lineages (Figure 8C,D) (Materials and methods). Whereas most populations identified at day 3 directly corresponded to those identified in the aggregate analysis, we found that F-Cyc could now be sub-divided into two populations, with only one exhibiting a clear cell cycle signature (Figure 8E, Supplementary file 7; Supplementary file 8). An intermediary population (Fibroblast - Cycling Intermediate; F-CI) showed upregulation of fibroblast activation markers including Postn, Cthrc1 and Acta2 (Figure 8—figure supplement 1A), but did not express markers of cell cycle, suggesting that it represents an additional population of activated fibroblasts, potentially competent for cell cycle entry. F-CI also upregulated genes involved in protein translation, a signature absent in F-Act (Figure 8E). The translation signature was maintained, albeit in attenuated form, in F-Cyc. Based on an iRF classifier trained to predict MI-day 3 populations (Figure 8—figure supplement 1B), we found no corresponding F-CI cluster at MI-day 7, indicating its transient nature (Figure 8—figure supplement 1C,D). Diffusion Map analysis also lent weight to the hypothesis that F-CI is a transitory population between F-Act and F-Cyc at MI-Day 3 (Figure 8C).

Figure 8. Time-point-specific analysis of GFP+ scRNA-seq.

(A,B) t-SNE visualization of GFP+ populations 3 days post sham/MI (A) and 7 days post sham/MI (B). (C,D) 3D Diffusion Map analysis of day 3 major populations (C) and day 7 major populations (D). (E) Heatmap of upregulated genes in day 3 injury-response populations. (F) Heatmap of differentially expressed genes between myofibroblast sub-populations. (G) Gene expression visualized in box and t-SNE plots for myofibroblast sub-population marker genes.

Figure 8.

Figure 8—figure supplement 1. Reanalysis of GFP-day 3 and GFP-day 7 data-sets separately.

Figure 8—figure supplement 1.

(A) Expression of F-CI marker genes on GFP-day 3 dot plot (B) AUC, sensitivity, specificity and precision for iRF classifier trained to predict GFP-day 3 cell populations after 10-fold cross-validation. (C) iRF prediction scores for GFP-day 7 cells as visualized on t-SNE plots. (D) Counts of GFP-day 7 cells predicted to correspond to GFP-day 3 populations as determined by iRF prediction (score >0.5). (E) Percentage of GFP-day 3 F-Cyc cells reassigned to alternative populations after removing cell cycle genes. (F) Percentage of GFP-day 7 F-Cyc cells reassigned to alternative populations after removing cell cycle genes. (G) 2D Diffusion Map plot of F-Act, MYO-1, MYO-2 and MYO-3 from the GFP-day 7 data-set.

We next removed genes annotated with the GO term ‘Cell Cycle’ and re-clustered populations. At MI-day 3, 88% of F-Cyc cells merged with F-CI, strongly supporting the hypothesis that F-CI is a pre-proliferative precursor of F-Cyc (Figure 8—figure supplement 1E). After removal of cell cycle genes at MI-day 7, 65% of F-Cyc cells remained as a distinct (proliferative) population, although 26% merged with F-Act and 6% merged with the MYO populations (Figure 8—figure supplement 1F). Taken together with Diffusion Map analysis, these data indicate that the closest population to the majority F-CI and cycling cells is F-Act rather than MYO, even though both express Acta2 and other MYO markers at high levels (Figure 4E). However, a minority of F-Cyc cells at MI-day 7 may be dividing MYO cells or in transition to MYO (see Discussion). Our data support the idea that F-CI cells are an activated form of fibroblast closely related to F-Cyc and derivative of F-Act. We hypothesize that they are primed for cell cycle entry and differentiation, but this requires further investigation.

In the MI-day 7 re-analysis, we found that MYO could be sub-divided into three clusters - MYO-1, MYO-2 and MYO-3 (Figure 8B,F; Supplementary file 9; Supplementary file 10), and comparing these clusters we noted genes corresponding to the contrasting functions of fibrosis inhibition and promotion. MYO-2 upregulated Tgfb1 (encoding TGF-β1), which is one of the strongest and most studied drivers of MYO formation (Figure 8G; Supplementary file 9), Scx(Scleraxis), which regulates ECM production and the myofibroblast phenotype downstream of TGF-β (Bagchi et al., 2016) and Thbs4(Thrombospondin 4), a regulator of cardiac fibrosis (Frolova et al., 2012). We sourced RNA-seq data from cultured mouse cardiac fibroblasts untreated or treated with TGF-β (Schafer et al., 2017) and extracted DE genes. As expected, the highest positive correlations with TGF-β treatment (log2 fold-changes) were in MYO and other activated fibroblast populations (Supplementary file 11). For inter-MYO comparisons, we found a significant positive correlation with TGF-β treatment in DE genes comparing MYO-2 v MYO-1 (Spearman’s correlation test, = 0.26, Padj = 1.93e-12) and MYO-3 v MYO-1 (Spearman’s correlation test, = 0.25, Padj = 1.75e-10), supporting the conclusion that MYO-2 and MYO-3 are pro-fibrotic.

In contrast, MYO-1 showed strong upregulation of anti-fibrosis genes included Wisp2, encoding matricellular protein CCN5 which can reverse established fibrosis (Jeong et al., 2016), and Sfrp2 encoding a soluble WNT receptor and antagonist of canonical WNT signalling, which is pro-fibrotic (Mirotsou et al., 2007) (Figure 8G). MYO-1 upregulated genes showed significant GO terms negative regulation of growth factor stimulus and negative regulation of cell proliferation (Figure 8F; Supplementary file 10), which included Sfrp1, implicated in inhibition of fibroblast proliferation and fibrosis (Sklepkiewicz et al., 2015), and Htra1 and Htra3, implicated in TGF-β signaling inhibition (Tocharus et al., 2004). Diffusion Map analysis of MYO sub-populations showed that MYO-1 and MYO-2 were distinct clusters with some overlap, suggesting a continuum of states, whereas MYO-3 did not appear to be a distinct population (Figure 8—figure supplement 1G).

Discussion

The mammalian heart is composed of a complex interdependent community of cells, although their interactions and flux are poorly characterized. Here, we present scRNA-seq data on >30,000 individual cardiac interstitial cells from sham, and MI days 3 and 7 hearts. Our interrogation of both the total interstitial population (TIP) and flow-sorted Pdgra-GFP+CD31- fibroblast lineage cells has given us a high-resolution map of cell lineage, state and flux in the healthy and injured heart, considerably extending preliminary studies (Kanisicak et al., 2016; Skelly et al., 2018; Gladka et al., 2018).

On the basis of these data, resident fibroblasts could be segregated into two major sub-populations denoted Sca1high (F-SH) and Sca1low (F-SL), both expressing canonical fibroblast markers such as Pdgfra, Pdgfra-GFP, Ddr2 and Col1a1. F-SH cells were enriched in S+P+ (SCA1+PDGFRα+) fibroblasts and clonal colony-forming units (Pinto et al., 2016; Chong et al., 2011; Noseda et al., 2015), and F-SH and F-SL showed distinct adhesive and secretory phenotypes, highlighting the likely functional differences between them.

We describe a novel activated fibroblast population, F-WntX, in sham hearts, which persists after MI. The related F-Trans is an intermediary population between F-SL and F-WntX (Figure 9). Aside from a low proportion of activated fibroblasts (F-Act) present in sham hearts, no other sham population (>75% of GFP+ cells) showed an activated phenotype. When we re-analyzed the scRNA-seq data of Skelly et al. on interstitial cells from uninjured hearts (Skelly et al., 2018), we identified all of the main fibroblast populations that we describe here in sham hearts (Figure 4—figure supplement 5A–C). F-Act and F-WntX expressed activation marker Postn in both studies; however, in contrast to our study, all populations identified by Skelly et al., including endothelial and immune cells, expressed the contractile marker Acta2. The reason for this is unclear, but is likely technical. Gladka et al. profiled cardiac populations after ischemia-reperfusion injury using scRNA-seq and highlighted Ckap4 as a novel marker of activated fibroblasts (Gladka et al., 2018); however, our data shows Ckap4 expression across virtually all cardiac stromal populations (Figure 4—figure supplement 5D), a discrepancy that may relate to the relatively low number of cells profiled in the Gladka et al. study.

Figure 9. Schematic summary of the flux and pseudotime differentiation dynamics of GFP+ populations between sham, MI-day 3 and MI-day 7.

Figure 9.

Populations are ordered in pseudotime from unactivated (top) to most activated/mature (bottom). Arrows connecting populations indicate direction of proposed differentiation/pseudotime trajectory. Colored arrows indicate whether the population appears to expand or diminish relative to the previous time-point.

Among fibroblasts, F-WntX uniquely expressed Wif1, encoding a canonical and non-canonical WNT signaling antagonist (Meyer et al., 2017; Bányai et al., 2012). WNT pathways play complex roles in cardiac biology and disease, impacting immune, vascular and pro-fibrotic pathways, and many drugs inhibiting WNT signaling are under investigation for their impacts on heart repair (Foulquier et al., 2018; Palevski et al., 2017). WIF1 additionally inhibits signaling through CTGF, a polyfunctional matricellular protein and positive driver of fibrosis (Travers et al., 2016; Surmann-Schmitt et al., 2012). WIF1 is a tumor suppressor inhibiting tumour angiogenesis through both WNT and VEGF pathways (Ko et al., 2014; Hu et al., 2009). Wif1 knockout mice show inhibition of Mo differentiation and abnormal chamber remodeling after MI (Meyer et al., 2017), while un-regulated transgenic WIF1 expression causes dilated cardiomyopathy (Lu et al., 2013), collectively indicating that correctly regulated WIF1 positively contributes to cardiac repair. One source of WNT proteins is inflammatory macrophages, and myeloid-specific deletion of the essential WNT transporter Wntless leads to improved cardiac functional recovery after MI involving an increase in reparative M2-like macrophages and angiogenesis (Palevski et al., 2017). In addition to WIF1, F-WntX cells showed upregulation of other WNT and TGFβ pathway antagonists (Figure 5A), overall flagging F-WntX cells as paracrine mediators of an anti-WNT/CTGF/TGFβ signaling milieu essential for cardiac repair.

WIF1 protein expression occurred in the border zone at MI-day 3, but not in sham or MI-days 1 and 7 hearts, consistent with previous findings (Meyer et al., 2017). We also detected WIF1 in ~4% of CD45+ immune cells infiltrating the border zone. We acknowledge that our IF studies may have underestimated the number of WIF1+ cells; for example, if we were unable to detect cells actively secreting WIF1 but lacking protein accumulation in the golgi. Contrary to our results, Meyer et al. showed WIF1 levels peaking at MI-day 1 and diminished by MI-day 7 using western blotting (Meyer et al., 2017). Whereas these differences need resolving, en face our data suggest that WIF1 expression is post-transcriptionally regulated with a peak around MI-day 3, consistent with our proposed function for WIF1 and the F-WntX population generally in inhibiting fibrosis and angiogenesis, and promoting differentiation of Mo, during the transition between the inflammatory and fibrotic phases of heart repair.

At MI-day 3, both F-SH and F-SL were significantly diminished, and we hypothesize that this occurs as they convert to an activated state (F-Act; Postn+Acta2negative-low). The scale of conversion suggests that fibroblasts remote from the infarct also become activated. Whereas F-Act upregulated genes that were associated with collagen fibrils and wound healing in common with MYO, F-Act cells are by far more closely related to resident fibroblasts than to MYO. A proportion of F-Act cells were actively proliferating at MI-day 3 and to a lesser extent at MI-day 7, consistent with the known peak of fibroblast proliferataion (Fu et al., 2018; Ivey et al., 2018). At MI-day 3, we also identified an activated, non-proliferating fibroblast population (F-CI) situated between F-Act and F-Cyc in trajectory plots, showing strong upregulation of genes supporting protein translation, and we hypothesize that this secondary activated state occurs in readiness for cell division and differentiation.

MYO was evident only at MI-day 7, where they represented >50% of total GFP+ cells. This may be an underestimate, as the Chromium microfluidic device biases against larger cells, and some MYO cells may be GFP-. Highlighting the limitations of using the contractile marker αSMA to define MYO (Tallquist and Molkentin, 2017), a proportion of F-WntX, F-Act, F-Cyc and MYO cells expressed its gene, Acta2. MYO cells massively upregulated a distinct network of genes related to cell adhesion, collagen fibril organization and angiogenesis, consistent with their known roles (Shannon et al., 2003Tallquist and Molkentin, 2017). Top ECM genes in these categories such as Postn, Fn1 and Col8a1 were expressed in virtually all MYO cells. However, our stage specific analysis of GFP+ cells at MI-day 7 allowed us to discern three distinct MYO sub-populations expressing pro-fibrotic (MYO-2 and MYO-3) or anti-fibrotic (MYO-1) states. MYO-1 expressed anti-TGF-β, anti-WNT and anti-proliferative signatures. Consistent with the view that MYO differentiation involves multiple cellular states, Fu et al. (2018) recently described a population of fibrobast-derived 'matrifibrocytes', quiescent cells which persist in the mature scar after MI.

The three major trajectories predicted by our Diffusion Map pseudotime analysis offer new insights into cardiac homeostasis and repair. The directionality of trajectories is suggested by the fact that all appear rooted in the resident F-SH and F-SL fibroblasts. One major trajectory arises specifically from F-SL and transits through F-Trans to F-WntX as the terminal state (Figure 9). Neither F-Trans nor F-WntX proliferate and may be primed for involvement in an injury response without the need for expansion. Up to MI-day 3, F-Act cells proliferate, but do not differentiate to MYO, showing that these events can be uncoupled. Differentiation to MYO occurs after MI-day 3, and our trajectory data suggest that this is largely from non-proliferating F-Act cells. We found no evidence for significant proliferation of MYO, although earlier time points need to be analyzed.

Resident fibroblasts F-SH and F-SL were depleted at MI-day 3, likely as a result of their activation and proliferation, and showed an apparent restoration by MI-day 7 in both TIP and GFP+ cells (falling just short of our stringent p-value of 0.01 for DPA) (Figure 1A,E; Figure 4A,D; Figure 8A,B). A caveat of pseudotime trajectories is that they may be bi-directional. If confirmed, the restoration of F-SH and F-SL between MI-days 3–7 (after the main fibroblast proliferative period [Fu et al., 2018; Ivey et al., 2018]) points to renewal involving phenotypic regression. Regression of myofibroblasts to a less activated state was proposed recently (Kanisicak et al., 2016), although our data suggest that the Postn-Cre lineage tracing mice used in that study to mark myofibroblasts labels most if not all F-Act cells (Figure 4E). Certainly, the mechanism of activation, proliferation, self-renewal, differentiation and de-differentiation of cardiac fibroblasts warrants deeper investigation.

The Mo/MΦ lineages of the heart have diverse functions, including in immunity, removal of debris and protection against autoimmunity during the early phases of MI, while promoting repair in latter phases, remodeling of the fetal coronary tree, neonatal heart regeneration and atrioventricular conduction (Lavine et al., 2014; Hulsmans et al., 2017; Leid et al., 2016; Aurora et al., 2014; Nahrendorf et al., 2007). The adult heart, like other organs, contains resident MΦ, some of which have their origins in erythromyeloid progenitors in the embryonic yolk sac and fetal liver (Leid et al., 2016) and which self-renew during homeostasis and injury (Heidt et al., 2014; Epelman et al., 2014; Dick et al., 2019). Other resident populations have differing degrees of monocyte dependence and some may eventually be supplanted during injury and aging by blood-born monocytes (Molawi et al., 2014; Dick et al., 2019). These resident MΦ likely have roles in defense against infection, antigen presentation and stimulation of T-cell responses, as well as efferocytosis (Epelman et al., 2014). In the neonatal heart, they are though to be essential for regeneration through stimulation of CM proliferation and angiogenesis (Lavine et al., 2014), and in the adult may be important for limiting fibrosis (Lavine et al., 2014; Dick et al., 2019), functions similar to pro-reparative M2 MΦ.

Our scRNA-seq data identified most of the known Mo/MΦ populations highlighted by targeted scRNA-seq of cardiac Mo/MΦ cells (Dick et al., 2019), including Ccr2+ and Ccr2- tissue resident MΦ, pro-inflammatory M1 Mo and MΦ at MI-day 3, and non-classical M2 MΦ at MI-day 7, as well as several minor populations and inflammatory fibroblasts. These data, combined with communication maps, provide a preliminary framework for further analysis of their relationships and flux in homeostasis, different disease models and after therapeutic interventions. Our Diffusion Maps lineage trajectory shows a continuum of states from M1Mo through M1MΦ to M2MΦ across the injury response, consistent with the recognised plasticity of blood born Mo (Lavine et al., 2018; Nahrendorf and Swirski, 2016). However, whether these states are determined exclusively within the injury environment remains to be determined. Our trajectory also shows convergence of M2MΦ with tissue-resident MΦ (Figure 2—figure supplement 1C). An emerging theme in the field is the similarity between yolk sac-derived tissue-resident MΦ and subsets of blood-born MΦ that appear during the reparative phase of MI (Dick et al., 2019). However, the latter may not fully adopt the gene expression signature of resident cells, nor fully compensate for their proposed functions, as suggested in the context of genetic ablation of Cc3cr1+ tissue-resident MΦ, although this may relate to the timing of their deployment (Dick et al., 2019). Irrespective of whether such cells are identical, it is noteworthy that pro-repairative macrophages appear to have multiple developmental origins, as found for virtually all other major adult heart lineages including CMs, ECs, SMCs, fibroblasts and adipocytes.

Our scRNA-seq data can be mined for expression of genes implicated in other forms of heart disease. For example, of 167 genes proximal to single nucleotide polymorphisms implicated in GWAS studies in atrial fibrillation (AF) (Roselli et al., 2018), 119 (71%) showed expression in our TIP single-cell data (for examples see Figure 1—figure supplement 8), suggesting possible involvement of stromal cells in AF risk.

We have identified substantial non-linear dynamics in the interactive cell communities of the heart (Figure 9), which, in this new light, can be further analyzed with lineage and functional tools. There remains a compelling clinical and economic rationale for finding new therapies for controlling the inflammation and fibrosis that accompanies virtually all forms of adult heart disease (Gourdie et al., 2016). In the long term, scRNA-seq analysis of cardiac homeostasis and disease will provide new entry points for discovering novel drugs and interventions supporting heart repair and regeneration.

Materials and methods

Key resources table.

Reagent type
(species) or resource
Designation Source or reference Identifiers Additional information
Gene (Mus musculus) Pdgfra NCBI NCBI Gene ID: 18595,
MGI:97530
Strain, strain
background
(Musmusculus, C57BL/6J)
Wild type, WT The Jackson Laboratory,
Stock Number: 000664,
RRID:IMSR:JAX:000664
Strain, strain
background
(Musmusculus, C57BL/6J)
Pdgfratm11(EGFP)Sor; PDGFRaGFP/+ The Jackson Laboratory,
Stock Number: 007669,
PMID: 12748302
MGI:2663656
Antibody APC-conjugated Rat
monoclonal anti-mouse
PDGFRa (CD140a)
eBioscience 17-1401-81,
Clone APA5
(1:200)
Antibody PE-Cy7-conjugated Rat
monoclonal anti-mouse
CD31 (PECAM-1)
eBioscience 25-0311-82,
Clone 390
(1:400)
Antibody PE-conjugated Rat
monoclonal anti-mouse
Sca1 (Ly6A/E)
BD Pharmingen 553108,
Clone D7
(1:400)
Antibody APC-Cy7-conjugated Rat
monoclonal anti-mouse
CD45
BD Pharmingen 557659,
Clone 30-F11
(1:400)
Antibody PE-conjugated Rat
monoclonal anti-mouse
F4/80
eBioscience 12-4801-82,
Clone BM8
(1:400)
Antibody APC-conjugated Rat
monoclonal anti-mouse
CD206 (MMR) Antibody
BioLegend 141708,
Clone C068C2
(1:200)
Antibody Chicken polyclonal
anti-GFP
Abcam ab13970 (1:200)
Antibody Rabbit polyclonal
anti-Wif1
Abcam ab186845 (1:1000)
Antibody Rat monoclonal
anti-CD31
Dianova DIA-310,
Clone SZ31
(1:100)
Antibody Rat monoclonal
anti-CD45
BD Biosciences 553076,
Clone 30-F11
(1:100)
Antibody Mouse monoclonal
anti-aSMA
Sigma A2547,
Clone 1A4
(1:100)
Antibody Rat monoclonal
anti-Ki67
Dako M7249,
Clone TEC-3
(1:100)
Antibody Mouse monoclonal
anti-GM130
BD Biosciences 610822,
Clone 35/GM130
(1:400)
Antibody Goat polyclonal
anti-Chicken Alexa 488
Life Technologies A11039 (1:500)
Antibody Goat polyclonal
anti-Rabbit Alexa 555
Life Technologies A21429 (1:500)
Antibody Goat polyclonal
anti-Rabbit Alexa 680
Life Technologies A21109 (1:500)
Antibody Goat polyclonal
anti-Rat Alexa 555
Life Technologies A21434 (1:500)
Antibody Donkey polyclonal
anti-Mouse Alexa 594
Life Technologies A21203 (1:500)
Antibody APC-conjugated Rat
monoclonal
anti-mouse PDGFRa
(CD140a)
eBioscience 17-1401-81,
Clone APA5
(1:200)
Commercial
assay or kit
Chromium Single Cell
30 Library and Gel
Bead Kit v2
10x Genomics 120237
Commercial
assay or kit
Chromium Single
Cell A Chip Kit
10x Genomics 120236
Commercial
assay or kit
Chromium i7 Multiplex Kit 10x Genomics 120262
Commercial
assay or kit
Nextera XT DNA
Sample Preparation
Kit (96 Samples)
Illumina FC-131–1096
Commercial
assay or kit
Nextera XT Index Kit v2 Illumina FC-131–2001
Commercial
assay or kit
Fluidigm Single-Cell
Auto Prep IFC chip
(5–10 um)
Fluidigm 100–5759
Commercial
assay or kit
SMART-Seq v4 Ultra
Low Input RNA Kit for
the Fluidigm C1 System
Takara 635026
Commercial
assay or kit
NextSeq 500/550
High Output Kit v2
Illumina FC-404–2002
Other LIVE/DEAD Viability/Cytoxicity
Kit, for mammalian cells
Thermo Fisher
Scientific
L-3224
Software, algorithm CellRanger 10x Genomics https://support.10xgenomics.com/single-cell-gene-expression/software/downloads/latest
Software, algorithm STAR PMID: 23104886 https://github.com/alexdobin/STAR;
RRID: SCR_015899
Software, algorithm Bowtie 2 PMID: 22388286 http://bowtie-bio.sourceforge.net/bowtie2/index.shtml;
RRID:SCR_005476
Software, algorithm featureCounts PMID: 24227677 http://subread.sourceforge.net;
RRID:SCR_012919
Software, algorithm Seurat PMID: 29608179 https://satijalab.org/seurat/;
RRID: SCR_007322
Software, algorithm Destiny PMID: 26668002 https://bioconductor.org/packages/release/bioc/html/destiny.html
Software, algorithm PANTHER PMID: 27899595 http://www.pantherdb.org;
RRID:SCR_004869
Software, algorithm Iterative Random Forest PMID: 29351989 https://cran.r-project.org/web/packages/iRF/index.html
Software, algorithm Differential Proportion
Analysis
This paper Source code 1 Materials and methods:
Differential proportion analysis
Software, algorithm Cell communication
analysis
This paper Source code 1 Materials and methods:
Ligand-receptor networks

Murine model

8-12 weeks old male mice were used in all experiments unless otherwise stated. For the single-cell RNA sequencing experiment, mice had a H2B-eGFP fusion gene knock-in at the endogenous Pdgfra locus (Pdgfratm11(EGFP)Sor; PdfgraGFP/+).

Surgically induced myocardial infarction

To induce acute MI, mice were anaesthetized by intraperitoneal injection of a combination of ketamine (100 mg/kg) and xylazine (20 mg/kg), and intubated. Hearts were exposed via a left intercostal incision followed by ligation of the left anterior descending coronary artery. Sham operated mice underwent surgical incision without ligation. Hearts were harvested for paraffin-embedding, or FACS analysis at 3 or 7 days post-surgery, as indicated in Results.

FACS experiments

TIP were isolated as previously described (Chong et al., 2011). Briefly, hearts were minced and incubated in collagenase type II (Worthington, USA) at 37°C before filtering through 40 μm strainers. Cells were resuspended in red cell lysis buffer, followed by dead cell removal, immunostaining for 15 min on ice with fluorophore-conjugated antibodies and two times wash with FACS buffer (1x PBS containing 2% fetal bovin serum) before acquisition.

We employed very stringent gating strategies to exclude doublets in the FACS analysis: FSC-H vs FSC-A, FSC-H vs FSC-W and SSC-H vs SSC-W cytograms were used to discriminate and gate out doublets/cell aggregates during sorting or from the analysis (Figure 1—figure supplement 5). To account for the autofluorescence generated by MI, we used a wild-type MI mouse as control to set up the gating strategy (Figure 1—figure supplement 6).

For the TIP fraction, total DAPI-negative live single cells were sorted in FACS buffer. For the GFP+ fraction, GFP+CD31-cells were sorted from the DAPI-negative live single cells. For Fluidigm experiments, SCA1+PDGFRα+CD31-(S+P+) cells were isolated and sorted as described above. For each sample, at least 10,000 final gate events were collected and stored for the later analyses.

Colony formation assay

TIP cells were isolated as described and stained with indicated antibodies. FACS sorted primary cells were seeded at a clonogenic density of 50 cells/cm2 (500 cells per well of 6-well plate) and were cultured in α-Minimal Essential Medium (α-MEM) containing 20% FBS+1% Pen/Strep at 37°C in a humidified 2% O2 and 5% CO2 incubator, with medium changes every 2–3 days. After 8-day culture, colonies were rinsed with phosphate-buffered saline (PBS), fixed with 2% paraformaldehyde (PFA) and stained with 0.05% (v/v) crystal violet dye in water. Differences in colony number and size were evaluated by a two-tailed one-sample t-test to test for variability between individual samples.

Sectioning, immunohistochemistry and confocal microscopy

Paraffin sections

Hearts were fixed in 4% PFA for 24 hr and processed at the Garvan Histopathology Center using a Leica Peloris II - Dual Retort rapid tissue processor (Germany). 10 μm thick longitudinal sections were dewaxed in xylene and rehydrated in decreasing concentrations of ethanol before being washed in distilled water.

Cryo-sections

Hearts were fixed in 4% PFA for 2.5 hr and washed in PBS before being incubated in 30% w/v Sucrose/PBS overnight at 4°C. Tissues were embedded in Tissue-Tek (Sakura, Cat #4583) and frozen on dry ice. 8 μm thick longitudinal sections were prepared for immunohistochemistry.

Immunohistochemistry

For paraffin-embedded tissue, antigens were unmasked using Tris-EDTA buffer (Tris base 10 mM, EDTA 1 mM, Tween-20 0.05%) boiled for 20 min. Samples were washed in PBS and treated with 5% BSA, 0.1% Triton X-100 for 1 hr at room temperature. Primary antibodies were incubated overnight at 4°C: chicken anti-GFP (Abcam, Cat #ab13970, 1/200), rabbit anti-WIF1 (Abcam, Cat #ab186845, 1/1000), rat anti-CD31 (Dianova, Cat #DIA-310, Clone S231, 1/100), rat anti-CD45 (BD Biosciences, Cat #553076, Clone 30-F11, 1/100), mouse anti-αSMA (Sigma, Cat #A2547, Clone 1A4, 1/100), rat anti-Ki67 (Dako, Cat #M7249, Clone TEC-3, 1/100) and mouse anti-GM130 (BD Biosciences, Cat #610822, Clone 35/GM130, 1/400). Alexa-conjugated secondary antibodies were incubated 1 hr at room temperature (1/500): goat anti-chicken Alexa 488 (Life Technologies, Cat #A11039), goat anti-rabbit Alexa 555 (Life Technologies, Cat #A21429), goat anti-rabbit Alexa 680 (Life Technologies, Cat #A21109), goat anti-rat Alexa 555 (Life Technologies, Cat #A21434) and donkey anti-mouse Alexa 594 (Life Technologies, Cat #A21203), and counterstained with Hoechst 33258 for 10 min at room temperature (Invitrogen, Cat #H3569, 1/10000). Finally, samples were mounted with ProLong Gold antifade reagent (Invitrogen, Cat #P36934) and imaged using a Zeiss LSM700 upright confocal microscope (Germany).

Images were processed using the ImageJ software (NIH, USA) and a minimum of 2600 cells were quantified in the border zone of 4 infarcted hearts.

Single-cell transcriptomics platform: 10x Chromium

The single-cell library preparation relied on a commercially available droplet method, the 10x Genomics Chromium System (10x Genomics Inc San Francisco, CA). The number of cells loaded on the system was calculated based on the desired number of captured cells following manufacturer’s instructions. scRNA-seq libraries were generated following capture. Chromium GFP+ day 3 samples were sequenced on the Illumina HiSeq 2500 and the remainder on the Illumina NextSeq 500 high output.

Single cell microfluidics: Fluidigm C1

This experiment was performed using the Fluidigm C1 platform (Fluidigm, San Francisco, CA). Immediately after cell sorting and counting, cells were loaded on the integrated fluidics circuits (IFC) C1 chip. Each capture site was carefully examined under a Leica fluorescence microscope in bright field, Red and Green fluorescence channels for cell doublets and viability, and to ensure the capture rate was satisfactory before cell lysis and cDNA preparation. The reverse transcription was performed on the chip using Clontech SMARTer Ultra Low Input RNA Kit V4 (Takara-Clontech, USA). After running the C1 Fluidigm system, single cell RNA libraries were generated from 100 to 300 pg (picogram) of cDNA, using the low throughput Nextera XT DNA library prep kit (Illumina, USA). Individual barcoded libraries were sequenced by Illumina HiSeq 2500 (125bp).

Processing of 10x Genomics Chromium scRNA-seq data

Raw scRNA-seq data was processed using the 10x Genomics CellRanger software (version 1.3.0). The BCL files obtained from the Illumina NextSeq platform were processed to Fastq files using the CellRanger mkfastq program. The Fastq files were mapped to the mm10 version 1.2.0 reference, downloaded from the 10x Genomics website, with the sequence for H2B-eGFP appended to the reference. The CellRanger count program was run on individual Fastq data-sets from the different conditions. The aggr program was run to generate aggregate unique molecular identifier (UMI) count matrices for the following experimental data-sets analyzed in this study: (1) TIP: sham-day 3, MI-day 3 and MI-day 7; (2) GFP+: sham-day 3, sham-day 7, MI-day 3 and MI-day 7, (3) GFP+ day 3: sham-day 3 and MI-day 3; (4) GFP+ day 7: sham-day 7 and MI-day 7.

Filtering, dimensionality reduction and clustering of scRNA-seq data

Bioinformatics processing of the scRNA-seq data was performed in R (R Development Core Team, 2018) using the Seurat package (Butler et al., 2018) with figures primarily generated using ggplot2 (Wickham, 2009). R scripts containing the steps used for processing and clustering the data for each individual data-set (TIP aggregate, GFP+ aggregate, GFP+ day 3 and GFP+ day 7) are available in Source code 1. For all data-sets, initial quality control filtering metrics were applied as follows. Cells with fewer than 200 detected expressed genes were filtered out. Genes that were expressed in less that 10 cells were filtered out. In order to control for dead or damaged cells, cells with over 5% of raw UMIs mapping to mitochondrial genes were filtered out. To further control for potential doublets in our data-sets, we visualized the distribution of expressed genes and UMI numbers and filtered out cells which were clear outliers.

UMIs were normalized to counts-per-ten-thousand, log-transformed, and a set of highly variable genes was identified by gating for mean expression level and dispersion level in a per-data-set manner. The log-normalized data was scaled, with variation due to total number of UMIs regressed out using a linear model. Principal component analysis was run on the scaled data for the set of previously defined highly variable genes. In order to identify the number of principal components (PCs) to use for clustering, we ran the JackStraw procedure implemented in Seurat that identifies statistically significant PCs. Based on running the JackStraw procedure with 1000 permutations, we defined significant PCs as those up to p<0.001, which were used as input to the Seurat graph-based clustering program, FindClusters. We experimented with modifying the number of PCs used for clustering but found that varying the number of PCs used caused only minor impact on the clustering solutions. The resolution parameter for FindClusters, which determined the number of returned clusters, was decided on a per-data-set basis after considering clustering output from a range of resolutions. The cells and clusters were visualized on a t-SNE dimensionality reduction plot generated on the same set of PCs used for clustering.

We inspected the clusters for hybrid gene expression signatures that could indicate captured cell doublets, or a signature of stress/apoptosis that could indicate cells damaged during the process of cell sorting and capture in the microfluidics device. Within TIP, our initial clustering analysis returned 29 clusters; within these we identified five minor clusters exhibiting hybrid gene expression signatures: F-EC, M2MΦ-EC, EC-L1, EC-L2 and BC-TC (Figure 1—figure supplement 4A–D). The small size of these populations meant that we could not exclude the possibility of doublets; the five hybrid clusters were therefore removed and all subsequent analysis (e.g. differential expression analysis) was performed on the remaining 24 clusters. Within the GFP+ data-set, we identified one cluster where GO analysis suggested cells in a stressed state; these cells were removed prior to down-stream analyses of the GFP+ data-sets.

Clustering of GFP+ day 7 data-set

We found when clustering the GFP+ day 7 data-set that the clustering solutions, depending on the resolution parameter, tended to either under-cluster (too few clusters) or over-cluster (too many clusters) the data when compared to clustering solutions for the full GFP+ aggregate or GFP+ day 3 data-set. In order to achieve a clustering solution that was directly comparable to the GFP+ aggregate and GFP+ day 3 data, we first over-clustered the data then ran an iterative procedure to ‘collapse’ transcriptionally similar clusters. We first generated a dendrogram of cluster similarity using the Seurat BuildClusterTree program, which builds a phylogenetic tree by first calculating average RNA expression across clusters, then performed hierarchical clustering on a distance matrix calculated from the averaged RNA profiles. We then ran AssessNodes to identify the clusters that were the most transcriptionally similar according to the Random Forest out-of-bag error calculations. The most similar clusters were merged and the process was repeated. We found four iterations of the above procedure produced clustering results directly comparable to the GFP+ aggregate and GFP+ day 3 clustering solutions.

Processing and filtering of Fluidigm data

Fastq files were mapped to the Gencode mouse mm10 reference using STAR aligner (Dobin et al., 2013) (version 2.5.2a). Reads marked unaligned by STAR were mapped using Bowtie 2 (Langmead and Salzberg, 2012) (version 2.3.1) with parameters –local –very-sensitive-local. BAM files generated by STAR and Bowtie2 were merged and read counts generated on the merged BAM using the Subread featureCounts (Liao et al., 2014) program with parameters -p -T 12 t exon -g gene_name.

Count data was normalized to counts per-million (CPM) and transformed to Log2(CPM + 1). Percent of RNA mapped to mitochondrial genes per cell was calculated. We filtered out cells that had below 1 million reads or greater than 20% of reads mapped to mitochondrial genes. This yielded 52 cells for experiment 1 and 52 cells for experiment 2.

Rank-based classifier for comparing data-sets

In order to circumvent the difficulties associated with clustering small numbers of cells, we instead used classification to map cell population identities from our analysis of the Chromium GFP+ experiment to Fluidigm. Due to significant differences in sequencing depth between the experiments, we built an iterative Random Forest (iRF) classifier trained on relative gene rankings (i.e. ranking genes in each cell from highest to lowest expressed), hypothesising that using ranks instead of expression should alleviate some of the difficulties in comparing data-sets of different sequencing depths. The classifier was built as follows. As the Fluidigm S+P+ experiment was performed on healthy hearts, we first removed the injury conditions from the GFP+ data-set and retained the populations most representative of the sham conditions: F-SL, F-Act, F-SH, F-Trans and F-WntX. We re-calculated a set of 706 highly variable genes for GFP+ by gating for genes with higher dispersion and mean expression and took the overlap with expressed genes (CPM > 0 in at least 1 cell) in Fluidigm. The expression matrix of highly variable genes was converted to a rank matrix, which was used as training data for an iRF classifier with 1000 trees and 3 iterations. We evaluated the accuracy of the classifier with 10-fold cross-validation and receiver operating characteristic (ROC) analysis. We found high accuracy across populations with area under the ROC curve (AUC) over 0.95 for all populations (Figure 4—figure supplement 2A). The classifier was applied to the Fluidigm data by first converting the expression matrix (highly variable genes) to ranks and returning the most probable cluster assignment. This was done for the two individual Fluidigm experiments (Figure 4—figure supplement 2B).

Comparing GFP+ day 3 and GFP+ day 7 populations

We compared the clusters identified in the GFP+ day 3 to GFP+ day 7 data-sets using two approaches. First was to build a multi-class iRF classifier for the GFP+ day 3 populations as above but including both sham and MI conditions. The classifier was trained on a set of 914 overlapping highly variable genes between the data-sets. We found the iRF classifier maintained reliable prediction accuracy even when including the injury conditions as determined by cross-validation and evaluation with multiple metrics including AUC, sensitivity, specificity and precision (Figure 8—figure supplement 1B). The GFP+ day 7 expression matrix was then converted to ranks and the iRF was applied to obtain population probabilities for each cell (Figure 8—figure supplement 1C). We additionally counted the number of cells with iRF score >0.5 (Figure 8—figure supplement 1D).

Differential expression

For calculating DE, we first identified genes expressed in at least 25% of cells for at least one of the populations being compared and with an absolute log2 fold-change difference of 0.5 (including a pseudo-count of 1). We then assigned p-values using the ‘bimodal’ test for DE (McDavid et al., 2013) implemented in the Seurat FindMarkers program. A Bonferroni-adjusted p-value of 1e-05 was used to determine significantly DE genes.

Trajectory (Diffusion Map) analysis

Trajectory analysis was performed using Diffusion Maps implemented in the Destiny R package (Angerer et al., 2016). For the set of populations being tested, we took the top upregulated genes for each population and the corresponding log-normalized expression matrix was input to the DiffusionMap program with default parameters. A fold-change (log2) cutoff of 1 was used to select upregulated genes with the exception of the GFP+ day 7 inter-MYO analysis, where a cutoff of 0.5 was used. We tested the stability of the DiffusionMap output by altering the numbers of input genes (i.e. modulating fold-change cutoff and using the full set of highly variable genes) but found the resulting trajectories to be consistent.

Gene ontology testing

Over-representation of GO terms in gene lists was calculated using the PANTHER web-service (Mi et al., 2017). The set of expressed genes in the relevant experiment was used as background. A false-discovery rate cut-off of 0.05 was used to determine statistical significance.

Identification of additional minor GFP+ populations

The EC (Endothelial Cell) population, identified by Pecam1 (CD31) expression, were observed in all conditions (Figure 4—figure supplement 3) and their transcriptome showed a strong positive correlation with previously isolated adult cardiac CD31+ ECs (Pearson’s correlation test, p-value=3.6e-15, p=0.48) (Quaife-Ryan et al., 2017). Detection of ECs was surprising, given the FACS gates to exclude CD31+ cells. Thus, the EC cluster may have low cell-surface CD31.

The MAC population, unique to MI-day 3, were Cd45+Cd68+ MΦ (Figure 4—figure supplement 3) and likely correspond to the minority population of Pdgfra-GFP+CD45+ cells identified at day 5 post-MI (Asli et al., 2017). We were also able to identify Pdgfra-GFP+CD45+ cells in MI-day 3 samples, as shown in Figure 6G. Very few cells in EC and MAC clusters showed GFP expression (Figure 4E) and were likely detected because of GFP perdurance. It is unclear at present whether they derive from intra-cardiac or extra-cardiac GFP+ cells.

Differential proportion analysis

We developed an approach for detecting changes in population proportions across conditions (Source code 1). For some number of conditions to be compared, clustering is first performed on an aggregate of all cells across conditions. Cells are assigned two labels: a group (G) label representing experimental group/condition and a cluster label (L). A count table is generated for each cluster per-condition, which can be converted to a proportion table. We define a statistic for the differential proportion test, Δpj, as the difference in cluster proportions between two conditions; that is Δpj=p1j-p2j for some cluster j and corresponding proportions in experiments 1 and 2. This workflow is illustrated in Figure 1—figure supplement 3A.

We next construct a null distribution for Δpj by randomly permuting cluster labels L for some w proportion of n total cells. Specifically, w*n cells are randomly selected, and their cell-type labels are replaced by a random sub-sample of labels drawn from the labels from all the cells (sampling without replacement). A new count and proportion table is then generated from this randomized data (Figure 1—figure supplement 3B). This process is repeated t times, and the resulting Δpj across the randomized data forms the null distribution. We then calculate empirical p-values representing either an increase or decrease in Δpj such that

Pincrease=1ti=1tIΔpiΔpj
Pdecrease=1ti=1tIΔpiΔpj

Where I(•) is the indicator function. A final p-value, Pj is determined as the minimum of Pincrease and Pdecrease. The most important parameter that needs to be set for DPA is w, where lower values will trend towards a stricter test (fewer significant hits) and higher values trend towards higher numbers of significant hits. For the following tests, and throughout this paper, we use w = 0.1.

As a negative control experiment, we evaluated DPA on our two GFP+ sham experiments (Figure 1—figure supplement 3C), which would be expected to demonstrate no major differences in population composition. We compared the results of DPA (w = 0.1, t = 100,000) to performing Fisher’s exact tests. We found Fisher’s exact test returned 7/11 of the populations as having significantly different proportions between the two sham experiments with p<0.05 (Figure 1—figure supplement 3D). In contrast, DPA identified only one population with significant proportion change between conditions with p=0.03 (Figure 1—figure supplement 3D). As p=0.03 represented the most significant change between sham conditions, we used a conservative p-value cutoff of 0.01 for comparing injury time-points as presented in Results.

We further evaluated DPA using simulation testing. We first designed a simulation experiment involving two replicate scRNA-seq experiments performed on one biological system with 10 cells types; that is where the underlying cell proportions are the same in both experiments. We simulated varying degrees of noise (e.g. experimental error or biological variability) by introducing an error rate e. Error was introduced when creating a simulated profile of each cell-type per-experiment by randomly adjusting each proportion, p, by ±e*p with the proportions finally readjusted to sum to 1. We randomly drew 5000 and 3000 cells from experiment 1 and experiment 2, respectively and performed DPA and Fisher’s exact test for each population. The procedure was repeated 100 times with error rates of 0.01, 0.05, 0.1, 0.15 and 0.2. We compared specificity measurements between Fisher’s exact test and DPA and found that DPA consistently made fewer false-positive calls than Fisher’s exact test, with difference in specificity between the two methods increasing with higher error rates (Figure 1—figure supplement 3E).

We next simulated a control vs condition experiment with 10 cell populations, where six populations change proportionally in the condition and four remain the same. We performed simulated experiments including error as above, drawing 4000 and 6000 cells for the control and condition experiments, respectively. For this experiment we have both true changes (positives) and non-changes (negatives) and could therefore evaluate both true-positive and false-positive detection rates. We found that both Fisher’s exact test and DPA correctly identified all true population changes with a sensitivity of 1 across all error rates (Figure 1—figure supplement 3F). When considering false-positives, DPA again outperformed Fisher’s exact test in specificity and precision measurements (Figure 1—figure supplement 3G,H). These results demonstrate that while DPA can identify true proportion changes with comparable sensitivity to Fisher’s exact test, it better controls for the detection of false-positive changes.

Ligand-receptor networks

In order to represent cell-cell communication networks via ligand-receptor interactions, we implemented a directed, weighted network with four layers of nodes as follows. The top layer of nodes refers to a set of source cell populations, defined as the cell populations expressing ligands. The second layer of nodes represents the set of ligands expressed by the source populations. A weighted edge connects Source:Ligand where the weight is the Log2 fold-change of the ligand in the source compared to the remaining populations. The third layer of nodes is the receptor targets of the ligands. These are determined from a map of ligand-receptor pairs, collated using both known ligand-receptor interactions and interactions predicted though protein localization and protein-protein interaction (PPI) information (Ramilowski et al., 2015). As this map contains human ligand-receptor interactions, we add a weight using mouse-specific protein-protein associations from the STRING database (Szklarczyk et al., 2017); these are represented as values between 0 and 1. The fourth layer of nodes represents the target cell populations. Receptors are connected to target populations where they are expressed with weight again determined by Log2 fold-change. We did not normalize the three edge weights so as to ensure that gene expression provides the greatest weighting. A path weight connecting a source to target via a ligand:receptor pair is calculated as the sum of weights along that path.

For our analysis, we initially considered all ligand:receptor pairs expressed in at least 10% of cells in a population. We then built a network using all 24 sub-populations identified in the TIP data-set. In order to filter out downregulated ligand-receptor connections, we set a minimum path weight of 1.5. An overall weight describing the strength of connection between a source and target population, ws:t, could then be calculated by summing all path weights between the source and target.

In order to identify Source:Target connections that have significantly higher summed path weights than would be expected by chance we generated random networks as follows. Given a Source:Target connection we identified the number of total unfiltered paths (T), the set of paths with weight greater than 1.5 (Gs:t) and the subsequent summed path weights (ws:t) for Gs:t. We then generated random networks by retaining the Ligand:Receptor edges (i.e. PPI connections) but randomly selecting T number of Source:Ligand edges and Receptor:Target edges and re-calculating Gs:t and ws:t (i.e. sub-sampling the fold-changes). This process was repeated m = 100,000 times and empirical p-values were calculated as

Pw=1mi=1mIwiws:t

As considering all possible combinations of Source:Target paths yields a large number of tests, we adjusted p-values with the Benjamini-Hochberg correction method and considered all edges with adjusted Pw <0.01 to be significant. Significant Source:Target edges were visualized as a hierarchical graph using Cytoscape (Shannon et al., 2003) with edge thickness determined by ws:t. The code for performing the above analysis on the TIP scRNA-seq is available in Source code 1.

Acknowledgements

We thank Elvira Forte, Gurpreet Kaur (Millennium Science) and Vikram Tallapragada for technical assistance, Rob Salomon, Dominik Kaczorowski and other GWCCG staff for support on FACS and single cell RNA-Seq platforms, and Karen Brennan and Michelle Catwright for animal support.

Funding Statement

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Contributor Information

Robert E Nordon, Email: r.nordon@unsw.edu.au.

Richard P Harvey, Email: r.harvey@victorchang.edu.au.

Edward Morrisey, University of Pennsylvania, United States.

Harry C Dietz, Howard Hughes Medical Institute and Institute of Genetic Medicine, Johns Hopkins University School of Medicine, United States.

Funding Information

This paper was supported by the following grants:

  • University of New South Wales to Nona Farbehi.

  • National Heart Foundation of Australia 100848 to Joshua WK Ho.

  • National Health and Medical Research Council 1105271 to Joshua WK Ho.

  • Stem Cells Australia SR110001002 to Richard P Harvey.

  • National Health and Medical Research Council 1074386 to Richard P Harvey.

  • Fondation Leducq 15CVD03 to Richard P Harvey.

  • St. Vincent's Clinic Foundation 100711 to Richard P Harvey.

  • Fondation Leducq 13CVD01 to Richard P Harvey.

  • National Health and Medical Research Council 573707 to Richard P Harvey.

  • National Health and Medical Research Council 1118576 to Richard P Harvey.

  • New South Wales Cardiovascular Research Network to Richard P Harvey.

Additional information

Competing interests

Reviewing editor, eLife.

No competing interests declared.

Author contributions

Conceptualization, Resources, Data curation, Software, Formal analysis, Validation, Investigation, Visualization, Methodology, Writing—original draft, Project administration, Writing—review and editing.

Conceptualization, Resources, Data curation, Software, Formal analysis, Validation, Investigation, Visualization, Methodology, Writing—original draft, Project administration, Writing—review and editing.

Conceptualization, Data curation, Funding acquisition, Validation, Investigation, Visualization, Project administration, Writing—review and editing.

Investigation, Writing—review and editing.

Conceptualization, Data curation, Investigation, Visualization, Writing—review and editing.

Conceptualization, Data curation, Supervision, Validation, Investigation.

Conceptualization, Data curation, Software, Formal analysis, Supervision, Funding acquisition, Investigation, Methodology, Writing—review and editing.

Conceptualization, Supervision, Funding acquisition, Investigation, Project administration, Writing—review and editing.

Conceptualization, Supervision, Funding acquisition, Investigation, Visualization, Writing—original draft, Project administration, Writing—review and editing.

Ethics

Animal experimentation: This research was performed following the guidelines, and with the approval, of the Garvan Institute of Medical Research/St. Vincent's Animal Experimentation Ethics Committee (research approvals 13/01, 13/02, 16/03 and 16/10).

Additional files

Source code 1. R code for processing and clustering of scRNA-seq data-sets, differential proportion analysis and cell communication network analysis.
elife-43882-code1.zip (1.5MB, zip)
DOI: 10.7554/eLife.43882.034
Supplementary file 1. Differentially expressed genes across TIP sub-populations.
elife-43882-supp1.xlsx (839.8KB, xlsx)
DOI: 10.7554/eLife.43882.035
Supplementary file 2. Differential proportion analysis p-value results for TIP and GFP+ sub-populations.
elife-43882-supp2.xlsx (6.8KB, xlsx)
DOI: 10.7554/eLife.43882.036
Supplementary file 3. Differentially expressed genes between Mo/MΦ sub-populations in TIP.
elife-43882-supp3.xlsx (138.9KB, xlsx)
DOI: 10.7554/eLife.43882.037
Supplementary file 4. Differentially expressed genes across GFP+ sub-populations.
elife-43882-supp4.xlsx (217.5KB, xlsx)
DOI: 10.7554/eLife.43882.038
Supplementary file 5. Differentially expressed genes across GFP+ Diffusion Map trajectories.
elife-43882-supp5.xlsx (118.9KB, xlsx)
DOI: 10.7554/eLife.43882.039
Supplementary file 6. GO Biological Process terms associated with GFP+ trajectory differentially expressed genes.
elife-43882-supp6.xlsx (62.4KB, xlsx)
DOI: 10.7554/eLife.43882.040
Supplementary file 7. Differentially expressed genes from GFP+ day 3 injury response populations.
elife-43882-supp7.xlsx (48.3KB, xlsx)
DOI: 10.7554/eLife.43882.041
Supplementary file 8. GO Biological Process terms associated with GFP+ day 3 injury response populations according to Diffusion Map trajectory: F-Act, F-CI and F-Cyc.
elife-43882-supp8.xlsx (32.9KB, xlsx)
DOI: 10.7554/eLife.43882.042
Supplementary file 9. Differentially expressed genes between myofibroblast sub-populations in GFP+ day 7 scRNA-seq.
elife-43882-supp9.xlsx (23.4KB, xlsx)
DOI: 10.7554/eLife.43882.043
Supplementary file 10. GO Biological Process terms associated with myofibroblast sub-populations in GFP+ day 7 scRNA-seq.
elife-43882-supp10.xlsx (13.8KB, xlsx)
DOI: 10.7554/eLife.43882.044
Supplementary file 11. Spearman correlation test comparisons between TGF-β -treated cardiac fibroblast RNA-seq and GFP+ day 7 sub-populations.
elife-43882-supp11.xlsx (13.8KB, xlsx)
DOI: 10.7554/eLife.43882.045
Transparent reporting form
DOI: 10.7554/eLife.43882.046

Data availability

Sequencing data have been deposited in the ArrayExpress database at EMBL-EBI (www.ebi.ac.uk/arrayexpress) under accession codes E-MTAB-7376 and E-MTAB-7365.

The following datasets were generated:

Farbehi N, Patrick R, Dorison A, Xaymardan M, Wystub-Lis K, Janbandhu V, Ho JWK, Nordon RE, Harvey RP. 2018. Single-cell RNA-seq of mouse cardiac interstitial cells 3 and 7 days after sham or myocardial infarction injury. ArrayExpress database. E-MTAB-7376

Farbehi N, Patrick R, Dorison A, Xaymardan M, Wystub-Lis K, Janbandhu V, Ho JWK, Nordon RE, Harvey RP. 2018. Single-cell RNA-seq of Pdgfra+/Sca1+/Cd31- mouse cardiac cells. ArrayExpress database. E-MTAB-7365

The following previously published datasets were used:

Schafer S, Viswanathan S, Widjaja AA. 2017. Integrated target discovery screens identify IL11 as novel therapeutic target for fibrosis. Gene Expression Omnibus. GSE97117

Skelly DA, Squiers GT, McLellan MA, Bolisetty MT, Robson P, Rosenthal NA, Pinto AR. 2017. Single cell RNA-Seq of the murine non-myocyte cardiac cellulome. ArrayExpress database. E-MTAB-6173

Quaife-Ryan GA, Sim CB, Ziemann M, Kaspi A. 2017. Multicellular Transcriptional Analysis of Mammalian Heart Regeneration. Gene Expression Omnibus. GSE95755

Bochmann L, Sarathchandra P, Mori F, Lara-Pezzi E, Lazzaro D. 2010. Transcription profiling of mouse cardiac muscle and epicardium after left coronary artery ligation and sharm operation. ArrayExpress database. E-MEXP-2446

References

  1. Angerer P, Haghverdi L, Büttner M, Theis FJ, Marr C, Buettner F. destiny: diffusion maps for large-scale single-cell data in R. Bioinformatics. 2016;32:1241–1243. doi: 10.1093/bioinformatics/btv715. [DOI] [PubMed] [Google Scholar]
  2. Asli N, Xaymardan M, Forte E, Waardenberg AJ, Cornwell J, Janbandhu V, Kesteven S, Chandrakanthan V, Malinowska H, Reinhard H, Yang SF, Pickett HA, Schofield P, Christ D, Ahmed I, Chong J, Heffernan C, Li J, Simonian M, Bouveret R, Srivastava S, Mishra RK, Dhawan J, Nordon R, Macdonald P, Graham RM, Feneley M, Harvey RP. Pdgfrα signaling in cardiac fibroblasts modulates quiescence, metabolism and self-renewal, and promotes anatomical and functional repair. bioRxiv. 2017 doi: 10.1101/225979. [DOI]
  3. Aurora AB, Porrello ER, Tan W, Mahmoud AI, Hill JA, Bassel-Duby R, Sadek HA, Olson EN. Macrophages are required for neonatal heart regeneration. Journal of Clinical Investigation. 2014;124:1382–1392. doi: 10.1172/JCI72181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Bagchi RA, Roche P, Aroutiounova N, Espira L, Abrenica B, Schweitzer R, Czubryt MP. The transcription factor scleraxis is a critical regulator of cardiac fibroblast phenotype. BMC Biology. 2016;14:21. doi: 10.1186/s12915-016-0243-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bajpai G, Schneider C, Wong N, Bredemeyer A, Hulsmans M, Nahrendorf M, Epelman S, Kreisel D, Liu Y, Itoh A, Shankar TS, Selzman CH, Drakos SG, Lavine KJ. The human heart contains distinct macrophage subsets with divergent origins and functions. Nature Medicine. 2018;24:1234–1245. doi: 10.1038/s41591-018-0059-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bányai L, Kerekes K, Patthy L. Characterization of a Wnt-binding site of the WIF-domain of Wnt inhibitory factor-1. FEBS Letters. 2012;586:3122–3126. doi: 10.1016/j.febslet.2012.07.072. [DOI] [PubMed] [Google Scholar]
  7. Basile DP, Yoder MC. Circulating and tissue resident endothelial progenitor cells. Journal of Cellular Physiology. 2014;229:10–16. doi: 10.1002/jcp.24423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Basu S, Kumbier K, Brown JB, Yu B. Iterative random forests to discover predictive and stable high-order interactions. PNAS. 2018;115:1943–1948. doi: 10.1073/pnas.1711236115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bochmann L, Sarathchandra P, Mori F, Lara-Pezzi E, Lazzaro D, Rosenthal N. Revealing new mouse epicardial cell markers through transcriptomics. PLOS ONE. 2010;5:e11429. doi: 10.1371/journal.pone.0011429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Butler A, Hoffman P, Smibert P, Papalexi E, Satija R. Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nature Biotechnology. 2018;36:411–420. doi: 10.1038/nbt.4096. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Chan CK, Lindau P, Jiang W, Chen JY, Zhang LF, Chen CC, Seita J, Sahoo D, Kim JB, Lee A, Park S, Nag D, Gong Y, Kulkarni S, Luppen CA, Theologis AA, Wan DC, DeBoer A, Seo EY, Vincent-Tompkins JD, Loh K, Walmsley GG, Kraft DL, Wu JC, Longaker MT, Weissman IL. Clonal precursor of bone, cartilage, and hematopoietic niche stromal cells. PNAS. 2013;110:12643–12648. doi: 10.1073/pnas.1310212110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Chong JJ, Chandrakanthan V, Xaymardan M, Asli NS, Li J, Ahmed I, Heffernan C, Menon MK, Scarlett CJ, Rashidianfar A, Biben C, Zoellner H, Colvin EK, Pimanda JE, Biankin AV, Zhou B, Pu WT, Prall OW, Harvey RP. Adult cardiac-resident MSC-like stem cells with a proepicardial origin. Cell Stem Cell. 2011;9:527–540. doi: 10.1016/j.stem.2011.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. D'Uva G, Aharonov A, Lauriola M, Kain D, Yahalom-Ronen Y, Carvalho S, Weisinger K, Bassat E, Rajchman D, Yifa O, Lysenko M, Konfino T, Hegesh J, Brenner O, Neeman M, Yarden Y, Leor J, Sarig R, Harvey RP, Tzahor E. ERBB2 triggers mammalian heart regeneration by promoting cardiomyocyte dedifferentiation and proliferation. Nature Cell Biology. 2015;17:627–638. doi: 10.1038/ncb3149. [DOI] [PubMed] [Google Scholar]
  14. Das A, Sinha M, Datta S, Abas M, Chaffee S, Sen CK, Roy S. Monocyte and macrophage plasticity in tissue repair and regeneration. The American Journal of Pathology. 2015;185:2596–2606. doi: 10.1016/j.ajpath.2015.06.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Dick SA, Macklin JA, Nejat S, Momen A, Clemente-Casares X, Althagafi MG, Chen J, Kantores C, Hosseinzadeh S, Aronoff L, Wong A, Zaman R, Barbu I, Besla R, Lavine KJ, Razani B, Ginhoux F, Husain M, Cybulsky MI, Robbins CS, Epelman S. Self-renewing resident cardiac macrophages limit adverse remodeling following myocardial infarction. Nature Immunology. 2019;20:29–39. doi: 10.1038/s41590-018-0272-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, Gingeras TR. STAR: ultrafast universal RNA-seq aligner. Bioinformatics. 2013;29:15–21. doi: 10.1093/bioinformatics/bts635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Duerrschmid C, Trial J, Wang Y, Entman ML, Haudek SB. Tumor necrosis factor: a mechanistic link between angiotensin-II-induced cardiac inflammation and fibrosis. Circulation. Heart Failure. 2015;8:352–361. doi: 10.1161/CIRCHEARTFAILURE.114.001893. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Emmens RW, Wouters D, Zeerleder S, van Ham SM, Niessen HWM, Krijnen PAJ. On the value of therapeutic interventions targeting the complement system in acute myocardial infarction. Translational Research. 2017;182:103–122. doi: 10.1016/j.trsl.2016.10.005. [DOI] [PubMed] [Google Scholar]
  19. Ensan S, Li A, Besla R, Degousee N, Cosme J, Roufaiel M, Shikatani EA, El-Maklizi M, Williams JW, Robins L, Li C, Lewis B, Yun TJ, Lee JS, Wieghofer P, Khattar R, Farrokhi K, Byrne J, Ouzounian M, Zavitz CC, Levy GA, Bauer CM, Libby P, Husain M, Swirski FK, Cheong C, Prinz M, Hilgendorf I, Randolph GJ, Epelman S, Gramolini AO, Cybulsky MI, Rubin BB, Robbins CS. Self-renewing resident arterial macrophages arise from embryonic CX3CR1(+) precursors and circulating monocytes immediately after birth. Nature Immunology. 2016;17:159–168. doi: 10.1038/ni.3343. [DOI] [PubMed] [Google Scholar]
  20. Epelman S, Lavine KJ, Beaudin AE, Sojka DK, Carrero JA, Calderon B, Brija T, Gautier EL, Ivanov S, Satpathy AT, Schilling JD, Schwendener R, Sergin I, Razani B, Forsberg EC, Yokoyama WM, Unanue ER, Colonna M, Randolph GJ, Mann DL. Embryonic and adult-derived resident cardiac macrophages are maintained through distinct mechanisms at steady state and during inflammation. Immunity. 2014;40:91–104. doi: 10.1016/j.immuni.2013.11.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Epelman S, Liu PP, Mann DL. Role of innate and adaptive immune mechanisms in cardiac injury and repair. Nature Reviews Immunology. 2015;15:117–129. doi: 10.1038/nri3800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Foulquier S, Daskalopoulos EP, Lluri G, Hermans KCM, Deb A, Blankesteijn WM. WNT signaling in cardiac and vascular disease. Pharmacological Reviews. 2018;70:68–141. doi: 10.1124/pr.117.013896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Frolova EG, Sopko N, Blech L, Popovic ZB, Li J, Vasanji A, Drumm C, Krukovets I, Jain MK, Penn MS, Plow EF, Stenina OI. Thrombospondin-4 regulates fibrosis and remodeling of the myocardium in response to pressure overload. The FASEB Journal. 2012;26:2363–2373. doi: 10.1096/fj.11-190728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Fu X, Khalil H, Kanisicak O, Boyer JG, Vagnozzi RJ, Maliken BD, Sargent MA, Prasad V, Valiente-Alandi I, Blaxall BC, Molkentin JD. Specialized fibroblast differentiated states underlie scar formation in the infarcted mouse heart. Journal of Clinical Investigation. 2018;128:2127–2143. doi: 10.1172/JCI98215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Gamboa NT, Taussky P, Park MS, Couldwell WT, Mahan MA, Kalani MYS. Neurovascular patterning cues and implications for central and peripheral neurological disease. Surgical Neurology International. 2017;8 doi: 10.4103/sni.sni_475_16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Gladka MM, Molenaar B, de Ruiter H, van der Elst S, Tsui H, Versteeg D, Lacraz GPA, Huibers MMH, van Oudenaarden A, van Rooij E. Single-Cell sequencing of the healthy and diseased heart reveals Cytoskeleton-Associated protein 4 as a new modulator of fibroblasts activation. Circulation. 2018;138:166–180. doi: 10.1161/CIRCULATIONAHA.117.030742. [DOI] [PubMed] [Google Scholar]
  27. Goldberg T, Hecht M, Hamp T, Karl T, Yachdav G, Ahmed N, Altermann U, Angerer P, Ansorge S, Balasz K, Bernhofer M, Betz A, Cizmadija L, Do KT, Gerke J, Greil R, Joerdens V, Hastreiter M, Hembach K, Herzog M, Kalemanov M, Kluge M, Meier A, Nasir H, Neumaier U, Prade V, Reeb J, Sorokoumov A, Troshani I, Vorberg S, Waldraff S, Zierer J, Nielsen H, Rost B. LocTree3 prediction of localization. Nucleic Acids Research. 2014;42:W350–W355. doi: 10.1093/nar/gku396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Gourdie RG, Dimmeler S, Kohl P. Novel therapeutic strategies targeting fibroblasts and fibrosis in heart disease. Nature Reviews Drug Discovery. 2016;15:620–638. doi: 10.1038/nrd.2016.89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Greicius G, Kabiri Z, Sigmundsson K, Liang C, Bunte R, Singh MK, Virshup DM. PDGFRα+ pericryptal stromal cells are the critical source of Wnts and RSPO3 for murine intestinal stem cells in vivo. PNAS. 2018;115:E3173–E3181. doi: 10.1073/pnas.1713510115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. He L, Huang X, Kanisicak O, Li Y, Wang Y, Li Y, Pu W, Liu Q, Zhang H, Tian X, Zhao H, Liu X, Zhang S, Nie Y, Hu S, Miao X, Wang QD, Wang F, Chen T, Xu Q, Lui KO, Molkentin JD, Zhou B. Preexisting endothelial cells mediate cardiac neovascularization after injury. Journal of Clinical Investigation. 2017;127:2968–2981. doi: 10.1172/JCI93868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Heidt T, Courties G, Dutta P, Sager HB, Sebas M, Iwamoto Y, Sun Y, Da Silva N, Panizzi P, van der Laan AM, van der Lahn AM, Swirski FK, Weissleder R, Nahrendorf M. Differential contribution of monocytes to heart macrophages in steady-state and after myocardial infarction. Circulation Research. 2014;115:284–295. doi: 10.1161/CIRCRESAHA.115.303567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Hilgendorf I, Gerhardt LM, Tan TC, Winter C, Holderried TA, Chousterman BG, Iwamoto Y, Liao R, Zirlik A, Scherer-Crosbie M, Hedrick CC, Libby P, Nahrendorf M, Weissleder R, Swirski FK. Ly-6Chigh monocytes depend on Nr4a1 to balance both inflammatory and reparative phases in the infarcted myocardium. Circulation Research. 2014;114:1611–1622. doi: 10.1161/CIRCRESAHA.114.303204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Hu J, Dong A, Fernandez-Ruiz V, Shan J, Kawa M, Martínez-Ansó E, Prieto J, Qian C. Blockade of Wnt signaling inhibits angiogenesis and tumor growth in hepatocellular carcinoma. Cancer Research. 2009;69:6951–6959. doi: 10.1158/0008-5472.CAN-09-0541. [DOI] [PubMed] [Google Scholar]
  34. Hu P, Liu J, Zhao J, Wilkins BJ, Lupino K, Wu H, Pei L. Single-nucleus transcriptomic survey of cell diversity and functional maturation in postnatal mammalian hearts. Genes & Development. 2018;32:1344–1357. doi: 10.1101/gad.316802.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Hulsmans M, Clauss S, Xiao L, Aguirre AD, King KR, Hanley A, Hucker WJ, Wülfers EM, Seemann G, Courties G, Iwamoto Y, Sun Y, Savol AJ, Sager HB, Lavine KJ, Fishbein GA, Capen DE, Da Silva N, Miquerol L, Wakimoto H, Seidman CE, Seidman JG, Sadreyev RI, Naxerova K, Mitchell RN, Brown D, Libby P, Weissleder R, Swirski FK, Kohl P, Vinegoni C, Milan DJ, Ellinor PT, Nahrendorf M. Macrophages facilitate electrical conduction in the heart. Cell. 2017;169:510–522. doi: 10.1016/j.cell.2017.03.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Ivey MJ, Kuwabara JT, Pai JT, Moore RE, Sun Z, Tallquist MD. Resident fibroblast expansion during cardiac growth and remodeling. Journal of Molecular and Cellular Cardiology. 2018;114:161–174. doi: 10.1016/j.yjmcc.2017.11.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Ivey MJ, Tallquist MD. Defining the cardiac fibroblast. Circulation Journal. 2016;80:2269–2276. doi: 10.1253/circj.CJ-16-1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Jablonski KA, Amici SA, Webb LM, Ruiz-Rosado JdeD, Popovich PG, Partida-Sanchez S, Guerau-de-Arellano M. Novel markers to delineate murine M1 and M2 macrophages. PLOS ONE. 2015;10:e0145342. doi: 10.1371/journal.pone.0145342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Jeong D, Lee MA, Li Y, Yang DK, Kho C, Oh JG, Hong G, Lee A, Song MH, LaRocca TJ, Chen J, Liang L, Mitsuyama S, D'Escamard V, Kovacic JC, Kwak TH, Hajjar RJ, Park WJ. Matricellular Protein CCN5 Reverses Established Cardiac Fibrosis. Journal of the American College of Cardiology. 2016;67:1556–1568. doi: 10.1016/j.jacc.2016.01.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Kanisicak O, Khalil H, Ivey MJ, Karch J, Maliken BD, Correll RN, Brody MJ, J Lin SC, Aronow BJ, Tallquist MD, Molkentin JD. Genetic lineage tracing defines myofibroblast origin and function in the injured heart. Nature Communications. 2016;7:12260. doi: 10.1038/ncomms12260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Kfoury Y, Scadden DT. Mesenchymal cell contributions to the stem cell niche. Cell Stem Cell. 2015;16:239–253. doi: 10.1016/j.stem.2015.02.019. [DOI] [PubMed] [Google Scholar]
  42. King KR, Aguirre AD, Ye YX, Sun Y, Roh JD, Ng RP, Kohler RH, Arlauckas SP, Iwamoto Y, Savol A, Sadreyev RI, Kelly M, Fitzgibbons TP, Fitzgerald KA, Mitchison T, Libby P, Nahrendorf M, Weissleder R. IRF3 and type I interferons fuel a fatal response to myocardial infarction. Nature Medicine. 2017;23:1481–1487. doi: 10.1038/nm.4428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Ko YB, Kim BR, Yoon K, Choi EK, Seo SH, Lee Y, Lee MA, Yang JB, Park MS, Rho SB. WIF1 can effectively co-regulate pro-apoptotic activity through the combination with DKK1. Cellular Signalling. 2014;26:2562–2572. doi: 10.1016/j.cellsig.2014.07.026. [DOI] [PubMed] [Google Scholar]
  44. Kovacic JC, Dimmeler S, Harvey RP, Finkel T, Aikawa E, Krenning G, Baker AH. Endothelial to mesenchymal transition in Cardiovascular Disease: JACC State-of-the-Art Review. Journal of the American College of Cardiology. 2019;73:190–209. doi: 10.1016/j.jacc.2018.09.089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Lai SL, Marín-Juez R, Moura PL, Kuenne C, Lai JKH, Tsedeke AT, Guenther S, Looso M, Stainier DY. Reciprocal analyses in zebrafish and medaka reveal that harnessing the immune response promotes cardiac regeneration. eLife. 2017;6:e25605. doi: 10.7554/eLife.25605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nature Methods. 2012;9:357–359. doi: 10.1038/nmeth.1923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Lavine KJ, Epelman S, Uchida K, Weber KJ, Nichols CG, Schilling JD, Ornitz DM, Randolph GJ, Mann DL. Distinct macrophage lineages contribute to disparate patterns of cardiac recovery and remodeling in the neonatal and adult heart. PNAS. 2014;111:16029–16034. doi: 10.1073/pnas.1406508111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Lavine KJ, Pinto AR, Epelman S, Kopecky BJ, Clemente-Casares X, Godwin J, Rosenthal N, Kovacic JC. The macrophage in cardiac homeostasis and disease: jacc macrophage in CVD series (Part 4) Journal of the American College of Cardiology. 2018;72:2213–2230. doi: 10.1016/j.jacc.2018.08.2149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Leid J, Carrelha J, Boukarabila H, Epelman S, Jacobsen SE, Lavine KJ. Primitive embryonic macrophages are required for coronary development and maturation. Circulation Research. 2016;118:1498–1511. doi: 10.1161/CIRCRESAHA.115.308270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Liao Y, Smyth GK, Shi W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics. 2014;30:923–930. doi: 10.1093/bioinformatics/btt656. [DOI] [PubMed] [Google Scholar]
  51. Lu D, Dong W, Zhang X, Quan X, Bao D, Lu Y, Zhang L. WIF1 causes dysfunction of heart in transgenic mice. Transgenic Research. 2013;22:1179–1189. doi: 10.1007/s11248-013-9738-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Lu D, Wang J, Li J, Guan F, Zhang X, Dong W, Liu N, Gao S, Zhang L. Meox1 accelerates myocardial hypertrophic decompensation through Gata4. Cardiovascular Research. 2018;114:300–311. doi: 10.1093/cvr/cvx222. [DOI] [PubMed] [Google Scholar]
  53. Mahmoud AI, O'Meara CC, Gemberling M, Zhao L, Bryant DM, Zheng R, Gannon JB, Cai L, Choi WY, Egnaczyk GF, Burns CE, Burns CG, MacRae CA, Poss KD, Lee RT. Nerves Regulate Cardiomyocyte Proliferation and Heart Regeneration. Developmental Cell. 2015;34:387–399. doi: 10.1016/j.devcel.2015.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. McDavid A, Finak G, Chattopadyay PK, Dominguez M, Lamoreaux L, Ma SS, Roederer M, Gottardo R. Data exploration, quality control and testing in single-cell qPCR-based gene expression experiments. Bioinformatics. 2013;29:461–467. doi: 10.1093/bioinformatics/bts714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Mescher AL. Macrophages and fibroblasts during inflammation and tissue repair in models of organ regeneration. Regeneration. 2017;4:39–53. doi: 10.1002/reg2.77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Meyer IS, Jungmann A, Dieterich C, Zhang M, Lasitschka F, Werkmeister S, Haas J, Müller OJ, Boutros M, Nahrendorf M, Katus HA, Hardt SE, Leuschner F. The cardiac microenvironment uses non-canonical WNT signaling to activate monocytes after myocardial infarction. EMBO Molecular Medicine. 2017;9:1279–1293. doi: 10.15252/emmm.201707565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Mi H, Huang X, Muruganujan A, Tang H, Mills C, Kang D, Thomas PD. PANTHER version 11: expanded annotation data from gene ontology and reactome pathways, and data analysis tool enhancements. Nucleic Acids Research. 2017;45:D183–D189. doi: 10.1093/nar/gkw1138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Mirotsou M, Zhang Z, Deb A, Zhang L, Gnecchi M, Noiseux N, Mu H, Pachori A, Dzau V. Secreted frizzled related protein 2 (Sfrp2) is the key Akt-mesenchymal stem cell-released paracrine factor mediating myocardial survival and repair. PNAS. 2007;104:1643–1648. doi: 10.1073/pnas.0610024104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Mohamed TMA, Ang YS, Radzinsky E, Zhou P, Huang Y, Elfenbein A, Foley A, Magnitsky S, Srivastava D. Regulation of cell cycle to stimulate adult cardiomyocyte proliferation and cardiac regeneration. Cell. 2018;173:104–116. doi: 10.1016/j.cell.2018.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Molawi K, Wolf Y, Kandalla PK, Favret J, Hagemeyer N, Frenzel K, Pinto AR, Klapproth K, Henri S, Malissen B, Rodewald HR, Rosenthal NA, Bajenoff M, Prinz M, Jung S, Sieweke MH. Progressive replacement of embryo-derived cardiac macrophages with age. The Journal of Experimental Medicine. 2014;211:2151–2158. doi: 10.1084/jem.20140639. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Moore-Morris T, Guimarães-Camboa N, Banerjee I, Zambon AC, Kisseleva T, Velayoudon A, Stallcup WB, Gu Y, Dalton ND, Cedenilla M, Gomez-Amaro R, Zhou B, Brenner DA, Peterson KL, Chen J, Evans SM. Resident fibroblast lineages mediate pressure overload-induced cardiac fibrosis. Journal of Clinical Investigation. 2014;124:2921–2934. doi: 10.1172/JCI74783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Nacu E, Glausch M, Le HQ, Damanik FF, Schuez M, Knapp D, Khattak S, Richter T, Tanaka EM. Connective tissue cells, but not muscle cells, are involved in establishing the proximo-distal outcome of limb regeneration in the axolotl. Development. 2013;140:513–518. doi: 10.1242/dev.081752. [DOI] [PubMed] [Google Scholar]
  63. Nahrendorf M, Swirski FK, Aikawa E, Stangenberg L, Wurdinger T, Figueiredo JL, Libby P, Weissleder R, Pittet MJ. The healing myocardium sequentially mobilizes two monocyte subsets with divergent and complementary functions. The Journal of Experimental Medicine. 2007;204:3037–3047. doi: 10.1084/jem.20070885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Nahrendorf M, Swirski FK. Abandoning M1/M2 for a Network Model of Macrophage Function. Circulation Research. 2016;119:414–417. doi: 10.1161/CIRCRESAHA.116.309194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Noseda M, Harada M, McSweeney S, Leja T, Belian E, Stuckey DJ, Abreu Paiva MS, Habib J, Macaulay I, de Smith AJ, al-Beidh F, Sampson R, Lumbers RT, Rao P, Harding SE, Blakemore AI, Jacobsen SE, Barahona M, Schneider MD. PDGFRα demarcates the cardiogenic clonogenic Sca1+ stem/progenitor cell in adult murine myocardium. Nature Communications. 2015;6:6930. doi: 10.1038/ncomms7930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Palevski D, Levin-Kotler LP, Kain D, Naftali-Shani N, Landa N, Ben-Mordechai T, Konfino T, Holbova R, Molotski N, Rosin-Arbesfeld R, Lang RA, Leor J. Loss of macrophage wnt secretion improves remodeling and function after myocardial infarction in mice. Journal of the American Heart Association. 2017;6:e004387. doi: 10.1161/JAHA.116.004387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Patterson M, Barske L, Van Handel B, Rau CD, Gan P, Sharma A, Parikh S, Denholtz M, Huang Y, Yamaguchi Y, Shen H, Allayee H, Crump JG, Force TI, Lien CL, Makita T, Lusis AJ, Kumar SR, Sucov HM. Frequency of mononuclear diploid cardiomyocytes underlies natural variation in heart regeneration. Nature Genetics. 2017;49:1346–1353. doi: 10.1038/ng.3929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Pinto AR, Ilinykh A, Ivey MJ, Kuwabara JT, D'Antoni ML, Debuque R, Chandran A, Wang L, Arora K, Rosenthal NA, Tallquist MD. Revisiting cardiac cellular composition. Circulation Research. 2016;118:400–409. doi: 10.1161/CIRCRESAHA.115.307778. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Plein A, Fantin A, Denti L, Pollard JW, Ruhrberg C. Erythro-myeloid progenitors contribute endothelial cells to blood vessels. Nature. 2018;562:223–228. doi: 10.1038/s41586-018-0552-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Quaife-Ryan GA, Sim CB, Ziemann M, Kaspi A, Rafehi H, Ramialison M, El-Osta A, Hudson JE, Porrello ER. Multicellular transcriptional analysis of mammalian heart regeneration. Circulation. 2017;136:1123–1139. doi: 10.1161/CIRCULATIONAHA.117.028252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. R Development Core Team . Vienna, Austria: 2018. http://www.r-project.org/ [Google Scholar]
  72. Ramilowski JA, Goldberg T, Harshbarger J, Kloppmann E, Kloppman E, Lizio M, Satagopam VP, Itoh M, Kawaji H, Carninci P, Rost B, Forrest AR. A draft network of ligand-receptor-mediated multicellular signalling in human. Nature Communications. 2015;6:7866. doi: 10.1038/ncomms8866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Roselli C, Chaffin MD, Weng LC, Aeschbacher S, Ahlberg G, Albert CM, Almgren P, Alonso A, Anderson CD, Aragam KG, Arking DE, Barnard J, Bartz TM, Benjamin EJ, Bihlmeyer NA, Bis JC, Bloom HL, Boerwinkle E, Bottinger EB, Brody JA, Calkins H, Campbell A, Cappola TP, Carlquist J, Chasman DI, Chen LY, Chen YI, Choi EK, Choi SH, Christophersen IE, Chung MK, Cole JW, Conen D, Cook J, Crijns HJ, Cutler MJ, Damrauer SM, Daniels BR, Darbar D, Delgado G, Denny JC, Dichgans M, Dörr M, Dudink EA, Dudley SC, Esa N, Esko T, Eskola M, Fatkin D, Felix SB, Ford I, Franco OH, Geelhoed B, Grewal RP, Gudnason V, Guo X, Gupta N, Gustafsson S, Gutmann R, Hamsten A, Harris TB, Hayward C, Heckbert SR, Hernesniemi J, Hocking LJ, Hofman A, Horimoto A, Huang J, Huang PL, Huffman J, Ingelsson E, Ipek EG, Ito K, Jimenez-Conde J, Johnson R, Jukema JW, Kääb S, Kähönen M, Kamatani Y, Kane JP, Kastrati A, Kathiresan S, Katschnig-Winter P, Kavousi M, Kessler T, Kietselaer BL, Kirchhof P, Kleber ME, Knight S, Krieger JE, Kubo M, Launer LJ, Laurikka J, Lehtimäki T, Leineweber K, Lemaitre RN, Li M, Lim HE, Lin HJ, Lin H, Lind L, Lindgren CM, Lokki ML, London B, Loos RJF, Low SK, Lu Y, Lyytikäinen LP, Macfarlane PW, Magnusson PK, Mahajan A, Malik R, Mansur AJ, Marcus GM, Margolin L, Margulies KB, März W, McManus DD, Melander O, Mohanty S, Montgomery JA, Morley MP, Morris AP, Müller-Nurasyid M, Natale A, Nazarian S, Neumann B, Newton-Cheh C, Niemeijer MN, Nikus K, Nilsson P, Noordam R, Oellers H, Olesen MS, Orho-Melander M, Padmanabhan S, Pak HN, Paré G, Pedersen NL, Pera J, Pereira A, Porteous D, Psaty BM, Pulit SL, Pullinger CR, Rader DJ, Refsgaard L, Ribasés M, Ridker PM, Rienstra M, Risch L, Roden DM, Rosand J, Rosenberg MA, Rost N, Rotter JI, Saba S, Sandhu RK, Schnabel RB, Schramm K, Schunkert H, Schurman C, Scott SA, Seppälä I, Shaffer C, Shah S, Shalaby AA, Shim J, Shoemaker MB, Siland JE, Sinisalo J, Sinner MF, Slowik A, Smith AV, Smith BH, Smith JG, Smith JD, Smith NL, Soliman EZ, Sotoodehnia N, Stricker BH, Sun A, Sun H, Svendsen JH, Tanaka T, Tanriverdi K, Taylor KD, Teder-Laving M, Teumer A, Thériault S, Trompet S, Tucker NR, Tveit A, Uitterlinden AG, Van Der Harst P, Van Gelder IC, Van Wagoner DR, Verweij N, Vlachopoulou E, Völker U, Wang B, Weeke PE, Weijs B, Weiss R, Weiss S, Wells QS, Wiggins KL, Wong JA, Woo D, Worrall BB, Yang PS, Yao J, Yoneda ZT, Zeller T, Zeng L, Lubitz SA, Lunetta KL, Ellinor PT. Multi-ethnic genome-wide association study for atrial fibrillation. Nature Genetics. 2018;50:1225–1233. doi: 10.1038/s41588-018-0133-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Schafer S, Viswanathan S, Widjaja AA, Lim WW, Moreno-Moral A, DeLaughter DM, Ng B, Patone G, Chow K, Khin E, Tan J, Chothani SP, Ye L, Rackham OJL, Ko NSJ, Sahib NE, Pua CJ, Zhen NTG, Xie C, Wang M, Maatz H, Lim S, Saar K, Blachut S, Petretto E, Schmidt S, Putoczki T, Guimarães-Camboa N, Wakimoto H, van Heesch S, Sigmundsson K, Lim SL, Soon JL, Chao VTT, Chua YL, Tan TE, Evans SM, Loh YJ, Jamal MH, Ong KK, Chua KC, Ong BH, Chakaramakkil MJ, Seidman JG, Seidman CE, Hubner N, Sin KYK, Cook SA. IL-11 is a crucial determinant of cardiovascular fibrosis. Nature. 2017;552:110–115. doi: 10.1038/nature24676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, Amin N, Schwikowski B, Ideker T. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Research. 2003;13:2498–2504. doi: 10.1101/gr.1239303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Shinde AV, Frangogiannis NG. Fibroblasts in myocardial infarction: a role in inflammation and repair. Journal of Molecular and Cellular Cardiology. 2014;70:74–82. doi: 10.1016/j.yjmcc.2013.11.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Skelly DA, Squiers GT, McLellan MA, Bolisetty MT, Robson P, Rosenthal NA, Pinto AR. Single-Cell Transcriptional Profiling Reveals Cellular Diversity and Intercommunication in the Mouse Heart. Cell Reports. 2018;22:600–610. doi: 10.1016/j.celrep.2017.12.072. [DOI] [PubMed] [Google Scholar]
  78. Sklepkiewicz P, Shiomi T, Kaur R, Sun J, Kwon S, Mercer B, Bodine P, Schermuly RT, George I, Schulze PC, D'Armiento JM. Loss of secreted frizzled-related protein-1 leads to deterioration of cardiac function in mice and plays a role in human cardiomyopathy. Circulation: Heart Failure. 2015;8:362–372. doi: 10.1161/CIRCHEARTFAILURE.114.001274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Srivastava D, DeWitt N. DeWitt NIn: Vivo Cellular Reprogramming: The Next Generation. Cell. 2016;166:1386–1396. doi: 10.1016/j.cell.2016.08.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Stubbs D, DeProto J, Nie K, Englund C, Mahmud I, Hevner R, Molnár Z. Neurovascular congruence during cerebral cortical development. Cerebral Cortex. 2009;19:i32–i41. doi: 10.1093/cercor/bhp040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Surmann-Schmitt C, Sasaki T, Hattori T, Eitzinger N, Schett G, von der Mark K, Stock M. The Wnt antagonist Wif-1 interacts with CTGF and inhibits CTGF activity. Journal of Cellular Physiology. 2012;227:2207–2216. doi: 10.1002/jcp.22957. [DOI] [PubMed] [Google Scholar]
  82. Swirski FK, Nahrendorf M. Cardioimmunology: the immune system in cardiac homeostasis and disease. Nature Reviews Immunology. 2018;18:733–744. doi: 10.1038/s41577-018-0065-8. [DOI] [PubMed] [Google Scholar]
  83. Szklarczyk D, Morris JH, Cook H, Kuhn M, Wyder S, Simonovic M, Santos A, Doncheva NT, Roth A, Bork P, Jensen LJ, von Mering C. The STRING database in 2017: quality-controlled protein-protein association networks, made broadly accessible. Nucleic Acids Research. 2017;45:D362–D368. doi: 10.1093/nar/gkw937. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Tallquist MD, Molkentin JD. Redefining the identity of cardiac fibroblasts. Nature Reviews Cardiology. 2017;14:484–491. doi: 10.1038/nrcardio.2017.57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Tanay A, Regev A. Scaling single-cell genomics from phenomenology to mechanism. Nature. 2017;541:331–338. doi: 10.1038/nature21350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Thielens NM, Tedesco F, Bohlson SS, Gaboriaud C, Tenner AJ. C1q: A fresh look upon an old molecule. Molecular Immunology. 2017;89:73–83. doi: 10.1016/j.molimm.2017.05.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Tocharus J, Tsuchiya A, Kajikawa M, Ueta Y, Oka C, Kawaichi M. Developmentally regulated expression of mouse HtrA3 and its role as an inhibitor of TGF-beta signaling. Development, Growth and Differentiation. 2004;46:257–274. doi: 10.1111/j.1440-169X.2004.00743.x. [DOI] [PubMed] [Google Scholar]
  88. Travers JG, Kamal FA, Robbins J, Yutzey KE, Blaxall BC. Cardiac fibrosis: the fibroblast awakens. Circulation Research. 2016;118:1021–1040. doi: 10.1161/CIRCRESAHA.115.306565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Ubil E, Duan J, Pillai IC, Rosa-Garrido M, Wu Y, Bargiacchi F, Lu Y, Stanbouly S, Huang J, Rojas M, Vondriska TM, Stefani E, Deb A. Mesenchymal-endothelial transition contributes to cardiac neovascularization. Nature. 2014;514:585–590. doi: 10.1038/nature13839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Villani AC, Satija R, Reynolds G, Sarkizova S, Shekhar K, Fletcher J, Griesbeck M, Butler A, Zheng S, Lazo S, Jardine L, Dixon D, Stephenson E, Nilsson E, Grundberg I, McDonald D, Filby A, Li W, De Jager PL, Rozenblatt-Rosen O, Lane AA, Haniffa M, Regev A, Hacohen N. Single-cell RNA-seq reveals new types of human blood dendritic cells, monocytes, and progenitors. Science. 2017;356:eaah4573. doi: 10.1126/science.aah4573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Wang J, Liu S, Heallen T, Martin JF. The Hippo pathway in the heart: pivotal roles in development, disease, and regeneration. Nature Reviews Cardiology. 2018;15:672–684. doi: 10.1038/s41569-018-0063-3. [DOI] [PubMed] [Google Scholar]
  92. Wickham H. Ggplot2: Elegant Graphics for Data Analysis. New York: Springer-Verlag; 2009. [Google Scholar]
  93. Williams JW, Giannarelli C, Rahman A, Randolph GJ, Kovacic JC. Macrophage Biology, Classification, and Phenotype in Cardiovascular Disease: JACC Macrophage in CVD Series (Part 1) Journal of the American College of Cardiology. 2018;72:2166–2180. doi: 10.1016/j.jacc.2018.08.2148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Zhao Q, Eichten A, Parveen A, Adler C, Huang Y, Wang W, Ding Y, Adler A, Nevins T, Ni M, Wei Y, Thurston G. Single-Cell Transcriptome Analyses Reveal Endothelial Cell Heterogeneity in Tumors and Changes following Antiangiogenic Treatment. Cancer Research. 2018;78:2370–2382. doi: 10.1158/0008-5472.CAN-17-2728. [DOI] [PubMed] [Google Scholar]
  95. Zheng Z, Lee KS, Zhang X, Nguyen C, Hsu C, Wang JZ, Rackohn TM, Enjamuri DR, Murphy M, Ting K, Soo C. Fibromodulin-deficiency alters temporospatial expression patterns of transforming growth factor-β ligands and receptors during adult mouse skin wound healing. PLOS ONE. 2014;9:e90817. doi: 10.1371/journal.pone.0090817. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Zheng GX, Terry JM, Belgrader P, Ryvkin P, Bent ZW, Wilson R, Ziraldo SB, Wheeler TD, McDermott GP, Zhu J, Gregory MT, Shuga J, Montesclaros L, Underwood JG, Masquelier DA, Nishimura SY, Schnall-Levin M, Wyatt PW, Hindson CM, Bharadwaj R, Wong A, Ness KD, Beppu LW, Deeg HJ, McFarland C, Loeb KR, Valente WJ, Ericson NG, Stevens EA, Radich JP, Mikkelsen TS, Hindson BJ, Bielas JH. Massively parallel digital transcriptional profiling of single cells. Nature Communications. 2017a;8:14049. doi: 10.1038/ncomms14049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Zheng Z, James AW, Li C, Jiang W, Wang JZ, Chang GX, Lee KS, Chen F, Berthiaume EA, Chen Y, Pan HC, Chen EC, Li W, Zhao Z, Zhang X, Ting K, Soo C. Fibromodulin reduces scar formation in adult cutaneous wounds by eliciting a fetal-like phenotype. Signal Transduction and Targeted Therapy. 2017b;2:17050. doi: 10.1038/sigtrans.2017.50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Zhou B, Honor LB, He H, Ma Q, Oh JH, Butterfield C, Lin RZ, Melero-Martin JM, Dolmatova E, Duffy HS, Gise A, Zhou P, Hu YW, Wang G, Zhang B, Wang L, Hall JL, Moses MA, McGowan FX, Pu WT. Adult mouse epicardium modulates myocardial injury by secreting paracrine factors. Journal of Clinical Investigation. 2011;121:1894–1904. doi: 10.1172/JCI45529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Zhou X, Michal JJ, Zhang L, Ding B, Lunney JK, Liu B, Jiang Z. Interferon induced IFIT family genes in host antiviral defense. International Journal of Biological Sciences. 2013;9:200–208. doi: 10.7150/ijbs.5613. [DOI] [PMC free article] [PubMed] [Google Scholar]

Decision letter

Editor: Edward Morrisey1

In the interests of transparency, eLife includes the editorial decision letter and accompanying author responses. A lightly edited version of the letter sent to the authors after peer review is shown, indicating the most substantive concerns; minor comments are not usually included.

Thank you for submitting your article "Single-cell expression profiling reveals dynamic flux of cardiac stromal, vascular and immune cells in health and injury" for consideration by eLife. Your article has been reviewed by three peer reviewers, one of whom is a member of our Board of Reviewing Editors, and the evaluation has been overseen by Harry Dietz as the Senior Editor. The reviewers have opted to remain anonymous.

The reviewers have discussed the reviews with one another and the Reviewing Editor has drafted this decision to help you prepare a revised submission.

All of the reviewers as well as the editor believe that your study has merit. The reviewers felt that a better synopsis of the informatic analysis that would both simplify and better illustrate the findings would be beneficial to the report. There were also multiple instances where additional data including gene/protein expression, positional information of cell types identified, and confirmation of gene expression differences reported in the scRNA-seq analysis are required. Additional textual revisions are also requested and outlined below.

Reviewer #1:

The authors report on a new fibroblast type/state which is enriched for Wnt inhibition through increased expression of Wnt inhibitors, in particular Wif1. While the data is undoubtedly important and useful for the cardiac repair field, there are several notable issues that limit the impact of this report.

1) The data analysis as presented is confusing and does not lend itself to easy interpretation. If one of the main points to focus on is the Wntx cell type/state, then this should be much more clearly presented in the figures. This should include a clear assessment by GO category or other statistically significant methods to show that this cell cluster is enriched in expression of Wnt inhibitors.

2) In relation to point 1, if the Wntx cell type is enriched in expression of secreted Wnt inhibitors, then these cells could act on a neighboring cell type rather than in a cell autonomous manner. The authors should provide data on whether the Wntx cell is located in any specific relationship to other cardiac cells types. The current data in Figure 5 are at too low a resolution to define cell-cell relationships as Wif1 could be acting at short range.

3) A weakness of these studies is that there is no clear integration across the 3 different time frames examined i.e. 0, 3, and 7 days. To assess true relationships and how cells change after injury, a pseudotime analysis combining the 3 times points could help define what basal homeostatic fibroblast population gives rise to the Wntx and other cell types found at days 3 and 7.

Reviewer #2:

This is a well-written paper that concentrates on delineating the distinct immune and fibrotic signatures present in the acute and reparative phases post cardiac injury. The data presented here give insight into the different subtypes of immune and fibrotic cells in the heart, which have otherwise been hard to identify in bulk cellular and transcriptomic studies. Although this is a largely descriptive study, it provides a wealth of information that will be useful in the future to garner mechanistic insights into the signaling networks involved during cardiac injury and repair.

1) The authors mention the presence of a small percentage of hybrid endothelial cells which display both endothelial and fibroblast markers. To exclude the possibility that this population arises due to captured cell doublets, the authors should perform co-immunostaining for these markers to determine if this intermediate cell type exists in the heart. Immunostaining will also provide important spatial information regarding the location of this cell type in the heart.

2) Similarly, the authors mention that a significant proportion of the M2-macrophage population expressed endothelial markers, suggesting their ability to transdifferentiate into endothelial cells. This conclusion would also be more compelling if coupled with some spatial information. Are these cells prevalent in the areas near the site of injury?

3) In the differential proportion analysis in Figure 1E, the authors should indicate the changes in the total fibroblast population along with the subtypes already shown. There appears to be a significant decrease in the F-SH and F-SL populations at MI-day3 and this not very clearly indicated. In the latter half of the paper, the authors attribute this to the conversion of these fibroblast types to F-Act. However, the FACS data presented in the supplement do not entirely support this conclusion. This should be addressed.

4) Parts of the Discussion section simply reiterate the results. It might be more useful if the provided a more in-depth discussion regarding the implications of the different immune signatures observed post injury in the context of the known literature.

5) The authors need to cite and discuss in some detail the recently published, related findings of Epelman and coworkers "Self-renewing resident cardiac macrophages limit adverse remodeling following myocardial infarction" Nature Immunology 20, 29-39 (2019).

Reviewer #3:

The authors have performed an interesting single cell RNA-seq using the widely used 10X platform on adult ventricular murine cardiac tissue with or without myocardial infarction (MI). The authors focused their study on non-cardiomyocytes (TIP cells) with an emphasis on the fibroblast lineage. Moreover, they extensively describe a population of fibroblasts that express high levels of Wnt signaling-associated transcripts like Wif1.

Overall, while the work is very solid and interesting, it would be improved by further validation experiments and functional data to support the profiling data, which is predictive. The data analysis is of good quality, however, there are several technical issues as discussed below.

1) The authors include and present what appear to be obvious cell doublets in their final data set as EC-L1, EC-L2, and F-EC clusters (Figure 1). Doublets are expected on the 10X platform at a certain percentage which can be as high as ~5%, depending on the number of cells loaded. Without validation, these clusters that are suspected to be doublets should be removed and information regarding their identities and reasons for removal be added to the Materials and methods section. Moreover, the M2MΦ cluster which "expressed canonical endothelial markers" is partly composed of MΦ-endothelial cell doublets. This is obvious given that the endothelial-like "~29%" separates clearly from the main myeloid cluster (Figure 1D). The clustering parameter therefore is inadequate and unable to separate this issue and should be corrected. Finally, suggestions of a connection to trans-differentiation of myeloid cells into the endothelial cells and the contributions of embryonic EMPs to cardiac vessels should be removed from the manuscript to avoid confusion or more experimental data must be presented.

2) Based on the methods described in this study, it is unclear why epicardial and endocardial cells were not detected in the analysis shown in Figure 1. Even single nucleus RNA-seq studies, which are much less powerful, can detect these rare cell types in the heart (Hu et al., 2018). Please clarify in the writing.

3) The authors use the M1 and M2 classifications of macrophages and monocytes throughout their manuscript. Recent publications (that have been accurately referenced in the manuscript) have detailed the transcriptional responses of myeloid cells after MI (King et al., 2017) and after ischemia-reperfusion (IR) injury (Bajpai et al., 2018) using high throughput single cell RNA-sequencing. The authors should consider using nomenclature conventions more consistent with these previous reports.

4) Results from ligand-receptor analysis would greatly benefit from validation at the protein level.

5) It is unclear why the authors performed additional scRNA-seq on Pdgfra lineage cells, and what information was gained from these experiments. Please clarify in the writing. Similarly, what were the findings from the Fluidigm scRNA-seq experiments? Did they differ from or confirm the droplet-based scRNA-seq results? Please clarify in the writing.

6) It is unclear if the authors are claiming that the F-SH cells, which they additionally denote as CFU-Fs are cardiac mesenchymal stem cells (MSCs)? This section of the Results needs to be more clearly written.

7) In Figure 5, no strong evidence is provided that the indicated images are in the borderzone. Please clarify.

eLife. 2019 Mar 26;8:e43882. doi: 10.7554/eLife.43882.061

Author response


Reviewer #1:

The authors report on a new fibroblast type/state which is enriched for Wnt inhibition through increased expression of Wnt inhibitors, in particular Wif1. While the data is undoubtedly important and useful for the cardiac repair field, there are several notable issues that limit the impact of this report.

1) The data analysis as presented is confusing and does not lend itself to easy interpretation. If one of the main points to focus on is the Wntx cell type/state, then this should be much more clearly presented in the figures. This should include a clear assessment by GO category or other statistically significant methods to show that this cell cluster is enriched in expression of Wnt inhibitors.

Thank you for this suggestion – we have accordingly made changes to both the figure presentation and structure of the text to more clearly present our findings on F-WntX, and to aid in interpretation of the data. Regarding figures, we have built a new figure (Figure 5) and modified Figure 6 so that they focus exclusively on the F-WntX cells. Figure 5 focusses specifically on the presentation of results from the bioinformatics analysis of the F-WntX population. This includes representative GO terms over-represented in genes up-regulated in F-WntX compared to other populations, including the term negative regulation of Wnt signalling (Figure 5C). In relation to comment 2 of reviewer 1, we have incorporated additional figure panels in Figure 5 focussing on the paracrine functions of F-WntX using bioinformatic analysis of ligand-receptor connections (Figure 5D, E). As also requested, we have included high-resolution immunofluorescence images to examine the spatial relationship of F-WntX cells to other cell types (Figure 6F-H).

In the text, we have accordingly re-structured the section on analysis of Pdgfra-GFP+ fibroblasts to better streamline the presentation of results. All analyses comparing the F-SL and F-SH populations are now presented together. Similarly, all bioinformatics analysis on F-WntX is now presented within the same section, in relation to Figure 5, followed by the analysis of WIF1 IF in relation to Figure 6. These are presented in sections Novel Pdgfra-GFP+ fibroblast populations and Localization and composition of WIF1+ cells, respectively. To further aid data interpretation, we have built a schematic of our interpretation of the flux and differentiation/pseudotime dynamics of the GFP+ populations (Figure 9), which is referred to in the Discussion.

2) In relation to point 1, if the Wntx cell type is enriched in expression of secreted Wnt inhibitors, then these cells could act on a neighboring cell type rather than in a cell autonomous manner. The authors should provide data on whether the Wntx cell is located in any specific relationship to other cardiac cells types. The current data in Figure 5 are at too low a resolution to define cell-cell relationships as Wif1 could be acting at short range.

As indicated in our response to Comment 1, we have performed additional co-immunostainings and have taken higher resolution images to investigate whether F-WntX cells can be observed to co-localise with specific cell types. These higher-resolution images are contained in Figure 6F-H. Based on 2D distance, Figure 6H shows frequent close proximity or contact between WIF1+ cells and endothelial cells marked by CD31. This result suggests a potential interaction between these two cells types which is also suggested by the cell-cell ligand-receptor connections. Such interactions were less evident for WIF1+ cells and those expressing α-SMA or CD45 (Figure 6F, G), which is consistent with the lack of a predicted communication between F-WntX cells and MYO or immune cells in the cell communication analysis.

While we hope we have addressed reviewer 1’s concerns with respect to the resolution of immunofluorescence micrographs, we acknowledge that the spatial data presented in not quantitative or 3D. Cardiac tissue is complex, being composed of large cardiomyocytes embedded within an extracellular matrix containing populations of endothelial cells, perivascular cells, fibroblasts, and immune, neural and other cell types, all in close proximity. These cells will be communicating with others via membrane contact (possibly though long membrane protrusions or nanotubes) or gradients of secreted proteins. With MI, tissue structure is degraded and in certain regions surviving cells are overwhelmed by invading immune cells during MI, making it harder to make sense of which cardiac cells are in closest proximity. This complexity precludes a more detailed analysis of the spatial and functional relationship between WIF1+ cells and other cardiac interstitial cells for this paper. These questions could be addressed in part using quantitative 3D imaging and a combination of cell type specific markers, and membrane and signalling markers, but we feel that such a high level of experimentation is outside the scope of this work.

We have modified the text in the Results section under Localization and composition of WIF1+ cells, and in the Discussion, to reflect the above observations on 2D proximity.

3) A weakness of these studies is that there is no clear integration across the 3 different time frames examined i.e. 0, 3, and 7 days. To assess true relationships and how cells change after injury, a pseudotime analysis combining the 3 times points could help define what basal homeostatic fibroblast population gives rise to the Wntx and other cell types found at days 3 and 7.

Individual tSNE plots representing sham and MI-days 3 and 7 data for TIP and Pdgfra-GFP+ populations have been presented at the beginning of each section (Figures 1A, 4A, 8A, B). Furthermore, population proportion data (Figures 1A and 4B; Figure 1—figure supplement 1A, B; Figure 1—figure supplement 4C; Figure 4—figure supplement 1F), population gene expression data (Figure 2C; Figure 1—figure supplement 3D) and cell trajectory analyses (Figures 7B, 8C, D) have been broken down according to sham and MI-days 3 and 7, as appropriate for interpretation. However, in order to manage the large amounts of subsequent data analyses (mostly gene expression data displayed on tSNE plots), we have generally had to display outputs on aggregate plots of sham/MI data. This includes the Diffusion Map presented in Figure 7A, which is nonetheless useful in that is indicates the proposed trajectory of cells from unactivated populations F-SL and F-SH through to F-WntX, F-Act, proliferating cells, and myofibroblasts. These can be interpreted by back-reference to the original tSNE populations plots showing the breakdown according to sham and MI-days 3 and 7 data. Thus, showing all sham/MI data would be unmanageable. However, we understand the reviewer’s point, and so to make data more accessible we have added population labels to some of the aggregate plots. Furthermore, our interpretation of these dynamics is now summarised in the schematic in Figure 9.

Reviewer #2:

This is a well-written paper that concentrates on delineating the distinct immune and fibrotic signatures present in the acute and reparative phases post cardiac injury. The data presented here give insight into the different subtypes of immune and fibrotic cells in the heart, which have otherwise been hard to identify in bulk cellular and transcriptomic studies. Although this is a largely descriptive study, it provides a wealth of information that will be useful in the future to garner mechanistic insights into the signaling networks involved during cardiac injury and repair.

1) The authors mention the presence of a small percentage of hybrid endothelial cells which display both endothelial and fibroblast markers. To exclude the possibility that this population arises due to captured cell doublets, the authors should perform co-immunostaining for these markers to determine if this intermediate cell type exists in the heart. Immunostaining will also provide important spatial information regarding the location of this cell type in the heart.

2) Similarly, the authors mention that a significant proportion of the M2-macrophage population expressed endothelial markers, suggesting their ability to transdifferentiate into endothelial cells. This conclusion would also be more compelling if coupled with some spatial information. Are these cells prevalent in the areas near the site of injury?

We thank the reviewer for the comments regarding doublets, which is an issue that was also raised by reviewer 3 in their Comment 1. In line with reviewer 3’s request that “Without validation, these clusters that are suspected to be doublets should be removed and information regarding their identities and reasons for removal be added to the Materials and methods section”, we have restructured the figures and text. First, we removed all “hybrid” populations from the initial tSNE plots and from subsequent data analyses to produce dendrograms, DE gene lists and GO terms etc, with an explanation given in the text and Materials and methods. However, we recognise the potential biological importance of true hybrid populations as they could betray differentiation or trans-differentiation events that are otherwise difficult to detect. Thus, in response to reviewer 2’s suggestion, we have dedicated a section at the end of the Discussion of TIP cell populations (Hybrid populations in TIP) where we acknowledge the problems of doublets and then present new data in support of two of the hybrid populations identified by scRNAseq. We performed flow cytometry experiments with stringent filtering to exclude doublets and aggregates (see also response to reviewer 3, Comment 1), which shows a minor population of Pdgfra-GFP+/CD31+ cells in both sham and MI hearts, providing support for the presence of F-EC. We also performed flow and detected a population of MΦ cells (identified as CD45+F4/80+CD206+) expressing a recognised marker of ECs (CD31), in support of the M2MΦ-EC population. These data are presented in Figure 1—figure supplements4-7. We acknowledge that we cannot rule out that doublets may still contribute to the hybrid F-EC and M2MΦ-EC populations; however, we believe that our new data warrant a specific discussion of hybrids in this manuscript to highlight the importance of further investigation.

3) In the differential proportion analysis in Figure 1E, the authors should indicate the changes in the total fibroblast population along with the subtypes already shown. There appears to be a significant decrease in the F-SH and F-SL populations at MI-day3 and this not very clearly indicated. In the latter half of the paper, the authors attribute this to the conversion of these fibroblast types to F-Act. However, the FACS data presented in the supplement do not entirely support this conclusion. This should be addressed.

These are good suggestions and we have included an additional supplementary figure panel (Figure 1—figure supplement 1C) that contains the proportion changes of the high-level cell types, including fibroblasts, by merging appropriate clusters (F-SL, F-SH, F-WntX and F-Act into fibroblasts, the EC1, EC2 and EC3 clusters into endothelial cells etc.). We have modified the text in the sub-section on DPA (Differential proportion analysis) to specifically mention the significant decrease in F-SH and F-SL proportions. As we focus more specifically on F-SL and F-SH, including their flux after injury in the section on the Pdgfra-GFP+ enriched cells, we also direct readers to this section for a more detailed description of the populations, including their flux after MI. We acknowledge that our flow cytometry data does not prove that the F-SL and F-SH populations definitely differentiate into F-Act, and we have toned down the claim that F-SL/F-SH diminishment occurs as a result of F-Act expansion.

4) Parts of the Discussion section simply reiterate the results. It might be more useful if the provided a more in-depth discussion regarding the implications of the different immune signatures observed post injury in the context of the known literature.

Thank you for this appropriate suggestion. We have substantially shortened the Discussion to avoid too much reiteration of results while retaining pertinent discussion of implications. We have also inserted a longer treatment of monocyte/macrophage populations.

5) The authors need to cite and discuss in some detail the recently published, related findings of Epelman and coworkers "Self-renewing resident cardiac macrophages limit adverse remodeling following myocardial infarction" Nature Immunology 20, 29-39 (2019).

We thank the reviewer for pointing out this recent detailed paper and we have now cited it extensively in the Results and Discussion. Specifically, we have compared the markers for the cardiac resident macrophage/DC sub-populations that were identified by Epelman and co-workers with our data. There is limited scope for a direct comparison as Epelman and co-workers enrich for macrophage/DC populations in their paper using cell surface markers and therefore have more power to identify resident macrophage sub-populations than we do within total cardiac interstitial cells, where tissue-resident macrophages represent only ~3% of cells in sham hearts and ~1% of cells in MI hearts. Pleasingly, however, we were able to discern both the Timd4-Lyve1-MHC-IIhighCcr2- and Timd4+Lyve1+MHC-IIlowCcr2- self-renewing tissue-resident sub-populations that Epelman et al. describe as the major sub-sets of our tissue-resident macrophage population (MAC-TR). We have therefore cited the Epelman paper in the context of these observations, and have included the expression of the discriminating markers as found in our data – Timd4 and Lyve1 in Figure 2—figure supplement 1A, and MHC-II and Ccr2 in Figure 2. We have inserted this into the sub-section, Monocyte/macrophage cell identities and dynamics, as follows:

“Recent work using flow cytometry and scRNA-seq has delineated several subsets of cardiac tissue-resident MΦ, including a self-renewing and pro-regenerative population with the signature TIMD4+LYVE1+MHC-IIlowCCR2-, that self-renew and are not replaced by blood monocytes even after injury [Lavine et al., 2018]. […] The additional major subset of CCR2- tissue-resident MΦ [Lavine et al., 2018 could also be recognised at the scRNAseq level as the Timd4-Lyve1-H2-Aa(MHC-II)highCcr2- subset of MAC-TR – this population has been shown to have a low monocyte dependence during homeostasis but is almost fully replaced by monocytes after MI [Lavine et al., 2018]. “

Reviewer #3:

The authors have performed an interesting single cell RNA-seq using the widely used 10X platform on adult ventricular murine cardiac tissue with or without myocardial infarction (MI). The authors focused their study on non-cardiomyocytes (TIP cells) with an emphasis on the fibroblast lineage. Moreover, they extensively describe a population of fibroblasts that express high levels of Wnt signaling-associated transcripts like Wif1.

Overall, while the work is very solid and interesting, it would be improved by further validation experiments and functional data to support the profiling data, which is predictive. The data analysis is of good quality, however, there are several technical issues as discussed below.

1) The authors include and present what appear to be obvious cell doublets in their final data set as EC-L1, EC-L2, and F-EC clusters (Figure 1). Doublets are expected on the 10X platform at a certain percentage which can be as high as ~5%, depending on the number of cells loaded. Without validation, these clusters that are suspected to be doublets should be removed and information regarding their identities and reasons for removal be added to the Materials and methods section. Moreover, the M2MΦ cluster which "expressed canonical endothelial markers" is partly composed of MΦ-endothelial cell doublets. This is obvious given that the endothelial-like "~29%" separates clearly from the main myeloid cluster (Figure 1D). The clustering parameter therefore is inadequate and unable to separate this issue and should be corrected. Finally, suggestions of a connection to trans-differentiation of myeloid cells into the endothelial cells and the contributions of embryonic EMPs to cardiac vessels should be removed from the manuscript to avoid confusion or more experimental data must be presented.

See also Comments 1 and 2 of reviewer 2. We thank the reviewers for their insightful suggestions and we have performed a number of additional analyses to address these concerns. As reviewer 3 suggests, we have increased the clustering resolution so that the hybrid M2-macrophage-EC population (M2MΦ-EC) separates as a distinct cluster from the remaining M2-macrophage subset (Figure 1—figure supplement 4A-D). This adjustment did not substantially change the discrimination of other populations – the result was 29 clusters in total, 5 of which we identified as minor hybrid populations including the previously identified EC-L1, EC-L2, F-EC and M2MΦ-EC cells, and a new hybrid population expressing B-cell and T-cell markers (BC-TC). As the reviewer also suggests, we have removed hybrid clusters from the analysis of populations presented in the manuscript (Materials and methods).

Furthermore, as indicated in our response to reviewer 2 Comments 1 and 2, we have performed additional FACS experiments following a stringent process of doublet removal (see Figure 1—figure supplement 5) to test whether we could provide support for hybrid populations F-EC and M2MΦ-EC (described in detail in Hybrid populations in TIP; Figure 1—figure supplement 6, 7). These FACS experiments indeed revealed a GFP+CD31+ population in MI hearts (Figure 1—figure supplement 6) and a population of CD31+ cells co-expressing CD45 and macrophage markers (Figure 1—figure supplement 7). The additional data provide evidence for the existence of these hybrid populations. We have included an additional Results sub-section – Hybrid populations in TIP – where we describe the hybrid populations detected, raise the issue of doublets, and present the supporting FACS data for the 2 populations.

2) Based on the methods described in this study, it is unclear why epicardial and endocardial cells were not detected in the analysis shown in figure 1. Even single nucleus RNA-seq studies, which are much less powerful, can detect these rare cell types in the heart (Hu et al., 2018). Please clarify in the writing.

The Hu et al. study using single nucleus (sn)RNAseq identified a minor (1%) population of cells as epicardial, but did not identify endocardial cells. Skelly et al., 2018, performed scRNAseq using the 10X Chromium platform on interstitial cells of healthy adult murine hearts, but did not identify epicardial or endocardial cells. As commented on by Hu et al., it is possible that snRNA-seq, which overcomes issues of cell dissociation, is better suited than scRNA-seq for obtaining rare populations such as epicardial cells. We have clarified this point in the text in reference to the two above mentioned papers, within the section discussing our detection of epicardial-derived cells among the Pdgfra-GFP+ population, as follows:

“Consistent with a previous cardiac scRNA-seq analysis [Skelly et al., 2018], we did not detect epicardial cells in TIP data, suggesting that these cells are under-sampled in our experiments – this is likely technical as epicardial cells have been detected readily in single-nucleus RNA-seq [Hu et al., 2018].”

3) The authors use the M1 and M2 classifications of macrophages and monocytes throughout their manuscript. Recent publications (that have been accurately referenced in the manuscript) have detailed the transcriptional responses of myeloid cells after MI (King et al., 2017) and after ischemia-reperfusion (IR) injury (Bajpai et al., 2018) using high throughput single cell RNA-sequencing. The authors should consider using nomenclature conventions more consistent with these previous reports.

We thank the reviewer for their suggestion. We have endeavoured where possible to use nomenclature consistent with the previous papers. Regarding King et al. we have adopted their nomenclature for IFN-inducible cells to define the corresponding population of myeloid cells in our data that express interferon-responsive genes (MAC-IFNIC population). Regarding Bajpai et al., the focus of their analysis was on resident macrophage populations, and as we noted in our response above to reviewer 2 Comment 5, we do not have sufficient representation of tissue-resident macrophages to identify the many sub-types of myeloid cells that they report in their study. We believe that the nomenclature we are using is suitable for the data and analysis contained within our manuscript. At every stage, we provide the defining gene expression signature for each population identified.

4) Results from ligand-receptor analysis would greatly benefit from validation at the protein level.

We thank the reviewer for the suggestion. As indicated in our response to reviewer 1 Comment 2, we have performed additional co-staining experiments to provide support for the ligand-receptor analysis. In one experiment, we look at the relationship between WIF1+ cells and neighbouring cell types (see response to reviewer 1 Comment 2 for details and revised Figure 6). In an additional analysis, we looked at the relationship between GFP+ fibroblasts and CD31+ endothelial cells as the ligand-receptor map predicted a significant paracrine relationship between fibroblast populations and endothelial cells. Our co-stainings showed that GFP+ cells were indeed frequently observed in close spatial relationship to or in contact with CD31+ cells (shown in Figure 3—figure supplement 1A-B), with the additional finding the GFP and CD31 never overlap. We have added the following description to the sub-section Cell-cell communication analysis in TIP:

“Strikingly, fibroblast populations (F-SH, F-SL, F-Act and F-Wntx) appeared to communicate exclusively with vascular (ECs and mural) and glial cells. In line with this result, immunofluorescence analysis of sham and MI-day 3 hearts showed that Pdgfra-GFP+ fibroblasts were observed in close spatial relationships or direct contact with CD31+ endothelial cells (Figure 3—figure supplement 1A-B).”

5) It is unclear why the authors performed additional scRNA-seq on Pdgfra lineage cells, and what information was gained from these experiments. Please clarify in the writing. Similarly, what were the findings from the Fluidigm scRNA-seq experiments? Did they differ from or confirm the droplet-based scRNA-seq results? Please clarify in the writing.

The Pdgfra-GFP+ experiment using the 10X Chromium platform allowed us to enrich for fibroblasts at MI time-points that are otherwise dominated by the presence of immune cells, which dilute out the fibroblast sub-types. For example, at MI-day 3, over 80% of cells were myeloid in our data (Figure 1—figure supplement 2A), severely limiting observations we could made about fibroblasts. In fact, we gained significant new insights into fibroblast subsets using the enrichment approach. The discrimination of F-Act, F-Cyc and F-CI could not have been achieved using the TIP data-set alone. Likewise, by enriching for Pdgfra-lineage cells at MI-day 7, we were able to identify for the first time myofibroblast sub-types with contrasting fibrotic and anti-fibrotic signatures. We have clarified this point in the sub-section Single-cell RNA-seq of the Pdgfra-GFP+ cardiac fibroblast lineage as follows:

“To circumvent the dominance of immune cells in TIP following MI, which dilute out other cell populations, and to focus on fibroblast sub-populations (Figure 1A), we performed single-cell expression profiling on PDGFRα+CD31- cardiac interstitial cells at days 3 and 7 post-sham or MI.”

The Fluidigm experiment was performed on FACS-sorted SCA1+/PDGFRA+/CD31- (S+P+) cells that we had previously shown contained fibroblast colony-forming cells (cCFU-F). The main aims therefore were to connect the 10X Chromium data with this subset and to provide a deeper sequencing resource of the cell-types within this fraction. Specifically, it allowed us to test the hypothesis that the F-SH population defined from the 10X Chromium scRNA-seq data corresponded in large part to S+P+. The small number of cells profiled on Fluidigm (two independent experiments, ~50 cells each) meant that we could not perform a meaningful unbiased clustering analysis, as with the Chromium scRNA-seq. Instead, we used a machine learning classification approach (Materials and methods: Rank-based classifier for comparing data-sets) allowing us to demonstrate that the F-SH cells are predicted to correspond to the majority of cells from the FACS-defined S+P+ fraction. This leads to a very important conclusion: that the main fibroblast populations, F-SH and F-SL, have different functional properties with respect to colony formation. We have modified the section in Single-cell RNA-seq of the Pdgfra-GFP+ cardiac fibroblast lineage to clarify the purpose of the Fluidigm experiment, as follows:

“In order to confirm the relationship between F-SH and S+P+, we performed deeper scRNA-seq on 103 FACS-purified S+P+ cells from uninjured wild type mice using the Fluidigm platform and predicted cell identity using an iterative Random Forest (iRF) classifier [Basu et al., 2018] trained on populations defined in our GFP+ experiments in sham conditions using the Chromium platform (Materials and methods; Figure 4—figure supplement 2A).”

6) It is unclear if the authors are claiming that the F-SH cells, which they additionally denote as CFU-Fs are cardiac mesenchymal stem cells (MSCs)? This section of the Results needs to be more clearly written.

We apologise for the confusion. To clarify, we claim that cCFU-F are enriched in the F-SH population – we are not claiming that all of F-SH corresponds to cCFU-F or MSCs.

Out claim is based on several findings. First, previous observation based on FACS fractionation of cardiac fibroblasts and clonal colony formation assays allowed us to show that colony-forming cells (cCFU-F) were enriched within a Sca1+Pdgfra+CD31- (S+P+) fraction compared to the P+S- fraction (Chong et al., 2011; see also Figure 4 —figure supplement 1A, B). Second, from the analysis of our Fluidigm scRNA-seq of FACS-sorted S+P+ cells, we show in the current manuscript that the majority of S+P+ cells are predicted to correspond to the F-SH population we identified from the 10X Chromium scRNA-seq data. As F-SH carries the majority of S+P+ cells, the implication is that F-SH also carries the majority of the cCFU-F. We have modified the text in this section to make our claim clearer:

“Together, these results show that the F-SH population contains a subset of cells expressing Pdgfra, Ly6a(Sca1) and Thy1(Cd90) that is enriched in cCFU-F, highlighting the distinct expression signatures and functional properties of F-SH and F-SL.”

7) In figure 5, no strong evidence is provided that the indicated images are in the borderzone. Please clarify.

We have included a new panel (Figure 6A) showing the clear morphological distinction between healthy and infarcted area in the border zone. This image shows the presence of WIF1+ cells in the border zone of the damaged tissue and their absence in healthy (“remote”) myocardium.

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Data Citations

    1. Farbehi N, Patrick R, Dorison A, Xaymardan M, Wystub-Lis K, Janbandhu V, Ho JWK, Nordon RE, Harvey RP. 2018. Single-cell RNA-seq of mouse cardiac interstitial cells 3 and 7 days after sham or myocardial infarction injury. ArrayExpress database. E-MTAB-7376
    2. Farbehi N, Patrick R, Dorison A, Xaymardan M, Wystub-Lis K, Janbandhu V, Ho JWK, Nordon RE, Harvey RP. 2018. Single-cell RNA-seq of Pdgfra+/Sca1+/Cd31- mouse cardiac cells. ArrayExpress database. E-MTAB-7365
    3. Schafer S, Viswanathan S, Widjaja AA. 2017. Integrated target discovery screens identify IL11 as novel therapeutic target for fibrosis. Gene Expression Omnibus. GSE97117
    4. Skelly DA, Squiers GT, McLellan MA, Bolisetty MT, Robson P, Rosenthal NA, Pinto AR. 2017. Single cell RNA-Seq of the murine non-myocyte cardiac cellulome. ArrayExpress database. E-MTAB-6173
    5. Quaife-Ryan GA, Sim CB, Ziemann M, Kaspi A. 2017. Multicellular Transcriptional Analysis of Mammalian Heart Regeneration. Gene Expression Omnibus. GSE95755 [DOI] [PMC free article] [PubMed]
    6. Bochmann L, Sarathchandra P, Mori F, Lara-Pezzi E, Lazzaro D. 2010. Transcription profiling of mouse cardiac muscle and epicardium after left coronary artery ligation and sharm operation. ArrayExpress database. E-MEXP-2446

    Supplementary Materials

    Figure 1—source data 1. Source data for FACS quantifications summarized in Figure 1—figure supplement 6D,E and Figure 1—figure supplement 7B,C.
    DOI: 10.7554/eLife.43882.012
    Figure 4—source data 1. Source data for quantification of colony counts summarized in Figure 4—figure supplement 2E.
    DOI: 10.7554/eLife.43882.023
    Figure 6—source data 1. Source data for quantification of marker-positive cells summarized in Figure 6I.
    DOI: 10.7554/eLife.43882.029
    Source code 1. R code for processing and clustering of scRNA-seq data-sets, differential proportion analysis and cell communication network analysis.
    elife-43882-code1.zip (1.5MB, zip)
    DOI: 10.7554/eLife.43882.034
    Supplementary file 1. Differentially expressed genes across TIP sub-populations.
    elife-43882-supp1.xlsx (839.8KB, xlsx)
    DOI: 10.7554/eLife.43882.035
    Supplementary file 2. Differential proportion analysis p-value results for TIP and GFP+ sub-populations.
    elife-43882-supp2.xlsx (6.8KB, xlsx)
    DOI: 10.7554/eLife.43882.036
    Supplementary file 3. Differentially expressed genes between Mo/MΦ sub-populations in TIP.
    elife-43882-supp3.xlsx (138.9KB, xlsx)
    DOI: 10.7554/eLife.43882.037
    Supplementary file 4. Differentially expressed genes across GFP+ sub-populations.
    elife-43882-supp4.xlsx (217.5KB, xlsx)
    DOI: 10.7554/eLife.43882.038
    Supplementary file 5. Differentially expressed genes across GFP+ Diffusion Map trajectories.
    elife-43882-supp5.xlsx (118.9KB, xlsx)
    DOI: 10.7554/eLife.43882.039
    Supplementary file 6. GO Biological Process terms associated with GFP+ trajectory differentially expressed genes.
    elife-43882-supp6.xlsx (62.4KB, xlsx)
    DOI: 10.7554/eLife.43882.040
    Supplementary file 7. Differentially expressed genes from GFP+ day 3 injury response populations.
    elife-43882-supp7.xlsx (48.3KB, xlsx)
    DOI: 10.7554/eLife.43882.041
    Supplementary file 8. GO Biological Process terms associated with GFP+ day 3 injury response populations according to Diffusion Map trajectory: F-Act, F-CI and F-Cyc.
    elife-43882-supp8.xlsx (32.9KB, xlsx)
    DOI: 10.7554/eLife.43882.042
    Supplementary file 9. Differentially expressed genes between myofibroblast sub-populations in GFP+ day 7 scRNA-seq.
    elife-43882-supp9.xlsx (23.4KB, xlsx)
    DOI: 10.7554/eLife.43882.043
    Supplementary file 10. GO Biological Process terms associated with myofibroblast sub-populations in GFP+ day 7 scRNA-seq.
    elife-43882-supp10.xlsx (13.8KB, xlsx)
    DOI: 10.7554/eLife.43882.044
    Supplementary file 11. Spearman correlation test comparisons between TGF-β -treated cardiac fibroblast RNA-seq and GFP+ day 7 sub-populations.
    elife-43882-supp11.xlsx (13.8KB, xlsx)
    DOI: 10.7554/eLife.43882.045
    Transparent reporting form
    DOI: 10.7554/eLife.43882.046

    Data Availability Statement

    Sequencing data have been deposited in the ArrayExpress database at EMBL-EBI (www.ebi.ac.uk/arrayexpress) under accession codes E-MTAB-7376 and E-MTAB-7365.

    The following datasets were generated:

    Farbehi N, Patrick R, Dorison A, Xaymardan M, Wystub-Lis K, Janbandhu V, Ho JWK, Nordon RE, Harvey RP. 2018. Single-cell RNA-seq of mouse cardiac interstitial cells 3 and 7 days after sham or myocardial infarction injury. ArrayExpress database. E-MTAB-7376

    Farbehi N, Patrick R, Dorison A, Xaymardan M, Wystub-Lis K, Janbandhu V, Ho JWK, Nordon RE, Harvey RP. 2018. Single-cell RNA-seq of Pdgfra+/Sca1+/Cd31- mouse cardiac cells. ArrayExpress database. E-MTAB-7365

    The following previously published datasets were used:

    Schafer S, Viswanathan S, Widjaja AA. 2017. Integrated target discovery screens identify IL11 as novel therapeutic target for fibrosis. Gene Expression Omnibus. GSE97117

    Skelly DA, Squiers GT, McLellan MA, Bolisetty MT, Robson P, Rosenthal NA, Pinto AR. 2017. Single cell RNA-Seq of the murine non-myocyte cardiac cellulome. ArrayExpress database. E-MTAB-6173

    Quaife-Ryan GA, Sim CB, Ziemann M, Kaspi A. 2017. Multicellular Transcriptional Analysis of Mammalian Heart Regeneration. Gene Expression Omnibus. GSE95755

    Bochmann L, Sarathchandra P, Mori F, Lara-Pezzi E, Lazzaro D. 2010. Transcription profiling of mouse cardiac muscle and epicardium after left coronary artery ligation and sharm operation. ArrayExpress database. E-MEXP-2446


    Articles from eLife are provided here courtesy of eLife Sciences Publications, Ltd

    RESOURCES