Main

The blood system represents a paradigm for how multiple mature cell lineages in an adult tissue are replenished from self-renewing multipotent stem cells1. Almost 90% of >100 billion cells replaced daily in humans are blood cells2, predominantly short-lived platelets, granulocytes and erythrocytes but also lymphoid lineages. In a steady state, the need for replenishment can be fulfilled through the amplification of multiple stages of increasingly lineage-restricted and highly proliferative progenitors until they become fully restricted to one blood cell lineage1.

Common to all established models of the adult hematopoietic stem and progenitor cell (HSPC) hierarchy is the implication of only one mandatory differentiation pathway from hematopoietic stem cells (HSCs) to each short-lived mature blood cell lineage1,3, translating into considerable time for an HSC to replenish fully lineage-restricted progenitors. While not representing a problem in steady-state hematopoiesis (as each progenitor stage is also continuously replenished), it could pose a challenge if intermediate progenitors are acutely eliminated in physiological settings or in response to therapeutic insults to the bone marrow (BM), resulting in transient but critical reductions in short-lived platelets and granulocytes, which can lead to considerable morbidity, hospitalization and transfusion burden4. In addition to hemostasis and thrombosis5, platelets have important roles in immune responses6,7. Substantial efforts have been made toward enhancing platelet replenishment following therapeutic and physiological challenges, but success has been limited4.

Recently, platelet-biased and platelet-restricted HSCs (P-HSCs) were identified8,9,10, and fast-track pathways for the replenishment of megakaryocyte-restricted progenitors (MkPs) have been implicated11,12,13,14. However, as relying on phenotypic or molecular (rather than functional) definitions of HSCs, it remains unclear whether the proposed accelerated pathways are initiated from true HSCs or downstream progenitor cells.

Although single-cell transplantations and steady-state lineage tracing have established HSC heterogeneity, there is evidence only for hierarchical relationships between HSCs with different lineage biases8,15,16, implicating shared rather than separate pathways for blood lineage replenishment. Herein, we pursued multiple functional and molecular single-cell approaches, combined with genetic lineage tracing, to investigate whether HSCs with distinct lineage biases and restrictions have nonhierarchical relationships and/or use distinct platelet progenitor pathways.

Results

Nonhierarchically related distinct platelet-replenishing HSCs

We previously showed that a large fraction of phenotypic (lineage (LIN)SCA1+cKIT+ (LSK)CD34CD150+CD48) HSCs express von Willebrand factor (Vwf) and are transcriptionally platelet primed; upon transplantation, Vwf+ HSCs replenish all blood lineages but in a platelet-biased manner8. Through a kinetic analysis of blood lineage replenishment in >1,000 mice transplanted with a single adult BM LSKCD34CD150+CD48 HSC, we identified Vwf-tdTomato+ long-term HSCs (LT-HSCs) that do not contribute to B or T lymphocytes and are platelet–erythroid–myeloid restricted, replenishing blood in a platelet-biased manner (platelet contribution ≥3-fold higher than erythroid and myeloid contributions); a smaller fraction of these HSCs are platelet-restricted, replenishing exclusively platelets in primary recipients9. However, platelet-restricted HSCs can also replenish low levels of erythroid and myeloid (granulocyte and monocyte) cells when transplanted into secondary recipients9. Nevertheless, they remain highly platelet-biased and fail to contribute to B and T lymphoid lineages9, suggesting that platelet-bias is a stable and HSC-intrinsic property. In contrast, most single Vwf-tdTomato HSCs replenish all lympho-myeloid lineages upon transplantation (multilineage HSCs (multi-HSCs)), typically in a lineage-balanced or lymphoid-biased manner9.

In the present study, Vwf-tdTomato+ P-HSCs were defined as single HSCs that, upon transplantation, stably (at multiple analysis time points) contribute ≥50-fold more to platelets than to erythrocytes and myeloid cells, with little or no (≤0.01%) B and T lymphocyte contributions. Single Vwf-tdTomato+ and Vwf-tdTomato LSKCD34CD150+CD48 cells were purified from the adult BM of CD45.2 Vwf-tdTomato/Gata1-eGFP mice9 and transplanted into irradiated CD45.1 recipient mice (Extended Data Fig. 1a). Because LT-HSCs have been shown not to express Gata1-eGFP17 but express the endothelial protein C receptor (EPCR/CD201)18, we used Gata1-eGFP and, in some instances, also CD201+ gating to enhance HSC purity among LSKCD34CD150+CD48 cells. As HSCs can be identified only through their functional properties18, single P-HSCs and multi-HSCs were eventually identified through stable and distinct blood lineage output, as defined above.

We hypothesized that if alternative differentiation pathways for platelet replenishment exist downstream of HSCs, they should be used differently by Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs replenishing blood cell lineages in a balanced or lymphoid-biased pattern (Extended Data Fig. 1b,c).

Flow cytometry analysis demonstrated that single Vwf-tdTomato multi-HSCs robustly replenished all commonly defined phenotypic HSPC compartments19,20,21,22 (Fig. 1a). In contrast, single Vwf-tdTomato+ P-HSCs consistently reconstituted only LSK FMS-like tyrosine kinase 3 (FLT3)CD150+CD48 LT-HSCs and LINSCA1cKIT+ (LK)CD150+CD41+ MkPs, and, in a few instances, also LSKFLT3CD150+CD48+ multipotent progenitor 2 (MPP2) and LKCD150+CD41CD105CD16/32 premegakaryocyte–erythroid (preMegE) progenitors (Fig. 1a and Extended Data Fig. 1d). Three established subsets of LSK progenitors (LSKFLT3CD150CD48+ MPP3, LSKFLT3+ MPP4 and LSKFLT3CD150CD48 short-term HSCs (ST-HSCs))21,22 were consistently and robustly replenished by Vwf-tdTomato multi-HSCs but never by P-HSCs (Fig. 1a,b). Functional in vitro single-cell clonal analysis demonstrated that all MPP subsets replenished by Vwf-tdTomato multi-HSCs possessed megakaryocyte potential (Fig. 1c and Extended Data Fig. 1e), suggesting that Vwf-tdTomato multi-HSCs, at least in part, might use a different pathway with more progenitor intermediates for platelet replenishment compared to Vwf-tdTomato+ P-HSCs.

Fig. 1: Distinct platelet replenishment kinetics from single HSCs.
figure 1

a, HSPC hierarchy reconstituted by single LSKGata1-eGFPCD34CD150+CD48 Vwf-tdTomato multi-HSCs (n = 8) or Vwf-tdTomato+ P-HSCs (n = 9). The numbers shown are the mean ± s.e.m. percentage contributions to each population. Orange, reconstitution in all mice. Pink, reconstitution in some mice (the fraction of reconstituted mice is indicated in the upper left of each circle); mean of positive mice. Only progenitor populations present in ≥1/3 of mice and with ≥0.1% average reconstitution are shown. For the full P-HSC hierarchy, see Extended Data Fig. 1d. Phenotypic definitions: LT-HSC, LSKFLT3CD150+CD48CD45.2+; ST-HSC, LSKFLT3CD150CD48CD45.2+; MPP2, LSKFLT3CD150+CD48+CD45.2+; MPP3, LSKFLT3CD150CD48+CD45.2+; MPP4, LSKFLT3+CD45.2+; MkP, LKCD150+CD41+CD45.2+; preMegE progenitor, LKCD41CD16/32CD150+CD105CD45.2+; colony-forming unit-erythroid (CFU-E), LKCD41CD16/32CD150CD105+Gata1-eGFP+; pregranulocyte–monocyte (preGM) progenitor, LKCD41CD16/32CD150CD105CD45.2+; GMP, LKCD41CD16/32+CD150CD105CD45.2+; platelets (P), CD150+CD41+TER119Vwf-tdTomato+Gata1-eGFP+ for Vwf-tdTomatotg/+ Gata1-eGFPtg/+ donors; erythrocytes (E), TER119+CD150CD41Vwf-tdTomatoGata1-eGFP+; myeloid (granulocyte and monocyte) cells (M), CD11b+NK1.1CD19CD4/CD8aCD45.1CD45.2+; donor-derived B lymphocytes (B), CD19+NK1.1CD4/CD8aCD11bCD45.1CD45.2+; donor-derived T lymphocytes (T), CD4/CD8a+NK1.1CD11bCD19CD45.1CD45.2+. b, HSCs and MPPs replenished by Vwf-tdTomato multi-HSCs (n = 8) or Vwf-tdTomato+ P-HSCs (n = 9). Representative profiles and mean ± s.e.m. percentages of the parent LSKCD45.2+ gate are shown. c, Granulocyte/monocyte (GM) and megakaryocyte (MK) in vitro lineage potentials (mean ± s.e.m.) of HSCs and MPPs replenished by Vwf-tdTomato multi-HSCs (n = 3). Data are from 580–720 plated wells per population with a similar distribution across three replicates. Each dot represents an independent experiment.

Source data

We assessed whether a hierarchical relationship might exist between Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs (Fig. 2a). The LSKCD150+CD48 phenotypic HSC compartment replenished in the BM of recipients of a single transplanted Vwf-tdTomato multi-HSC remained exclusively or predominantly Vwf-tdTomato (Fig. 2b). In contrast, when replenished by a single Vwf-tdTomato+ P-HSC, the generated LSKCD150+CD48 cells remained predominantly Vwf-tdTomato+, although often containing a fraction of Vwf-tdTomato cells (Fig. 2b). As the HSC identity cannot be reliably defined by phenotype alone18, we next performed gold-standard secondary long-term reconstitution experiments with Vwf-tdTomato and Vwf-tdTomato+ cells purified from the BM of primary recipients reconstituted by a single Vwf-tdTomato multi-HSC or Vwf-tdTomato+ P-HSC (Fig. 2a,b). In 11 primary recipients, single Vwf-tdTomato multi-HSCs replenished Vwf-tdTomato LSKCD150+CD48 cells capable of balanced or lymphoid-biased multilineage contribution in secondary recipients but not Vwf-tdTomato+ LSKCD150+CD48 cells with secondary long-term contribution to platelets; therefore, no P-HSCs were produced (Fig. 2c,d and Supplementary Tables 1 and 2).

Fig. 2: Nonhierarchical relationship between distinct HSCs.
figure 2

a, Experimental outline of hierarchical HSC transplantations. Further details of cell phenotypes and numbers are provided in Supplementary Tables 1–3. b, Top, percentage of Vwf-tdTomato+ cells within LSKCD150+CD48CD45.2+ cells replenished in the BM of CD45.1 primary (1°) recipients transplanted with a single CD45.2 LSKGata1-eGFPCD34CD150+CD48 Vwf-tdTomato multi-HSC (n = 33) or Vwf-tdTomato+ P-HSC (n = 17). Dots represent individual mice, and lines represent mean ± s.e.m. Bottom, representative flow cytometry profiles. c, Left, representative histogram (mean ± s.e.m., n = 4) of Vwf-tdTomato expression in LSKCD150+CD48CD45.2+ cells replenished in primary CD45.1 recipients by a single CD45.2 Vwf-tdTomato multi-HSC. Middle, primary blood reconstitution at 16–37 weeks (wk) and secondary (2°) reconstitution at 16–18 weeks after transplantation of Vwf-tdTomato LSKCD150+CD48CD45.2+ cells sorted from primary recipients (mean ± s.e.m., n = 4 from four experiments; in secondary recipients, each dot represents the average reconstitution of one to two mice per primary recipient). Right, interpretation of results regarding (non)hierarchical replenishment of multi-HSCs and P-HSCs. Further details of cell phenotypes and numbers are provided in Supplementary Table 1. d, Left, representative histogram (mean ± s.e.m., n = 7) of Vwf-tdTomato expression in LSKCD150+CD48CD45.2+ cells replenished in primary recipients of a single transplanted LSKGata1-eGFPCD34CD150+CD48 Vwf-tdTomato multi-HSC. Middle, primary and secondary blood reconstitution 16–25 and 16–22 weeks after transplantation, respectively (mean ± s.e.m., n = 7 primary recipients from five experiments; in secondary recipients, each dot represents the average reconstitution of one to three mice per primary recipient). Right, interpretation of results regarding (non)hierarchical replenishment of multi-HSCs and P-HSCs. Further details of cell phenotypes and numbers are provided in Supplementary Table 2. e, Left, representative histogram (mean ± s.e.m., n = 3) of Vwf-tdTomato expression in LSKCD150+CD48CD45.2+ cells replenished in primary recipients of a single transplanted LSKGata1-eGFPCD34CD150+CD48 Vwf-tdTomato+ P-HSC. Middle, primary and secondary blood reconstitution 16–37 and 16–18 weeks after transplantation, respectively (mean ± s.e.m., n = 3 primary recipients from three experiments; in secondary recipients, each dot represents the average reconstitution of one to two mice per primary recipient). Right, interpretation of results regarding (non)hierarchical replenishment of multi-HSCs and P-HSCs. Further details of cell phenotypes and numbers are provided in Supplementary Table 3.

Source data

Although it is difficult to envision how P-HSCs could replenish multi-HSCs, given that replenishment remains stably platelet-biased and with little or no lymphoid contribution even upon secondary transplantation9, we also performed secondary transplantations with purified Vwf-tdTomato+ and Vwf-tdTomato LSKCD150+CD48 cells replenished in primary recipients by single Vwf-tdTomato+ P-HSCs. Single Vwf-tdTomato+ P-HSCs replenished Vwf-tdTomato+ LSKCD150+CD48 cells in primary recipients, which, upon secondary transplantation, replenished blood in a platelet-biased manner without detectable lymphoid output; in contrast, Vwf-tdTomato LSKCD150+CD48 cells did not provide secondary long-term myelo-lymphoid reconstitution and, therefore, contained no multi-HSCs (Fig. 2e and Supplementary Table 3). In one unique case, in which a single Vwf-tdTomato+ P-HSC also produced a sizable fraction of Vwf-tdTomato LSKCD150+CD48 cells, both the Vwf-tdTomato+ and Vwf-tdTomato cells were capable of long-term secondary blood replenishment but exclusively in a platelet-biased manner without any lymphoid output (Extended Data Fig. 1f and Supplementary Table 3), confirming that P-HSCs are unable to replenish multi-HSCs. Together with Vwf-tdTomato multi-HSCs being incapable of producing Vwf-tdTomato+ P-HSCs, these findings demonstrate that Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs are not hierarchically related and, therefore, should replenish platelets through different pathways.

Molecularly distinct megakaryocyte replenishment pathways

While the above experiments established that Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs are nonhierarchically related and might use distinct progenitor pathways for platelet replenishment, they also suggested that these cells, in part, might pass through shared progenitor stages (Fig. 1a). To compare the cellular trajectories in a more unbiased and in-depth manner, we performed single-cell whole-transcriptome (Smart-seq3) analysis23 of HSPCs long-term replenished by single Vwf-tdTomato+ P-HSCs (2,290 cells from seven reconstituted mice) and Vwf-tdTomato multi-HSCs (2,478 cells from eight reconstituted mice) (Fig. 3a and Supplementary Fig. 1a). Single cells replenished by Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs displayed similar quality control metrics (Supplementary Fig. 1b,c). After adjustment for batch effects using the mutual nearest-neighbor approach24 (Supplementary Fig. 1d–g), we performed dimensional reduction using uniform manifold approximation and projection (UMAP) based on the 2,000 most variable genes (Fig. 3b and Supplementary Fig. 2a–c). The multiple replicate recipients of a single Vwf-tdTomato+ P-HSC or Vwf-tdTomato multi-HSC showed consistently distinct contributions to different HSPC compartments (Fig. 3b and Extended Data Fig. 2), with minimal overlap within shared phenotypically or molecularly defined HSC and progenitor compartments (Fig. 3b and Extended Data Figs. 2, 3a,b and 4a,b), using different ranges of highly variable genes (HVGs) (Extended Data Fig. 4c) and following dimensional reduction by t-distributed stochastic neighbor embedding (tSNE) (Extended Data Fig. 4d). This suggests that Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs replenish molecularly distinct progenitor pathways for platelets.

Fig. 3: Distinct molecular platelet differentiation pathways.
figure 3

a, Left, experimental design (partly created with Biorender.com) for single-cell RNA sequencing of HSPCs generated by single Vwf-tdTomato+ P-HSCs (n = 7) or Vwf-tdTomato multi-HSCs (n = 8). Right, mean (dots indicate individual mice) contribution to blood lineages. b, UMAP of LINcKIT+ cells replenished by single Vwf-tdTomato+ P-HSCs (blue; n = 7 mice, 2,290 cells) or Vwf-tdTomato multi-HSCs (red; n = 8 mice, 2,478 cells). HSC, GMP, MkP and CFU-E cells were classified based on molecular signatures (Extended Data Fig. 3b). ce, AUC heatmaps for lineage signatures in single MkPs (c; n = 133 cells) and other HSPCs (d; n = 2,157 cells) replenished by single Vwf-tdTomato+ P-HSCs (seven mice) and in preMegE progenitors with an MkP and/or erythroid AUC score of >0.1 (e; n = 212 cells) replenished by single Vwf-tdTomato multi-HSCs (eight mice). Red rectangle, preMegE progenitors derived from Vwf-tdTomato multi-HSCs with combined MkP–erythroid signatures without myeloid and lymphoid signatures. f, UMAP after removing erythroid- and myeloid-restricted progenitors, visualized by donor type, molecular HSC (MolO > 0.22), molecular MkP (AUC > 0.25) and pseudotime order. g, DEGs (red; adjusted P < 0.05, absolute log2(fold change) > 0.5) when comparing MolO HSCs replenished by Vwf-tdTomato+ P-HSCs (n = 1,047 cells) and Vwf-tdTomato multi-HSCs (n = 97 cells). h, Gene-set enrichment normalized enrichment scores (NES; false discovery rate (FDR) q value < 0.1) of HALLMARK pathways based on DEGs detected in g. i, Expression (log2) of DEGs (adjusted P < 0.05, combined Wilcoxon/Fisher’s exact test) related to mTORC1 signaling when comparing MolO HSCs derived from Vwf-tdTomato+ P-HSCs (n = 1,047 cells, seven mice) and Vwf-tdTomato multi-HSCs (97 cells, eight mice). Boxes, first and third quartiles; line, median; whiskers, ±1.5× interquartile range; dots, outlier cells. The percentages of cells with detected gene expression (Methods) are shown. j, Fold-change (log2) tradeSeq fitted expression values of the top 70 DEGs (adjusted P < 0.01, log2(fold change) > 1, patternTest tradeSeq function) along pseudotime when comparing cells replenished by Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs. k, Pearson correlation (center line) along pseudotime comparing the expression of the top 70 DEGs and 70 randomly selected non-DEGs between cells replenished by Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs. Shading indicates the 95% confidence interval (CI). l, Normalized gene expression along pseudotime for cells shown in f. Lines show the mean expression count from the generalized additive model fit using tradeSeq. m, FLT3 expression in LSKCD45.2+ cells generated by single Vwf-tdTomato multi-HSCs (n = 8) and Vwf-tdTomato+ P-HSCs (n = 9). Representative profiles with mean ± s.e.m. percentages of parent gates and representative histograms are shown.

Source data

In agreement with the strong and stable platelet-biased replenishment, no bipotent cells with shared expression of both MkP and erythroid gene signatures were detected in MkPs or any other HSPCs replenished by Vwf-tdTomato+ P-HSCs (Fig. 3c,d). In contrast, and in agreement with previous studies19, apparent bipotent preMegE progenitors with combined megakaryocyte and erythroid but no myeloid or lymphoid gene expression were replenished by Vwf-tdTomato multi-HSCs (Fig. 3e).

After removing committed erythroid and myeloid progenitors replenished by Vwf-tdTomato multi-HSCs, we identified cells with high area under the curve (AUC) scores for HSC (‘molecular overlap population’ (MolO))25 or MkP signatures19 (Methods), as expected at the start and end of pseudotime, respectively (Fig. 3f and Supplementary Fig. 2d,e). The MolO score decreased along pseudotime for cells replenished by Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs, and this was also observed for the AUC score for HSCs with a low contribution to mature blood lineages and megakaryocyte bias26 (Extended Data Fig. 5a,b). In contrast, the AUC scores for signatures enriched in multilineage HSCs and HSCs with high lineage output26 followed the same pattern as the megakaryocyte signature (Extended Data Fig. 5a,b), increasing along pseudotime for both Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs, compatible with also capturing progenitors. Although no major differences were observed when comparing the enrichment of published HSC signatures21,25,26,27,28,29 in replenished HSCs, Vwf-tdTomato+ P-HSCs showed significantly higher AUC scores for functional HSC signatures (MolO and serial engrafter) and for restricted/biased lineage output (low output and megakaryocyte bias) (Extended Data Fig. 5c). MkP gene signatures19 were also enriched in HSCs originating from Vwf-tdTomato+ P-HSCs, whereas the granulocyte–monocyte progenitor (GMP) signature was higher in HSCs replenished from Vwf-tdTomato multi-HSCs (Extended Data Fig. 5d).

We identified 375 differentially expressed genes (DEGs; absolute log2(fold change) > 0.5, adjusted P < 0.05) within MolO-defined HSCs replenished by Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs (Fig. 3g and Supplementary Table 4). Mammalian target of rapamycin complex 1 (mTORC1) signaling was among the pathways most enriched in MolO HSCs replenished by Vwf-tdTomato+ P-HSCs (Fig.3g,h), including Fads1 (ref. 30), Fads2 (ref. 31), Ldha (ref. 32) and Nupr1 (ref. 33) (Fig. 3i and Supplementary Table 4), a pathway critical for HSC quiescence and self-renewal34. Other pathways associated with mTORC1 signaling35 were also enriched, including interferon response and MYC targets (Fig. 3h).

TradeSeq analysis36 established that 217 of 11,989 genes detected in at least 10% of all cells were differentially expressed between the two pathways along pseudotime (adjusted P < 0.01 and tradeSeq median absolute log2(fold change) > 1) (Fig. 3j and Supplementary Table 4). Pearson correlation analysis of the 70 top-ranked DEGs along pseudotime showed a more similar gene expression profile for cells located at trajectory start (HSCs) and end (MkPs) than those at intermediate stages, whereas 70 randomly selected genes demonstrated consistent and very high correlation (Fig. 3k and Supplementary Table 4). Taken together, DEGs along pseudotime define the separation of the two pathways from HSCs to MkPs.

Genes encoding markers previously assigned to distinct HSPC stages were among the top-ranked DEGs (Fig. 3g), including Vwf, Flt3 (encoding the receptor FLT3 expressed on MPP subsets, including those with little or no megakaryocyte potential)37 and Cd48 (encoding cell-surface CD48, suggested to define distinct MkP subsets)14, each showing distinct separation along pseudotime (Fig. 3l). Differential RNA expression of Flt3 and Cd48 correlated closely with the corresponding protein expression (Figs. 1b and 3m and Extended Data Fig. 3a).

In further agreement with the replenishment of distinct MkPs, 345 DEGs (absolute log2(fold change) > 0.5, false discovery rate (FDR)-adjusted P < 0.05) were identified when comparing molecularly defined MkPs replenished from Vwf-tdTomato+ P-HSCs (P-MkPs) and Vwf-tdTomato multi-HSCs (multi-MkPs) (Fig. 4a, Extended Data Fig. 6a and Supplementary Table 4) but, importantly, not in the gene expression signature defining MkPs (Extended Data Fig. 5d). One of the most differentially expressed MkP genes, Phactr1 (Fig. 4b), has been linked to increased platelet numbers38 and risk for early-onset cardiovascular thrombosis39. A significant enrichment of genes involved in fatty acid metabolism and cholesterol homeostasis was observed in P-MkPs (Fig. 4c,d and Supplementary Table 4), of relevance for the reported association between hypercholesterolemia and platelet homeostasis40, as well as for the disruption of cholesterol efflux activating MkPs41. P-MkPs were also enriched for coagulation genes, including Tbxa2r (encoding thromboxane 2, promoting platelet activation, and targeted by acetylsalicylic acid to prevent platelet overactivation42), and genes associated with complement (Fig. 4c,d and Supplementary Table 4), including prolycarboxypeptidase (Prcp; promoting enhanced coagulation through plasma prekallikrein43). P-MkPs also showed upregulated expression of interferon-α and -γ response genes (Fig. 4c,d and Supplementary Table 4), including interferon-induced transmembrane protein genes (Ifitm1 and Ifitm3) promoting immune-mediated platelet activation44, and inflammation-induced genes linked to the regulation of platelet activity (Lgals3bp (ref. 45) and Ifih1 (ref. 46)). As for HSCs (Fig. 3h), genes associated with hypoxia and MYC targets were enriched in P-MkPs (Fig. 4c). Multiple genes assigned to fatty acid metabolism (Aldh1a1, Car2), cholesterol homeostasis (Clu, Fads2) and coagulation (Vwf, Clu, Tbxa2r) were more highly expressed along pseudotime in the Vwf-tdTomato+ P-HSC trajectory before MkP generation (Fig. 4e), suggesting that some of the differences observed in MkPs replenished from the two pathways are already programmed at the HSC stage. Notably, the expression of the coagulation-related gene Clu was exclusive to the Vwf-tdTomato+ P-HSC pathway (Fig. 4e).

Fig. 4: Transcriptional characterization of MkPs replenished by single Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs.
figure 4

a, DEGs (red; adjusted P < 0.05 and absolute log2(fold change) > 0.5, combined Wilcoxon/Fisher’s exact test) when comparing molecular MkPs replenished by single Vwf-tdTomato multi-HSCs (multi-MkPs; n = 177) or Vwf-tdTomato+ P-HSCs (P-MkPs; n = 119). Cd48 and Vwf are highlighted in red. b, Expression (log2) of Phactr1, the top DEG (P < 0.05, log2(fold change) > 0.5, combined Wilcoxon/Fisher’s exact test) when comparing multi-MkPs and P-MkPs. The percentages of cells with detected expression (Methods) are indicated above the violin plots. c, Normalized gene-set enrichment score for HALLMARK pathways of DEGs enriched (FDR q value < 0.1) in multi-MkPs (red) and P-MkPs (blue). d, Expression (log2) of DEGs (adjusted P < 0.05, combined Wilcoxon/Fisher’s exact test) associated with fatty acid metabolism, cholesterol homeostasis, coagulation, interferon response and complement when comparing multi-MkPs (red; n = 177 cells, eight mice) and P-MkPs (blue; n = 119 cells, seven mice). Boxes, first and third quartiles; line, median; whiskers, the largest values within the ±1.5× interquartile range; dots, outliers. The percentages of cells with detected gene expression (Methods) are shown above the boxes. e, Expression (log2) of selected genes in all LINcKIT+ single cells generated by single Vwf-tdTomato multi-HSCs or Vwf-tdTomato+ P-HSCs along pseudotime. Dots represent individual cells, and lines represent LOESS (locally estimated scatterplot smoothing) curves of the expression for the HSC subtype (gray shading indicates the 95% CI).

To identify DEGs facilitating future identification and enrichment of MkPs distinct for the two differentiation pathways, genes encoding cell-surface antigens were further explored (Fig. 4a and Supplementary Table 4). A significant upregulation of Cd24a and Vwf (driving the expression of Vwf-tdTomato) was observed in molecularly defined P-MkPs, whereas Cd48 and Itga2 (encoding CD49b) were upregulated in multi-MkPs and negative in almost all P-MkPs (Fig. 5a). Flow cytometric index information confirmed the differential expression of the CD48 protein and Vwf-tdTomato in MkPs (Extended Data Fig. 6b), further validated in separate experiments together with CD24 and CD49b expression. In agreement with the transcriptional data, the expression of Vwf-tdTomato and cell-surface CD24 was distinctly upregulated in P-MkPs, whereas CD48 and CD49b were expressed in most multi-MkPs but virtually absent from P-MkPs (Fig. 5b,c).

Fig. 5: Phenotypic characterization of MkPs replenished by single transplanted Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs.
figure 5

a, Normalized expression (log2) of the indicated DEGs (adjusted P < 0.05, combined Wilcoxon/Fisher’s exact test) encoding Vwf and specified cell-surface proteins (Itga2 encodes CD49b) in single molecularly defined multi-MkPs (n = 177) or P-MkPs (n = 119). The percentage of cells expressing each gene (Methods) is indicated below each violin plot. b,c, Histograms (left), mean fluorescence intensity (MFI; middle) and percentage positive cells (right) based on flow cytometry analysis of the expression of Vwf-tdTomato reporter and CD24 (b; corresponding gene expression upregulated in P-MkPs) and of CD48 and CD49b (c; corresponding gene expression upregulated in multi-MkPs) in LKCD150+CD41+CD45.2+ MkPs replenished by single Vwf-tdTomato multi-HSCs (n = 3) or Vwf-tdTomato+ P-HSCs (n = 3). Histograms (percentage positive cells) show the expression on gated MkPs replenished by a platelet-restricted Vwf-tdTomato+ P-HSC and a Vwf-tdTomato multi-HSC. Dots represent individual mice, and lines represent mean ± s.e.m.

Source data

Our findings agree with previous studies suggesting that CD48 expression might define a distinct subset of MkPs12,13. We found P-MkPs to be uniformly CD48neg–lo at the transcriptional and protein levels, but a fraction of multi-MkPs were also negative for Cd48 and CD48 (Fig. 5 and Extended Data Fig. 6). While compatible with multi-HSCs partly replenishing CD48 MkPs overlap** with CD48 P-MkPs, this was not the case, as CD48+ and CD48 multi-MkPs showed highly overlap** DEGs when individually compared to CD48 P-MkPs, including for Cd24a, Itga2 and Vwf (Fig. 6a,b and Supplementary Table 4). The same pattern of DEGs was observed when the comparison was based on Cd48 mRNA expression (Fig. 6c and Supplementary Table 4), whereas very few DEGs were detected when comparing multi-MkPs negative or positive for Cd48 (Fig. 6d). Thus, while Cd48/CD48 expression specifically identifies multi-MkPs, Cd48/CD48 multi-MkPs are also molecularly distinct from P-MkPs.

Fig. 6: Replenishment of molecularly distinct MkPs by Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs.
figure 6

a, DEGs (adjusted P < 0.05, combined Wilcoxon/Fisher’s exact test, absolute log2(fold change) > 0.5; blue, genes overlap** with the DEGs comparing total P-MkPs and total multi-MkPs in Fig. 4a; red, genes not overlap** with the DEGs in Fig. 4a) when comparing CD48+ (left; n = 122 cells) or CD48 (right; n = 51 cells) molecularly defined MkPs replenished by single Vwf-tdTomato multi-HSCs (multi-MkPs) to CD48 molecularly defined MkPs replenished by single Vwf-tdTomato+ P-HSCs (P-MkPs; n = 101 cells). Cd48, Vwf, Cd24a and Itga2 are highlighted in blue. b, Distribution of Vwf log2(mRNA expression) and percentage of Vwf transcript-positive cells (left) and Vwf-tdTomato reporter fluorescence distribution (right) in CD48 and CD48+ molecularly defined multi-MkPs and CD48 P-MkPs. The numbers of analyzed single cells are indicated below the violin plots. c, DEGs (adjusted P < 0.05, combined Wilcoxon/Fisher’s exact test, absolute log2(fold change) > 0.5; blue, genes overlap** with the DEGs comparing total P-MkPs and total multi-MkPs in Fig. 4a; red, genes not overlap** with the DEGs in Fig. 4a) when comparing Cd48 transcript-positive (left) or Cd48 transcript-negative (right) multi-MkPs (n = 102 and 71 cells, respectively) to Cd48 transcript-negative P-MkPs (n = 114 cells). Cd48, Vwf, Cd24a and Itga2 are highlighted in blue. d, DEGs (adjusted P < 0.05, combined Wilcoxon/Fisher’s exact test, absolute log2(fold change) > 0.5; blue, genes overlap** with the DEGs comparing total P-MkPs and total multi-MkPs in Fig. 4a; red, genes not overlap** with the DEGs in Fig. 4a) when comparing Cd48 transcript-positive (n = 102 cells) to Cd48 transcript-negative (n = 71 cells) multi-MkPs. A detailed list of detected DEGs is provided in Supplementary Table 4.

Taken together, single-cell RNA-sequencing analyses of HSPCs replenished by single transplanted Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs unravel molecularly distinct progenitor differentiation trajectories for platelet replenishment, including transcriptionally and phenotypically distinct MkPs.

Usage of alternative platelet replenishment pathways

One of the most striking differences revealed by single-cell RNA sequencing was the virtual absence of Flt3 RNA expression in the entire pathway initiated by Vwf-tdTomato+ P-HSCs, contrasting with the high Flt3 expression in the Vwf-tdTomato multi-HSC pathway from the earliest stages of differentiation (Fig. 3l,m). Previous Flt3Cre fate-map** studies demonstrated that replenishment of all blood lineages, including platelets, occurs through Flt3-expressing stages in steady-state hematopoiesis47. Although we confirmed erythrocytes, granulocytes, and B and T lymphocytes to be almost 100% Flt3Cre-tdTomato+ in steady-state Flt3Cretg/+ R26Tom/+ mice, a fraction (10%) of platelets were consistently Flt3Cre-tdTomato (Fig. 7a and Extended Data Fig. 7a,b), compatible with steady-state platelets, unlike other blood lineages, being partly produced through an Flt3 pathway, initiated by P-HSCs. To test this possibility directly, we investigated to what degree single Vwf-eGFP multi-HSCs and platelet-restricted Vwf-eGFP+ P-HSCs from Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+ mice replenish Flt3Cre-tdTomato+ and Flt3Cre-tdTomato platelets after transplantation. In agreement with usage of an Flt3+ pathway, platelets and other blood cell lineages long-term replenished by single Vwf-eGFP multi-HSCs were almost entirely Flt3Cre-tdTomato+, whereas Vwf-eGFP+ platelet-restricted P-HSCs replenished almost exclusively Flt3Cre-tdTomato platelets (Fig. 7b and Extended Data Fig. 7c–e). This provided further support for a nonhierarchical relationship between Vwf-eGFP multi-HSCs and Vwf-eGFP+ P-HSCs and a strict separation between Vwf-eGFP multi-HSC and Vwf-eGFP+ P-HSC platelet replenishment pathways, as also supported by Vwf-eGFP multi-HSCs but not Vwf-eGFP+ P-HSCs replenishing FLT3+ Flt3Cre-tdTomato+ BM MPPs (Extended Data Fig. 7f).

Fig. 7: Fate map** of alternative platelet replenishment pathways.
figure 7

a, Flt3Cre-tdTomato labeling (mean ± s.e.m.) of blood lineages in steady-state Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+ mice (n = 7; 8–13 weeks old): platelets (CD150+CD41+TER119), erythrocytes (TER119+CD150CD41), myeloid (granulocyte) cells (CD11b+GR1+CD41NK1.1CD19CD4/CD8a), B lymphocytes (CD19+CD41NK1.1CD4/CD8aCD11bGR1) and T lymphocytes (CD4/CD8a+CD41NK1.1CD11bGR1CD19). Dots represent individual mice. Tom, tdTomato. b, Representative profiles of Flt3Cre-tdTomato labeling in platelets (left) and mean ± s.e.m. labeling of all lineages (right) replenished by a single CD45.2 Vwf-eGFP multi-HSC (top; n = 14) and Vwf-eGFP+ platelet-restricted P-HSC (bottom; n = 14) 18–21 weeks after transplantation.

Source data

Our present and previous findings9 are compatible with Vwf-tdTomato+ P-HSCs replenishing MkPs through fewer progenitor intermediates than Vwf-tdTomato multi-HSCs. In agreement with this, Gata1+ progenitors produced from transplanted Vwf+ P-HSCs replenish platelets with faster kinetics than Gata1+ progenitors from Vwf+ multi-HSCs48. To investigate whether this might also translate into faster steady-state kinetics of MkP replenishment through the P-HSC than the multi-HSC progenitor pathway, we explored published single-cell RNA-sequencing data, in which the kinetics of progenitor replenishment were assessed after recombination induction in Hoxb5CreERT2/+ R26Tom/+ reporter mice, specifically labeling the HSC compartment49. Interestingly, a subset of MkPs were the first lineage-restricted progenitors replenished by labeled HSCs49. Compared to MkPs replenished later, this early wave of MkPs showed an upregulation of genes also upregulated in P-MkPs and a downregulation of genes upregulated in multi-MkPs (Supplementary Fig. 3). Collectively, these findings raise the possibility that, upon insults to the hematopoietic system resulting in loss of MkPs, usage of the P-HSC pathway might more rapidly replenish platelets than multi-HSCs. Thus, we treated Flt3Cretg/+ R26Tom/+ mice with cyclophosphamide (CP), a cytotoxic agent that efficiently reduces megakaryocytes and MkPs4,50. We observed a small reduction in platelets and a more striking reduction in BM MkPs after CP treatment, accompanied by a clear decrease in the fraction of Flt3Cre-tdTomato+ platelets (from 95% before treatment to 82% at 7 days and 80% at 18 days after CP), followed by a return toward steady-state levels at 45 days; in contrast, Flt3Cre-tdTomato+ labeling of other lineages was unaffected at any time point (Fig. 8a and Extended Data Fig. 8a–c). This suggests that the Flt3 P-HSC pathway might have an important and lineage-specific role in accelerated platelet replenishment after CP treatment. LINcKIT+FLT3+ BM progenitors remained close to 100% Flt3Cre-tdTomato+ after CP treatment (Extended Data Fig. 8d), suggesting that the decreased labeling of blood platelets is not a consequence of decreased Flt3Cre recombination in FLT3+ BM progenitors. To exclude unspecific effects of CP on Cre recombination, we treated VavCretg/+ R26Tom/+ mice, in which Cre expression is under the control of the pan-hematopoietic Vav promoter51, resulting in all hematopoietic cells in the BM and blood being completely labeled. In these mice, platelets (and all other lineages) remained almost 100% VavCre-tdTomato+ following CP treatment (Extended Data Fig. 9). The nucleic acid-binding fluorescent dye thiazole orange (TO) labels enriched RNA content of newly generated reticulated platelets52. In a steady state, the majority (>90%) of TO+ platelets were Flt3Cre-tdTomato+ and, as previously reported53, only a minority of platelets were newly generated TO+ platelets. Following the CP challenge, the Flt3Cre-tdTomato fraction of TO+ platelets increased. Moreover, on day 4 after CP injection, TO+ reticulated platelets represented a significantly larger fraction of Flt3Cre-tdTomato platelets than Flt3Cre-tdTomato+ platelets (Fig. 8b,c and Extended Data Fig. 10a–c).

Fig. 8: Fate map** of alternative platelet replenishment pathways upon hematopoietic challenges.
figure 8

a, Flt3Cre-tdTomato labeling of blood lineages (as in Fig. 7a) in 7- to 11-week-old Flt3Cretg/+ R26Tom/+ mice after CP treatment on day 0. Analysis at baseline (day −3; n = 22) and on day 4 (n = 14), day 7 (n = 14), day 18 (n = 12) and day 45 (n = 7). Lines connect the mean of each time point. In platelets, ****P = 3.23 × 10−6 for day 4, ****P = 7.67 × 10−86 for day 7, ****P = 7.19 × 10−99 for day 18 and ****P = 6.88 × 10−45 for day 45 compared to baseline; ****P = 6.18 × 10−17 between days 18 and 45. Linear mixed-model two-sided analysis with P-value adjustment by the Benjamini–Hochberg procedure. b, Flt3Cre-tdTomato labeling (mean ± s.e.m.) of TO+ cells in 7- to 11-week-old Flt3Cretg/+ R26Tom/+ mice after CP treatment. Baseline (day −3), n = 16; day 4, n = 10; day 7, n = 10; day 18, n = 10; day 45, n = 7. Compared to baseline, ****P = 3.28 × 10−8 for day 4, ****P = 1.87 × 10−6 for day 7, ****P = 2.63 × 10−7 for day 18 and ***P = 1.23 × 10−4 for day 45. Two-way analysis of variance (ANOVA) with Bonferroni correction. Dots represent individual mice. c, TO labeling (mean ± s.e.m.) in the same mice as in b. Data represent the percentages of TO+ cells. ****P = 3.18 × 10−8 for day 4. Two-way ANOVA with Bonferroni correction. Dots represent individual mice. d, Flt3Cre-tdTomato labeling (as in c) upon 5FU treatment (day 0). Analysis at baseline (day −7 or −2) and on days 10, 17 and 24 after 5FU (n = 6). In platelets, ****P = 1.19 × 10−47 for day 10, ****P = 5.68 × 10−42 for day 17 and ****P = 3.95 × 10−23 for day 24 compared to baseline; ****P = 3.60 × 10−5 for day 17 and ****P = 4.10 × 10−27 for day 24, both compared to day 10. Compared to baseline, ****P = 4.05 × 10−18 for day 24 in erythrocytes and ****P = 6.30 × 10−11 for day 17 in myeloid cells. For myeloid cells, ****P = 9.60 × 10−7 for day 24 when compared to day 17. Linear mixed-model two-sided analysis with P-value adjustment by the Benjamini–Hochberg procedure. e, Platelet counts in Flt3Cretg/+ R26Tom/+ mice at baseline (day −10 or −3; n = 3) and on day 3 after anti-CD42b antibody treatment (n = 5). Dots represent individual mice, and lines represent mean ± s.e.m. f, Flt3Cre-tdTomato labeling (as in c) after anti-CD42b treatment (day 0). Analysis at baseline (day −10 or −3) and on days 3, 7 and 11 after anti-CD42b treatment (n = 5). A marginal significance was observed on day 11 (*P = 0.0466). Linear mixed-model two-sided analysis with P-value adjustment by the Benjamini–Hochberg procedure.

Source data

We next administered 5-fluorouracil (5FU), another myeloablative agent shown to reduce MkPs rapidly54. In line with this, we observed a transient decrease in platelets and a significant reduction in Flt3Cre-tdTomato+ platelets (from 95% to 58%) on day 10 after 5FU treatment (Fig. 8d), whereas FLT3+ progenitors in the BM were 100% Flt3Cre-tdTomato+ (Extended Data Fig. 10d,e). This was followed by a significant rebound to 78% tdTomato+ platelets on day 24 following 5FU treatment (Fig. 8d). Notably, we also observed a smaller yet significant decrease in Flt3Cre-tdTomato+ fractions of erythrocytes and myeloid cells (but not lymphocytes; Fig. 8d), probably reflecting that P-HSCs can also replenish lower levels of erythrocytes and myeloid cells.

Finally, we tested a challenge that specifically depletes platelets rather than progenitors, as, following such a challenge, rapid platelet replenishment would probably be primarily accomplished from existing progenitors (from both pathways) rather than HSCs; consequently, the contribution by the two pathways could be expected to be largely unaltered. We induced acute platelet depletion by administering an anti-CD42b antibody to Flt3Cretg/+ R26Tom/+ mice. As previously reported8, acute thrombocytopenia was observed 3 days after anti-CD42b treatment (Fig. 8e), with no impact on other lineages (Extended Data Fig. 10f) or the balance between Flt3Cre-tdTomato+ and Flt3Cre-tdTomato platelets (Fig. 8f). Unlike the loss of MkPs in response to CP treatment, an expansion of MkPs was observed 3 days after platelet depletion (Extended Data Fig. 10g), suggesting that a rapid expansion of MkPs might underlie the subsequent platelet recovery.

Together, these findings suggest that a rapid and transient increase in platelet replenishment can be achieved through the P-HSC pathway in response to challenges that reduce progenitors of the megakaryocyte lineage.

Discussion

While previous studies provided evidence only for hierarchical kinships between HSCs with different lineage biases8,15,16, we here establish a nonhierarchical relationship between Vwf-tdTomato HSCs stably replenishing all myeloid and lymphoid blood cell lineages without a platelet bias and Vwf-tdTomato+ P-HSCs replenishing only, or almost exclusively, platelets. Moreover, in contrast to established hierarchical models of hematopoiesis implicating only one mandatory differentiation pathway from HSCs for each lineage, we uncovered two distinct pathways for platelet replenishment. Rather than representing alternative differentiation pathways from the same HSC, these two pathways are initiated by distinct HSCs. Previously established signature scores for multi-HSCs and high-output HSCs26 increased along the pseudotime differentiation trajectory from HSCs toward MkPs, suggesting that these signatures also capture progenitor cells, in line with the reported deficient blood contribution upon secondary transplantation of HSCs marked by these signatures26. While the exact roles of Vwf-tdTomato multi-HSCs and Vwf-tdTomato+ P-HSCs in sustaining platelet homeostasis remain unclear, phenotypic and single-cell RNA-sequencing analyses demonstrated that Vwf-tdTomato+ P-HSCs bypass several stages of MPPs used by Vwf-tdTomato multi-HSCs. Moreover, single HSC transplantations showed that Vwf-eGFP+ P-HSCs, unlike Vwf-eGFP multi-HSCs, generate platelets without passing through Flt3-expressing progenitor stages, further corroborating the existence of alternative platelet replenishment pathways from HSCs with distinct lineage biases. Although, as previously shown47, all other blood lineages were almost exclusively produced through an Flt3+ pathway in the steady state, a significant fraction of platelets had not passed through Flt3-expressing progenitors. Upon suppression of megakaryopoiesis in Flt3Cretg/+ R26Tom/+ mice in response to CP treatment, we observed a rapid and transient increase in platelets not having passed through the Flt3+ pathway, whereas all cells of other blood lineages remained fully labeled, suggesting enhanced replenishment of platelets through the Flt3 pathway from P-HSCs. A similar enhanced usage of the Flt3 pathway was observed with 5FU, but with a smaller contribution also to myeloid and erythroid cells. This aligns with most P-HSCs, although being heavily platelet biased, also contributing to a lesser degree to granulocytes/monocytes and erythrocytes (more prominent upon challenge through secondary transplantation)9. In contrast, upon specific depletion of platelets, the relative contributions of the Flt3 and Flt3+ pathways were not significantly affected, compatible with preexisting MkPs, rather than HSCs, being responsible for the rapid platelet replenishment. Collectively, these findings suggest that the slower but more potent platelet replenishment by multi-HSCs through an Flt3+ pathway, including multiple stages of progenitor amplification before megakaryocyte commitment, is the default pathway in steady state; in contrast, a shorter and faster Flt3 progenitor pathway initiated by P-HSCs becomes more prominent shortly after challenges that reduce relevant MkPs in the BM.

Single-cell RNA-sequencing analysis demonstrated that even fully Mk-restricted progenitors (MkPs)19 are molecularly distinct in the two pathways. This agrees with the recent identification of MkP and megakaryocyte heterogeneity13,14,55, raising the possibility that the two pathways might also replenish platelets with distinct properties56.

Single-cell RNA sequencing provided insights into DEGs encoding cell-surface antigens that should facilitate the identification and further characterization of P-MkPs and multi-MkPs also in wild-type mice. Cd24a was highly upregulated in P-MkPs, also at the protein level, overlap** with differential expression of Vwf-tdTomato. In contrast, CD49b and CD48 expression was almost exclusive to multi-MkPs. CD48 has been reported to be expressed on an MkP subset14 and a distinct megakaryocyte subset with proposed immunoregulatory functions, including pathogen recognition and phagocytosis55.

To what degree the two pathways result in the replenishment of functionally distinct megakaryocytes or platelets remains to be investigated. It would also be important to exclude that neither of the two pathways produces dysregulated platelets. Being the first evidence of alternative differentiation pathways from distinct and nonhierarchically related HSCs, for any short-lived blood cell lineage, it raises the possibility that other short-lived myelo-erythroid blood cell lineages might also be replenished through more than one pathway. While no evidence exists for alternative platelet replenishment pathways in human hematopoiesis, this possibility is supported by findings compatible with the existence of human P-HSCs57,58. The identification of a fast-track platelet replenishment pathway initiated by a distinct class of HSCs could provide a platform for combatting transplantation- and drug-induced thrombocytopenia through means to stimulate this pathway or by expanding P-HSCs.

Methods

Animals

Animal experiments performed at the University of Oxford were approved by the Oxford Clinical Medicine Ethical Review Committee, and those performed at the Karolinska Institutet were approved by the regional review committee for animal ethics (Stockholms djurförsöksetiska nämnd). All experimental and mouse breeding procedures were performed in accordance with the UK Home Office and Swedish Jordbruksverket regulations.

Young adult (7–14 weeks old) Vwf-tdTomato/Gata1-eGFP mice9,17 (Vwf-tdTomatotg/+ Gata1-eGFPtg/+) and Flt3Cre/Rosa26tdTomato(Ai9)/Vwf-eGFP/Gata1-eGFP mice8,17,59 (Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+) on a C57BL/6OlaHsd (University of Oxford) or C57BL/6JrJ (Karolinska Institutet) background were used as BM donors in single HSC transplantations. Seven- to 18-week-old wild-type CD45.1 B6.SJL-Ptprca Pepcb/BoyJ (University of Oxford) and B6.SJL-Ptprca Pepcb/BoyCrl (Karolinska Institutet) mice were used as recipients in primary and secondary transplantations, as donors of unfractionated BM competitor cells, and for BM analysis after anti-CD42b antibody treatment. Recipient mice that did not survive or had to be killed before 16–18 weeks after the primary or secondary transplantation were excluded from analyses. Flt3Cre/Rosa26tdTomato(Ai14) mice (Flt3Cretg/+ R26Tom/+; 7–11 weeks old) and VavCre/Rosa26tdTomato(Ai14) mice51 (VavCretg/+ R26Tom/+; 8–23 weeks old) on a C57BL/6JrJ background were used for fate-map** and CP and 5FU treatment experiments. Flt3Cre/Rosa26tdTomato(Ai9)/Vwf-eGFP/Gata1-eGFP (Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+; 8–13 weeks old) mice on a C57BL/6OlaHsd background were also used for steady-state fate-map** and anti-CD42b treatment experiments. Mice were housed in individually ventilated cages at the Oxford JR facility (12/12 h light/dark cycle, 19–24 °C and 45–65% humidity) and the Karolinska Institutet KM facility (12/12 h light/dark cycle, 22 ± 1 °C and 50% humidity).

Single-cell transplantations

Single-cell sorting of adult BM HSCs was performed using a FACSAria II or FACSAria Fusion cell sorter (BD Biosciences), prepared by crushing pelvic and leg bones (and optionally also sternum and spine bones) into PBS with 5% fetal calf serum (FCS; Sigma-Aldrich) and 2 mM EDTA (Sigma-Aldrich). Single phenotypically defined HSCs (Extended Data Fig. 1a) were sorted from Vwf-tdTomatotg/+ Gata1-eGFPtg/+ mice (Vwf-tdTomato+ and Vwf-tdTomato fractions of LSKGata1-eGFPCD34CD150+CD48 or LSKGata1-eGFPCD34CD150+CD48CD201+ cells) and Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+ mice (Vwf/Gata1-eGFP+ and Vwf/Gata1-eGFP fractions of LSKCD34CD150+CD48 or LSKCD34CD150+CD48CD201+ cells).

Comparison of single-cell expression and coexpression of Vwf-eGFP and Gata1-eGFP in BM LSKCD34CD150+CD48 cells showed that <10% express Gata1-eGFP (Extended Data Fig. 7c). Moreover, Gata1-eGFP and CD201 are mutually exclusive in LSKCD34CD150+CD48 cells, indicating that eGFP expression in LSKCD34CD150+CD48CD201+ cells mainly reflects Vwf-eGFP+ HSCs (Extended Data Fig. 7c). Regardless of their cell-surface phenotype, P-HSCs and multi-HSCs were defined based on their long-term lineage replenishment pattern as established by blood lineage analysis at multiple time points (see the next sections).

In experiments with mice that coexpress Vwf-eGFP and Gata1-eGFP, eGFP was used for sorting of single eGFP+ and eGFP HSCs for transplantation and to identify donor-derived blood platelets (which express both Vwf and Gata1) and erythrocytes (which express Gata1) in the transplantation recipients.

Single HSCs were sorted by an automated cell deposition unit, refrigerated at 4 °C, into 96-well round-bottom plates (Corning) with 100 μl per well of Iscove’s modified Dulbecco’s medium (IMDM, Gibco) with 20% BIT-9500 serum substitute (Stem Cell Technologies), 100 U ml−1 penicillin and 0.1 mg ml−1 streptomycin (100× Pen/Strep, Hyclone), 2 mM l-glutamine (Gibco) and 0.1 mM 2-mercaptoethanol (Sigma-Aldrich). Single index-sorted HSCs were mixed with 2–3 × 105 wild-type CD45.1 unfractionated BM competitor cells (100 μl per well) and transplanted by intravenous lateral tail-vein injection into lethally irradiated CD45.1 mice (10–10.5 Gy, cesium-137 or X-ray). BM cell counts were measured manually with a hemacytometer and/or an automated cell counter (Sysmex XP-300 or ABX Pentra ES 60).

Blood reconstitution analysis

Peripheral blood was collected from a lateral tail vein into lithium–heparin or K3 EDTA microvettes (Sarstedt). The platelet supernatant was collected after centrifugation of blood samples at 100g for 10 min at room temperature. Then, it was mixed with a small fraction (0.5–1 μl) of red precipitate for combined analysis of platelets and erythrocytes. The remaining precipitate was incubated 1:1 with dextran (Sigma-Aldrich, Mr 450,000–650,000) 2% w/v in PBS for 20–30 min at 37 °C. Erythrocytes were lysed by incubation in ammonium chloride solution (Stem Cell Technologies) for 2 min at room temperature. Leukocyte samples were incubated with purified CD16/32 (Fc-block) for 10–15 min at 4 °C. Then, they were stained with anti-mouse antibodies for 15–20 min at 4 °C in PBS with 1–5% FCS and 2 mM EDTA. Samples were analyzed using LSRII and Fortessa cytometers (BD Biosciences). See Supplementary Table 5 for antibody details.

Donor-derived platelets were defined as follows: CD150+CD41+TER119Vwf-tdTomato+Gata1-eGFP+ for Vwf-tdTomatotg/+ Gata1-eGFPtg/+ donors and CD150+CD41+TER119Vwf/Gata1-eGFP+ for Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+ donors. Donor-derived erythrocytes: TER119+CD150CD41Vwf-tdTomatoGata1-eGFP+ for Vwf-tdTomatotg/+ Gata1-eGFPtg/+ donors and TER119+CD150CD41Vwf/Gata1-eGFP+ for Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+ donors. Donor-derived myeloid (granulocyte/monocyte) cells: CD11b+NK1.1CD19CD4/CD8aCD45.1CD45.2+. Donor-derived B cells: CD19+NK1.1CD4/CD8aCD11bCD45.1CD45.2+. Donor-derived T cells: CD4/CD8a+NK1.1CD11bCD19CD45.1CD45.2+. The granulocyte/monocyte identity of CD11b+NK1.1CD19CD4/CD8aCD45.1CD45.2+ cells from reconstituted recipient mice was confirmed by cytospins stained with eosin Y/azure A/methylene blue (Richard-Allan Scientific Three-Step Stain Set, Thermo Fisher Scientific) (Supplementary Fig. 4).

Categorization of reconstitution patterns

All five mature blood cell lineages (platelets, erythrocytes, myeloid cells, B cells and T cells) were considered. We considered primary and secondary recipients to be reconstituted by HSCs if the donor contribution to platelets was ≥0.1% at ≥16–18 weeks after transplantation9. Mice reconstituted by single Vwf+ P-HSCs and Vwf multi-HSCs were defined at ≥16–18 weeks after transplantation as follows. Vwf+ P-HSC: donor platelets ≥0.1%; donor platelet percentage ≥50-fold higher than donor erythrocytes and myeloid cells; and donor B and T cells undetectable (≤0.01%). Where indicated (Fig. 7b and Extended Data Fig. 7e,f), the stricter group of platelet-restricted HSCs was considered, in which donor erythrocyte, myeloid cell, B cell and T cell lineages were all below the detection level (≤0.01%). Vwf multi-HSC: donor platelets ≥0.1%; donor erythrocytes, myeloid cells, B cells and T cells all >0.01%; and donor platelets, erythrocytes and myeloid cells all ≤2-fold higher than B and T cells. Mice reconstituted with multilineage patterns with ≥2-fold platelet, platelet–erythroid and platelet–erythroid–myeloid bias were excluded from the Vwf multi-HSC group, as such biases are typical of Vwf+ multi-HSCs9.

Reconstitution analysis of HSPCs

BM HSPC reconstitution analysis was performed using FACSAria Fusion, LSRII and LSR Fortessa flow cytometers (BD Biosciences) after crushing pelvic and leg bones (and optionally also sternum and spine bones) into PBS with 5% FCS and 2 mM EDTA, followed by cKIT enrichment according to the manufacturer’s instructions (CD117 MicroBeads and magnetic activated cell sorting (MACS) LS columns, Miltenyi Biotec). cKIT-enriched BM cells were incubated with purified CD16/32 (Fc-block) for 15–20 min at 4 °C, followed by anti-mouse antibody staining for 15–20 min at 4 °C. For the myeloid progenitor panel, cells were incubated with fluorophore-conjugated CD16/32 before further staining. See Supplementary Table 5 for antibody details.

Phenotypic BM populations were defined as follows: LSK, LINSCA1+cKIT+; LK, LINSCA1cKIT+; LT-HSC, LSKFLT3CD150+CD48; ST-HSC, LSKFLT3CD150CD48; MPP2, LSKFLT3CD150+CD48+; MPP3, LSKFLT3CD150CD48+; MPP4, LSKFLT3+; MkP, LKCD150+CD41+; preMegE progenitor, LKCD41CD16/32CD150+CD105; CFU-E, LKCD41CD16/32CD150CD105+; preGM progenitor, LKCD41CD16/32CD150CD105; GMP, LKCD41CD16/32+CD150CD105. Cells within each population were considered donor-derived cells when CD45.2+CD45.1 or when Gata1-eGFP+ in the case of CFU-E cells due to their low CD45 expression.

In vitro lineage potentials

Donor-derived (CD45.2+CD45.1) phenotypic HSC and MPP populations, as defined above, were bulk sorted into Eppendorf tubes with X-VIVO 15 medium containing gentamycin and l-glutamine (Lonza) and supplemented with 10% FCS, 0.1 mM 2-mercaptoethanol, 25 ng ml−1 mouse stem cell factor (PeproTech), 25 ng ml−1 human thrombopoietin (PeproTech), 10 ng ml−1 human FLT3 ligand (Immunex) and 5 ng ml−1 mouse interleukin-3 (PeproTech). The volume of sorted cells was diluted as needed and manually distributed at an average of one cell per well into Terasaki microplates (Thermo Fisher Scientific) at 20 μl per well. The growth of granulocytes/macrophages and megakaryocytes was scored under an inverted microscope after 8 days of culture at 37 °C and 5% CO2 in a humidified incubator. Granulocyte/macrophage and megakaryocyte scores were also confirmed in a representative subset of wells through the analysis of cytospins stained with eosin Y/azure A/methylene blue (Richard-Allan Scientific Three-Step Stain Set, Thermo Fisher Scientific).

Secondary hierarchical transplantations

Secondary transplantations were performed 16–38 weeks after the primary single-HSC transplantation, and a final blood analysis of primary recipients was performed a maximum of 2 weeks before sorting to confirm the reconstitution pattern. BM cell suspensions from primary recipients of a single P-HSC or multi-HSC were prepared by crushing pelvic and leg bones (and optionally also sternum and spine bones) into PBS with 5% FCS and 2 mM EDTA, followed by cKIT enrichment according to the manufacturer’s instructions (CD117 MicroBeads and MACS LS columns, Miltenyi Biotec). Phenotypically defined donor-derived (CD45.2+CD45.1) HSCs were sorted into supplemented IMDM as described above, mixed with 2–3 × 105 wild-type CD45.1 unfractionated BM competitor cells and transplanted by intravenous lateral tail-vein injection into lethally irradiated CD45.1 secondary recipient mice (10–10.5 Gy, cesium-137 or X-ray).

In some experiments (Supplementary Table 1), all sortable donor-derived (CD45.2+CD45.1) LSK Vwf-tdTomato+ cells were transplanted while the Vwf-tdTomato cells were transplanted, kee** the same Vwf-tdTomato+/Vwf-tdTomato ratio observed within the donor-derived phenotypic HSC compartment (LSKCD150+CD48CD45.2+). In other experiments (Supplementary Tables 2 and 3), all sortable donor-derived (CD45.2+CD45.1) HSCs (LSKCD150+CD48 split into Vwf-tdTomato+ and Vwf-tdTomato fractions or Gata1-eGFPLSKCD150+CD48 split into Vwf-tdTomato+/CD201+, Vwf-tdTomato+/CD201, Vwf-tdTomato/CD201+ and Vwf-tdTomato/CD201 fractions) were transplanted from each primary donor.

Smart-seq3 single-cell library preparation and sequencing

Single CD45.2+CD45.1LINcKIT+ HSPCs generated in vivo from Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs were sorted from single-HSC-transplanted mice 24–31 weeks after transplantation (Fig. 3a). In all mice, the lineage reconstitution pattern in the blood was established from at least two time points, the last one performed ≤2 weeks before the isolation of cells for Smart-seq3 processing. From mice reconstituted by Vwf-tdTomato+ P-HSCs, an unbiased isolation of all CD45.2+CD45.1 donor-derived LINcKIT+ cells was performed in combination with collection of index-sorting information. From mice reconstituted by Vwf-tdTomato multi-HSCs, in addition to collection of index-sorting information, a targeted selection of distinct donor-derived HSPCs was performed to ensure that all LINcKIT+ compartments were sufficiently represented in the Smart-seq3 analysis.

To detect potential batch variations between individual plates and individual experiments, we combined the BM samples from three 12-week-old wild-type CD45.1 mice and stored them as frozen aliquots. One BM vial was thawed for each experiment, in which each 384-well sorted plate contained 24 GMP and 24 LSK cells from this internal batch control.

As previously described23, single donor-derived LINcKIT+ cells and internal control GMP and LSK cells were sorted into 384-well plates with 3 μl of Smart-seq3 lysis buffer containing 0.5 U μl−1 RNase inhibitor (Takara), 0.1% Triton X-100 (Sigma-Aldrich), 0.5 mM dNTPs each (Thermo Fisher Scientific), 0.5 µM Smart-seq3 oligonucleotide-dT primer (5′-biotin-ACGAGCATCAGCAGCATACGAT30VN-3′, IDT) and 5% polyethylene glycol (Sigma-Aldrich). The concentrations of dNTPs, oligonucleotide-dT primer and polyethylene glycol were calculated based on a 4-μl total volume after the addition of 1 μl of reverse transcription mix. Immediately after the sort, each plate was spun down and stored at −80 °C. Generation of the cDNA library was done as previously described23, according to detailed protocols published on protocols.io (https://doi.org/10.17504/protocols.io.bcq4ivyw), with 22 cycles of PCR amplification and using 100 pg of amplified cDNA for tagmentation. The amplified tagmented libraries were pooled, bead purified and analyzed using a high-sensitivity DNA chip (Agilent Bioanalyzer) and quantified using a Qubit fluorometer (Thermo Fisher Scientific). Pooled libraries were sequenced on a NovaSeq S-prime (batch 3) or S4 (batches 1 and 2) 150 × 2-bp paired-end flow cell. FASTQ files from batch 1 were downsampled to 25% to match the sequencing depth from batches 2 and 3 to achieve similar sensitivity in gene detection across batches.

Processing and quality control of RNA-sequencing data

FASTQ files were demultiplexed and mapped to the mouse genome (Genome Reference Consortium Mouse Build 38, Ensembl release 99) using zUMIs (versions 2.9.4cq, 2.9.6 and 2.9.7)23,60, and unique molecular identifier (UMI) reads for each gene were quantified. In line with previous studies61, the exclusion of cells from analysis was based on UMI count distribution; only cells with >10,000 UMIs and <5% mitochondrial reads were included in the analysis (Supplementary Fig. 1a–c). In total, 5,014 cells derived from single transplanted Vwf-tdTomato+ P-HSCs or Vwf-tdTomato multi-HSCs were sequenced. After filtering, 4,768 cells were included in the downstream analyses. Counts were normalized using the logNormCounts function from the scran package62. After normalization, a pseudocount of 1 was added, and the data were log2 transformed.

To identify HVGs, we used the modelGeneVar function from the scran package (Supplementary Fig. 1d). Batches were then integrated using a mutual nearest-neighbor approach (batchelor::fastMNN)24 by first performing a multisample principal component analysis (PCA) and then finding the mutual nearest neighbors in this PCA space. To visualize any potential technical batch effects from each sequencing run, we plotted the internal control LSK and GMP cells, as well as the cells isolated from transplanted mice, before and after batch correction on the PCA plots (Supplementary Fig. 1e,f) used to generate the tSNE and UMAP visualization plots. Batch integration performance was evaluated using the iLISI and cLISI metrics63 (Supplementary Fig. 1g). UMAP plots were generated using a range of HVGs (500–5,000 genes; Extended Data Fig. 4c) demonstrating a preserved pattern. The 2,000 most HVGs were selected for both UMAP and tSNE plots (Supplementary Fig. 2a–c and Extended Data Fig. 4c,d).

Cell-surface marker expression was extracted from FlowJo workspaces using flowWorkspace64 to parse the gating hierarchies. Cells lacking index-sorting information were assigned as nonindexed, and cells falling outside the defined gates were classified as nongated.

Gene signature analysis

Molecular signatures were computed using the rank-based AUC score implemented in the AUCell package65. HSCs and MkPs derived from Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs were compared based on their AUC scores for previously published signatures of lineage-restricted progenitors19 and cell-cycle status66 and for multiple HSC-associated signatures, including low-output HSC26, high-output HSC26, megakaryocyte-biased HSC26, multi-HSC26, HSC1 cluster26, serial-engrafter HSC26, LT-HSC21, stem score28, surface marker overlap25, MolO25, RA-CFP-dim HSC29 and dormant HSC versus active HSC27.

Differential gene expression analysis

Differential gene expression analysis was performed with a combination of the Wilcoxon test for differences in expression levels and Fisher’s exact test for expression frequencies, as previously described67. HSCs were defined as single cells with an AUCell score of >0.22 for the MolO signature25, and MkPs were defined as single cells with an AUCell score of >0.25 for the MkP signature19. Gene expression differences with an adjusted P value (combined Fisher) of <0.05 and an absolute value of log2(fold change) of >0.5 were considered significantly differentially expressed. Genes with one or more reads were classified as expressed within single cells. HALLMARK (v2023.2) gene-set enrichment analysis on the DEGs was performed by the preranked test in GSEA software version 4.3.3, using log2(fold changes) as the rank based on genes detected in more than ten cells.

Pseudotime analysis

Total counts were normalized, and committed erythroid and myeloid progenitor cells were excluded based on >2 log2(expression) values of Car1 and Mpo, shown to define committed erythroid and myeloid lineage progenitors19, respectively (Extended Data Fig. 4a,b). The cell-cycle phase effect in the progenitor compartment was removed while retaining the difference from quiescent stem cells by regressing the difference between the S-phase score and the G2M-phase score, as described in the Seurat package68. Variable features were selected, and batches were integrated as described above before dimensionality reduction with UMAP (Supplementary Fig. 2).

Pseudotime was computed using diffusion pseudotime69 with default parameters, implemented in the SCANPY70 package through the reticulate package in R. One cell with high expression of the MolO score (cell-id: AAGCCGTTGTCCATTG) was used as the starting cell. Differential gene expression analysis on the trajectories, with diffusion pseudotime as the independent variable, was performed using the tradeSeq package36. The filtered non-normalized count matrix was used as the input. Cells were assigned to either the P-HSC or multi-HSC trajectory with weight 1. Generalized additive models were fit with the default setting of six knots. Each gene was tested for differential expression as a function of pseudotime between P-HSC and multi-HSC using the patternTest function36. The genes were ordered according to their Wald statistic for downstream analysis and visualization. DEGs were identified within the top 70 genes when comparing cells replenished by Vwf-tdTomato+ P-HSCs and Vwf-tdTomato multi-HSCs, sorted based on the Wald statistic with an adjusted P value of <0.01 and mean fold change of >1, and used to calculate the Pearson correlation between the two patterns. Correlation estimates and 95% CIs were calculated at 100 points along pseudotime. The same procedure was applied to the 70 most variable genes, representing randomly selected genes, to create a background correlation trajectory.

Reanalysis of MkPs produced by Hoxb5CreERT2-labeled HSCs in steady-state mice

A recent publication49 combined genetic fate map** in Hoxb5CreERT2/+ R26Tom/+ reporter mice (a model for specific labeling of HSCs upon tamoxifen treatment) with single-cell RNA sequencing, in which the authors sequenced (Smart-seq2 or 10× single-cell RNA sequencing) HSPCs from steady-state Hoxb5CreERT2/+ R26Tom/+ mice at different time points following recombination induced by tamoxifen. We obtained the normalized gene expression (kindly provided by the authors) and compared molecularly defined MkPs replenished at early (days 3, 7 and 12) and late (days 112, 161 and 269) time points after tamoxifen for the expression of the most highly DEGs identified between MkPs derived from Vwf-tdTomato+ P-HSCs (P-MkPs) and from Vwf-tdTomato multi-HSCs (multi-MkPs), focusing on genes with an absolute log2(fold change) value of >1 and a P value of <10−15 in the comparison between P-MkPs and multi-MkPs (Fig. 4a and Supplementary Table 4). Of the 30 genes meeting these criteria, genes not detected in the Hoxb5CreERT2 R26Tom/+ dataset were excluded, resulting in five and six genes upregulated and downregulated in P-MkPs, respectively, when compared to multi-MkPs. A one-tailed Wilcoxon test was used for the analysis, and multiple testing was adjusted using the Benjamini–Hochberg method.

In vivo Flt3Cre and VavCre fate map**

Blood and BM samples were processed as described above and analyzed using LSRII and Fortessa cytometers (BD Biosciences). See Supplementary Table 5 for antibody details.

In 8- to 13-week-old Flt3Cretg/+ R26Tom/+ Vwf-eGFPtg/+ Gata1-eGFPtg/+ steady-state mice and in recipients of single HSCs, platelets and erythrocytes were gated based on the eGFP+ signal to exclude possible debris and to identify donor-derived cells, respectively. Leukocytes were gated CD41 in steady-state and CP- and 5FU-treated mice and Vwf/Gata1-eGFP in single-HSC-transplanted mice (even if Gata1 is expressed in some myeloid cells17) to exclude transfer of the Flt3Cre-tdTomato signal from adhering platelets.

To exclude cases with inefficient Flt3Cre recombination, we checked the coexpression of cell-surface FLT3 and Flt3Cre-tdTomato in the BM of single-HSC transplantation donors. Steady-state Flt3Cre and VavCre mice were only used if erythroid, myeloid and B and T blood cells were ≥98% Flt3Cre-tdTomato+.

In vivo treatments

For CP treatment, 7- to 11-week-old Flt3Cretg/+ R26Tom/+ mice and 8- to 23-week-old VavCretg/+ R26Tom/+ mice received a single intraperitoneal injection of 200 mg kg−1 CP (European Pharmacopoeia, C3250000) or PBS control on day 0 (ref. 71). Blood was analyzed on days −3 (baseline), 4, 7, 18 and 45.

For 5FU treatment, 8- to 10-week-old Flt3Cretg/+ R26Tom/+ mice received a single intraperitoneal injection of 150 mg kg−1 5FU (Accord Healthcare) on day 0 (ref. 72). Blood was analyzed on days −7 or −2 (baseline), 5 (only blood cell counts), 10, 17 and 24. For both treatments, blood cell counts were measured with a Sysmex XP-300 automated cell counter, and peripheral blood and BM were processed as described above for transplantation experiments. CP and 5FU were obtained from the pharmacy at the Karolinska University Hospital.

For platelet depletion (anti-CD42b) treatment, 12- to 13-week-old Flt3Cretg/+ R26Tom/+ mice received a single intravenous injection of 2 mg kg−1 of an anti-CD42b antibody (R300, Emfret Analytics)73. Platelets and erythrocytes were processed as described above. For BM analysis, 11- to 12-week-old CD45.1 mice were compared on day 3 after anti-CD42b treatment to untreated CD45.1 mice. BM cell suspensions were prepared by crushing femurs, tibiae and pelvic bones from both legs of each treated and untreated mouse into PBS with 5% FCS and 2 mM EDTA. BM cells were incubated with fluorophore-conjugated CD16/32 for 15–20 min at 4 °C, followed by anti-mouse antibody staining for 15–20 min at 4 °C. BM cellularity and blood cell counts were measured with a Sysmex XP-300 automated cell counter. See Supplementary Table 5 for antibody details.

Analysis of reticulated platelets with TO

Platelet/erythrocyte cell suspensions, processed and antibody stained as described above, were incubated with 1 ml of BD Retic-COUNT reagent74 (BD Biosciences) or with 1 ml PBS with 5% FCS and 2 mM EDTA—samples and negative controls, respectively—for 1 h at room temperature in the dark (according to the manufacturer’s instructions, excluding fixation) and directly analyzed using Fortessa cytometers (BD Biosciences).

Data analysis

Statistical comparisons were performed as specified in the figure legends, using R scripts, GraphPad Prism 9 software and QuickCalcs online tools (https://www.graphpad.com/quickcalcs).

Kinetics, in which the blood of the same mouse was sampled at multiple time points (replenishment kinetics after transplantation, blood cell counts after CP treatment and Cre-driven tdTomato labeling after CP treatment), were analyzed with mixed-effects models using the lme4 R package. Mouse identity was assigned a random effect to account for intermouse variation. All combinations of time points and experimental conditions were assigned a fixed effect. Significance testing of all fixed-effects contrasts was done using the emmeans R package, and P values were adjusted using the Benjamini–Hochberg procedure.

Mixed-effects model analysis was also performed with a similar setup in the built-in statistical package of GraphPad Prism 9 for Cre-driven tdTomato labeling of reticulated platelets.

Flow cytometry data were acquired using BD FACSDiva version 9.0 software and analyzed using BD FlowJo version 10.8.1 software.

Reporting summary

Further information on research design is available in the Nature Portfolio Reporting Summary linked to this article.