• Inactivation of Suz12 results in a rapid and marked exhaustion of the HSC pool.

  • Lymphoid development is completely dependent on PRC2, but numerous myeloid lineages develop in the absence of PRC2.

Polycomb repressive complex 2 (PRC2) is a chromatin modifier that regulates stem cells in embryonic and adult tissues. Loss-of-function studies of PRC2 components have been complicated by early embryonic dependence on PRC2 activity and the partial functional redundancy of enhancer of zeste homolog 1 (Ezh1) and enhancer of zeste homolog 2 (Ezh2), which encode the enzymatic component of PRC2. Here, we investigated the role of PRC2 in hematopoiesis by conditional deletion of suppressor of zeste 12 protein homolog (Suz12), a core component of PRC2. Complete loss of Suz12 resulted in failure of hematopoiesis, both in the embryo and the adult, with a loss of maintenance of hematopoietic stem cells (HSCs). In contrast, partial loss of PRC2 enhanced HSC self-renewal. Although Suz12 was required for lymphoid development, deletion in individual blood cell lineages revealed that it was dispensable for the development of granulocytic, monocytic, and megakaryocytic cells. Collectively, these data reveal the multifaceted role of PRC2 in hematopoiesis, with divergent dose-dependent effects in HSC and distinct roles in maturing blood cells. Because PRC2 is a potential target for cancer therapy, the significant consequences of modest changes in PRC2 activity, as well as the cell and developmental stage-specific effects, will need to be carefully considered in any therapeutic context.

Hematopoiesis is the process of blood cell formation that generates approximately 1 trillion blood cells per day in humans1  that is sustained by rare HSCs resident in the bone marrow (BM). The balance between HSC self-renewal and differentiation is highly dynamic, allowing for the production of phenotypically distinct blood cells at steady state, as well as rapid expansion of specific cell types during stress. This balance is a tightly regulated process governed by both cell-extrinsic factors (eg, cytokine signaling and the BM microenvironment)2,3  and cell-intrinsic cues (eg, transcription factors and chromatin modifiers).4-6  Hence, it is not surprising that disruption to this delicate balance can result in life-threatening hematologic disorders, such as leukemia.

Polycomb group proteins are a major class of epigenetic regulators that repress target genes by coordinating changes in the chromatin structure via posttranslational modifications of histone tails. They exists in several multiprotein complexes and are important in fine-tuning gene expression.7  PRC2 consists of 3 core components: embryonic ectoderm development (Eed), suppressor of zeste 12 protein homolog (Suz12), and either enhancer of zeste homolog 2 (Ezh2) or its close homolog enhancer of zeste homolog 1 (Ezh1). Ezh1 and Ezh2 both possess histone methyltransferase activity for di- and tri-methylation of histone H3 at position lysine 27 (H3K27me2/3). Although these core factors are essential for PRC2 complex stability and function, the presence of several accessory factors including Rbbp4/7, Jarid2, Mtf2, Phf1, and Phf19 have been implicated in the recruitment of PRC2 to its target genes and modulation of its activity.8  PRC2 core components are required for embryonic development: Eed, Suz12, and Ezh2 knockout mice develop abnormally and die prior to gastrulation.9-11  Moreover, PRC2 is essential for proper differentiation of stem cells from diverse lineages including embryonic, skin, liver, and neuronal stem cells.12-17 

Conditional knockout studies in mice have enabled dissection of the roles of PRC2 in the hematopoietic system. In the lymphoid lineages, Ezh2 deletion resulted in impairment of B-lymphoid development,18-20  germinal center B-cell formation,21  and terminal differentiation of naïve CD4+ T-cells,22  but it was largely dispensable for T-lymphoid development in the thymus.23  Although several studies have shown that conditional inactivation of Ezh2 appears to have minimal impact on HSC function in adult mice, the role of Ezh2 in fetal hematopoiesis is less clear. Mochizuki-Kashio et al18  argued that fetal hematopoiesis relies exclusively on Ezh2-PRC2, and this dependency shifts toward Ezh1-PRC2 during the transition from neonates to adults, evidenced by increased Ezh1 expression and a concomitant decline in Ezh2 expression in aging HSCs. Although Ezh1 is not required for fetal hematopoiesis, Hidalgo et al24  further demonstrated that Ezh1-deficient adult mice progressively developed BM failure as a result of abnormal activation of senescence programs in HSCs. Xie et al20  recently generated an Eed conditional knockout mouse and showed that Eed is essential for adult but not fetal HSC function. HSCs overexpressing Ezh2 also possess enhanced capacity for serial transplantation in vivo, and recipient mice develop myeloproliferative-like disorders.25,26  Somewhat paradoxically, a number of studies have demonstrated that partial inhibition of PRC2 also enhances HSC function2,27  and can predispose to leukemia.28,29  Collectively, these observations illustrate that hematopoietic cells respond sensitively to altered PRC2 dosage, and any slight deviation can dramatically alter the hematopoietic system. Indeed, a spectrum of mutations in EED, EZH2, and SUZ12 has been described in many human cancers including lymphoid and myeloid malignancies, highlighting the association of altered PRC2 activity with oncogenesis.22,30-35  Thus, although PRC2 has become a compelling therapeutic target in malignancy, a more detailed understanding of the complex functional consequences of altered PRC2 activity in the hematopoietic system will be a prerequisite for effective manipulation of PRC2 activity in disease.

Here, we have generated a novel conditional knockout allele of Suz12 in mice. We show that unlike partial loss of Suz12 function, deletion of Suz12 in hematopoiesis results in dysfunction of both fetal and adult HSCs, which become unable to contribute effectively to blood cell production. Moreover, via restricted deletion in specific blood cell lineages, we establish that Suz12 is dispensable for production of myeloid cells, megakaryocytes, and platelets, while it is essential for B- and T-lymphoid development.

Mice

Experiments were approved by the Walter and Eliza Hall Institute of Medical Research Animal Ethics Committee. A targeting vector was generated carrying loxP sites flanking exon 5 of the murine Suz12 gene that included a neomycin-selection cassette flanked by 2 Flip recombinase target sites (Figure 1A). To verify correct targeting, genomic DNA from transfected C57BL/6 ES cells were digested with HindIII and the fragments were separated on 0.8% agarose gel to detect targeted clones, which was confirmed by Southern blotting (8.4 kb). A neomycin-resistant embryonic stem cell clone carrying the targeted locus was then used to generate chimeric mice. The neomycin cassette was removed by crossing to FlpE transgenic mice36  to generate the Suz12 floxed (fl) allele, herein referred to as Suz12fl. A polymerase chain reaction (PCR)-based strategy was developed for further genotyping to distinguish the wild-type (WT), FL, and recombined (Δ) alleles (PCR primers: forward: CAAAAACCTTGTCTCGGAAA, reverse: TCATCATCGGGATATGGGTA). The DelCre,37 LysMCre,38 Pf4Cre,39 Rag1Cre,40 Rosa26CreERT241  (herein referred to as CreERT2), and VavCre42  mice have been previously described. All studies were performed with mice on a C57BL/6 genetic background.

Figure 1

Conditional targeting of the mouse Suz12 gene. (A) The WT (+) Suz12 locus, indicating exons 4-7 is shown, with the fl conditional allele, and recombined knockout allele (Δ) following cre-mediated excision. The WT Suz12 locus was modified after the insertion of a targeting vector containing 2 loxP sites flanking exon 5 of the Suz12 gene, which also contained a neomycin cassette flanked by 2 Flip recombinase target sites located after the 3′-end of exon 5 and just upstream of the second loxP site. Expected fragment sizes in Southern blot analysis and PCR-based genotyping are shown. (B) PCR analysis of genomic DNA extracted from BM and spleen of Suz12/CreERT2 mice 8 days after tamoxifen treatment, using a primer combination that distinguishes WT, fl, and recombined (Δ) alleles. (C) Immunoblot analysis of Suz12 and Ezh2 protein in the thymus, spleen, and BM of Suz12fl/+/CreERT2 and Suz12fl/fl/CreERT2 mice 8 days after tamoxifen treatment. Actin was used as loading control. (D) Immunoblot analysis of Suz12, Ezh2, and H3K27me3 in cultured fetal liver cells from Suz12fl/fl/CreERT2 E14 embryos. Cell lysates were prepared 24 hours after 4-hydroxy tamoxifen (4-OH Tam]) or vehicle treatment. Total histone H3 and actin were used as loading controls.

Figure 1

Conditional targeting of the mouse Suz12 gene. (A) The WT (+) Suz12 locus, indicating exons 4-7 is shown, with the fl conditional allele, and recombined knockout allele (Δ) following cre-mediated excision. The WT Suz12 locus was modified after the insertion of a targeting vector containing 2 loxP sites flanking exon 5 of the Suz12 gene, which also contained a neomycin cassette flanked by 2 Flip recombinase target sites located after the 3′-end of exon 5 and just upstream of the second loxP site. Expected fragment sizes in Southern blot analysis and PCR-based genotyping are shown. (B) PCR analysis of genomic DNA extracted from BM and spleen of Suz12/CreERT2 mice 8 days after tamoxifen treatment, using a primer combination that distinguishes WT, fl, and recombined (Δ) alleles. (C) Immunoblot analysis of Suz12 and Ezh2 protein in the thymus, spleen, and BM of Suz12fl/+/CreERT2 and Suz12fl/fl/CreERT2 mice 8 days after tamoxifen treatment. Actin was used as loading control. (D) Immunoblot analysis of Suz12, Ezh2, and H3K27me3 in cultured fetal liver cells from Suz12fl/fl/CreERT2 E14 embryos. Cell lysates were prepared 24 hours after 4-hydroxy tamoxifen (4-OH Tam]) or vehicle treatment. Total histone H3 and actin were used as loading controls.

Close modal

Analysis of hematopoietic cells by flow cytometry

Single-cell suspensions from BM, spleen, thymus, mesenteric lymph nodes, and E13.5 fetal livers were prepared in balanced salt solution (150 mM NaCl, 3.7 mM KCl, 2.5 mM CaCl2, 1.2 mM MgSO4, 7.4 mM HEPES-NaOH, 1.2 mM KH2PO4, 0.8 mM K2HPO4) supplemented with 5% (vol/vol) fetal calf serum. For analysis of blood, red cells were lysed by treatment with lysis buffer (150 mM NH4Cl, 0.1 mM EDTA, and 12 mM NaHCO3) and the remaining cells were washed with balanced salt solution/5% fetal calf serum. Monoclonal antibodies used were: B220 [clone RA3-6B2], CD4 [GK1.5], CD8a [53-6.7], CD16/32 [24G.2], CD19 [1D3], CD25 [PC61.5], CD34 [RAM34], CD43 [S7], CD44 [IM7], CD45.1 [A20], CD45.2 [104], CD48 [HM48-1], CD127/IL7Rα [A7R34], CD135/Flt3 [A2F10], CD150 [9D1], c-Kit [2B8], Gr-1 [RB6-8C5], IgM [II/41], Mac-1 [M1-70], Ly6D [49-H4], Ly6G [Ly6G], Sca-1 [D7], and Ter119 [Ly76]. Antibodies were conjugated to Ag-presenting cell, AlexaFluor-590, AlexaFluor-700, brilliant violet 421, FITC, PE, PECy7, PerCPCy5.5, or biotin, followed by streptavidin-conjugated phycoerythrin-TexasRed (BD Biosciences, eBioscience, or Biolegend). Propidium iodide or fluorogold was used for dead cell exclusion. Samples were processed using LSR II and LSR Fortessa (BD Biosciences) instruments with data analysis performed using FlowJo software (Treestar Inc., Ashland).

Hematology

Peripheral blood was collected from the retro-orbital plexus for automated cell count (Advia3120, Bayer). Megakaryocyte and macrophage cultures are described in detail in the supplemental Methods, available on the Blood Web site.

Transplantation and tamoxifen administration

Competitive BM transplantation studies were performed using 1 × 106 test cells (CD45.2) and 1 × 106 WT competitor cells (CD45.1) from 8- to 10-week-old mice. BM cells were injected intravenously into 8- to 10-week-old irradiated CD45.1/CD45.2 or CD45.1 recipients (2 × 550 Gy). Three recipient mice were used per test donor (n = 3-4 donors). For secondary transplants, 2 × 106 cells from individual primary transplant recipients were injected into 3 CD45.1 recipients and analyzed 12 weeks later. For Suz12/CreERT2 studies, recipient mice were bled 5-weeks posttransplant to assess baseline chimerism, and 1-week later, tamoxifen (Sigma) or vehicle (10% ethanol in peanut oil; Sigma) was administered via oral gavage for 3 consecutive days (4.2 mg/day). For gene expression analysis, a single dose of tamoxifen (3 mg) was administered by gavage.

Immunoblotting

Cells were homogenized in radio immunoprecipitation assay buffer (1% NP-40, 0.1% sodium dodecyl sulfate, 0.5% sodium deoxycholate, 150 mM NaCl, 50 mM Tris-HCl) and supplemented with Complete Protease Inhibitors (Roche). Proteins were separated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, transferred to polyvinylidene fluoride membranes, and blotted with antibodies against Suz12 (P-15, Santa Cruz), Ezh2 (AC22, Cell Signaling Technology), H3K27me3 (07-449, Millipore), histone H3 (AS3, Millipore), and actin (I-19, Santa Cruz).

Statistical analysis

Data were analyzed using GraphPad Prism 5.0. When multiple statistical tests were performed, the new threshold for statistical significance was established by dividing the standard significance threshold by the number of tests (P value < .05). A two-tailed Student t test was performed in 2 group comparisons. When comparing multiple groups, one-way analysis of variance was performed, followed by Tukey’s post-hoc test.

Functional validation of Suz12 deletion in conditional mutant mice

Suz12Δ/+ mice were generated by crossing animals bearing a Suz12fl allele with DelCre transgenic mice. The DelCre transgene was bred out and intercrosses established between Suz12Δ/+ mice. No Suz12Δ/Δ offspring were weaned from these matings (supplemental Table 1), consistent with the embryonic lethality previously reported in homozygous Suz12 knockout43  and gene-trap11  mice. To confirm conditional Suz12 deletion, we generated Suz12fl/fl/CreERT2 mice, where deletion of Suz12 exon 5 can be induced in the presence of tamoxifen. For 3 consecutive days, tamoxifen (4.2 mg/day) was administered, and organs were harvested 8 days after the last injection. Genomic PCR analysis showed efficient excision of the loxP-flanked exon in both the BM and spleen (Figure 1B). Immunoblot analysis confirmed the absence of Suz12 protein and a concomitant reduction in Ezh2 protein in the BM, spleen, and thymus (Figure 1C). Similarly, when freshly isolated Suz12fl/fl/CreERT2 E13.5 lineage-depleted fetal liver cells were cultured in the presence of 4-hydroxytamoxifen (1 × 10−7 M) or vehicle (100% ethanol) for 24 hours, immunoblot analysis confirmed the absence of Suz12 protein, as well as a significant reduction in Ezh2 and global H3K27me3 levels in 4-hydroxytamoxifen–treated cells (Figure 1D).

Suz12 is required for fetal hematopoiesis

Previous reports have demonstrated that the PRC2 catalytic subunit Ezh2 is essential for fetal hematopoiesis, but is dispensable for adult BM HSC function,18  due to compensation by Ezh1.24  In contrast, a recent study suggests that Eed is not required for fetal hematopoiesis.20  To assess the specific role of Suz12 in fetal hematopoiesis, Suz12Δ/+/VavCreT mice were bred with Suz12fl/fl mice to restrict Suz12 deletion to hematopoietic cells. Genotyping revealed no Suz12fl/Δ//VavCreT mice were present at weaning (supplemental Table 2A), but these animals were present and viable at embryonic day E13.5 (supplemental Table 2B). Suz12fl/Δ//VavCreT embryos displayed significantly reduced fetal liver cellularity, and a concomitant decrease in the numbers of hematopoietic stem cells (lineage Sca-1+ c-Kit+ [LSK] CD48 CD150+) and myeloid progenitor (common myeloid progenitor [CMP], granulocyte macrophage progenitor [GMP], and megakaryocyte erythroid progenitor [MEP]) populations (Figure 2A), as well as compromised erythroid development (Figure 2B; supplemental Figure 1).

Figure 2

Suz12 is required for fetal hematopoiesis. (A) Representative plots of flow cytometric analyses of hematopoietic stem and progenitor cells in fetal livers of Suz12fl/+, Suz12fl/Δ, Suz12fl/+/VavCreT, and Suz12fl/Δ/VavCreT E13.5 embryos. Numbers from gated populations represent the proportions of total viable fetal liver cells of the indicated populations: lineage Sca-1+ c-Kit+ (LSK) cells (top panel), LSK CD150+ CD48 cells (middle panel) and myeloid progenitors (CMP, MEP, and GMP; lower panel) are shown. The absolute numbers of hematopoietic stem and progenitor cells in E13.5 fetal livers are summarized in the adjacent bar charts. (B) Representative flow cytometric plots are shown of erythroid differentiation in fetal livers of Suz12fl/+, Suz12fl/Δ, Suz12fl/+/VavCreT, and Suz12fl/Δ/VavCreT E13.5 embryos. The absolute cell numbers in each stage of erythroid differentiation are summarized in the adjacent bar chart. Data represent mean ± standard deviation from n = 5-14 fetal livers per genotype. One-way analysis of variance followed by Tukey’s post-hoc test was performed to compare differences between genotypes. *P < .05; **P < .01.

Figure 2

Suz12 is required for fetal hematopoiesis. (A) Representative plots of flow cytometric analyses of hematopoietic stem and progenitor cells in fetal livers of Suz12fl/+, Suz12fl/Δ, Suz12fl/+/VavCreT, and Suz12fl/Δ/VavCreT E13.5 embryos. Numbers from gated populations represent the proportions of total viable fetal liver cells of the indicated populations: lineage Sca-1+ c-Kit+ (LSK) cells (top panel), LSK CD150+ CD48 cells (middle panel) and myeloid progenitors (CMP, MEP, and GMP; lower panel) are shown. The absolute numbers of hematopoietic stem and progenitor cells in E13.5 fetal livers are summarized in the adjacent bar charts. (B) Representative flow cytometric plots are shown of erythroid differentiation in fetal livers of Suz12fl/+, Suz12fl/Δ, Suz12fl/+/VavCreT, and Suz12fl/Δ/VavCreT E13.5 embryos. The absolute cell numbers in each stage of erythroid differentiation are summarized in the adjacent bar chart. Data represent mean ± standard deviation from n = 5-14 fetal livers per genotype. One-way analysis of variance followed by Tukey’s post-hoc test was performed to compare differences between genotypes. *P < .05; **P < .01.

Close modal

Suz12 is required for adult HSC maintenance and self-renewal

To determine the role of Suz12 in adult hematopoiesis, we transplanted BM cells from Suz12fl/fl/CreERT2 or Suz12+/+/CreERT2 (test, CD45.2) mice, mixed with an equal number of cells from wt (competitor, CD45.1) marrow, into irradiated CD45.1/CD45.2 recipients (Figure 3A). Peripheral blood was analyzed 5 weeks after transplantation to establish the baseline chimerism, which was equivalent in both groups. A week later, tamoxifen or vehicle was administered via oral gavage and the percentage of test donor-derived cells in the peripheral blood was monitored at 6, 10, and 16 weeks after tamoxifen treatment. Recipients of the mixture of Suz12+/+/CreERT2 and WT BM receiving either tamoxifen or vehicle showed no significant changes in the level of chimerism at all 3 time points (Figure 3B, top panel). In contrast, recipients of the Suz12fl/fl/CreERT2 and wt BM mixture that were treated with tamoxifen showed progressive loss of test donor-derived CD45.2 cells over the 16-week period, an effect that was not evident in vehicle-treated mice (Figure 3B, bottom panel). Similarly, analysis of major hematopoietic organs 16 weeks after tamoxifen treatment revealed almost no contribution from test donor-derived CD45.2+ cells in tamoxifen-treated recipients of the Suz12fl/fl/CreERT2 and wt BM mixture (Figure 3D), including the BM LSK population (Figure 3C), suggesting that Suz12 is required for steady-state maintenance and function of adult HSCs. Analysis of gene expression in LSK cells and lineagec-Kit+ cells sorted 7 days after tamoxifen treatment confirmed marked induction of Cdkn2a (p16 and p19), Cdkn1a, and other PRC2 target genes (supplemental Figure 2), which was consistent with the changes in gene expression observed in Eed knockout HSCs.20  The strong induction of Cdkn2a and Cdkn1a likely contribute to the significant functional defects in Suz12-deficient HSCs.

Figure 3

Suz12 is critical for maintaining adult HSC self-renewal and homeostasis. (A) Experimental design for competitive BM transplantation assays. (B) Analysis of test donor contribution (CD45.2+) to the peripheral blood of mice receiving either Suz12fl/fl/CreERT2 or Suz12+/+/CreERT2 BM cells with tamoxifen (TAM) or vehicle (ethanol [EtOH]) treatment. (C) Representative flow cytometric plots of hematopoietic stem and progenitor cell populations in BM of mice that received Suz12fl/fl/CreERT2 donor BM 16 weeks after tamoxifen (TAM) or vehicle (EtOH) treatment. The numbers represent the proportions of cells gated from the indicated populations. (D) Analysis of test donor contribution (CD45.2+) to the spleen (SPL), mesenteric lymph nodes (MLN), thymus (THY), BM, lineage Sca-1+ c-Kit+ (LSK) cells and BM lineage Sca-1 c-Kit+ myeloid progenitors (MP) in mice receiving either Suz12fl/fl/CreERT2 or Suz12+/+/CreERT2 BM cells and tamoxifen (TAM) or vehicle (EtOH) treatment. Data represent mean ± standard deviation from Suz12+/+/CreERT2 (n = 3 donors) and Suz12fl/fl/CreERT2 (n = 4 donors) mice. Each data point represents the average test donor contribution from 3 recipients of BM from an individual donor. A two-tailed Student t test was performed to test statistical significance between vehicle (EtOH) and tamoxifen (TAM)-treated groups. BMT, bone marrow transplant; NS, nonsignificant; PB, peripheral blood. *P < .05; **P < .005; ***P < .001.

Figure 3

Suz12 is critical for maintaining adult HSC self-renewal and homeostasis. (A) Experimental design for competitive BM transplantation assays. (B) Analysis of test donor contribution (CD45.2+) to the peripheral blood of mice receiving either Suz12fl/fl/CreERT2 or Suz12+/+/CreERT2 BM cells with tamoxifen (TAM) or vehicle (ethanol [EtOH]) treatment. (C) Representative flow cytometric plots of hematopoietic stem and progenitor cell populations in BM of mice that received Suz12fl/fl/CreERT2 donor BM 16 weeks after tamoxifen (TAM) or vehicle (EtOH) treatment. The numbers represent the proportions of cells gated from the indicated populations. (D) Analysis of test donor contribution (CD45.2+) to the spleen (SPL), mesenteric lymph nodes (MLN), thymus (THY), BM, lineage Sca-1+ c-Kit+ (LSK) cells and BM lineage Sca-1 c-Kit+ myeloid progenitors (MP) in mice receiving either Suz12fl/fl/CreERT2 or Suz12+/+/CreERT2 BM cells and tamoxifen (TAM) or vehicle (EtOH) treatment. Data represent mean ± standard deviation from Suz12+/+/CreERT2 (n = 3 donors) and Suz12fl/fl/CreERT2 (n = 4 donors) mice. Each data point represents the average test donor contribution from 3 recipients of BM from an individual donor. A two-tailed Student t test was performed to test statistical significance between vehicle (EtOH) and tamoxifen (TAM)-treated groups. BMT, bone marrow transplant; NS, nonsignificant; PB, peripheral blood. *P < .05; **P < .005; ***P < .001.

Close modal

Heterozygosity for Suz12 enhances HSC self-renewal and lymphocyte production

Previous findings have shown that hematopoietic stem and progenitor cells are very sensitive to PRC2 dosage, with partial reduction2,20,28  or elevation25,26  allowing significantly increased HSC function. Moreover, mice carrying a heterozygous loss-of-function allele of Suz12 have elevated blood lymphocytes attributed to increased lymphoid progenitor activity.2,44  Consistent with these findings, HSCs from tamoxifen-treated Suz12fl/+/CreERT2 mice showed enhanced self-renewal capacity, outcompeting HSCs from vehicle-treated Suz12fl/+/CreERT2 mice in serial competitive BM transplantation assays (supplemental Figure 3A). Suz12Δ/+ mice had higher blood leukocyte counts (supplemental Figure 3B) and this phenotype was observed in various combinations of either germline or hematopoietic-specific heterozygous deletion of Suz12 (supplemental Table 3).

Suz12 is required for early lymphopoiesis

To specifically explore the role of Suz12 in lymphopoiesis, we conditionally inactivated Suz12 in the early stages of lymphoid development using Rag1Cre mice.40  Adult Suz12fl/fl/Rag1CreKI mice appeared normal and healthy, but displayed lymphopenia (supplemental Table 4) and had dramatically reduced spleen and thymus cellularity (supplemental Figure 4A). Genotyping of T-lymphocytes in the blood of Suz12fl/fl/Rag1CreKI mice suggested that the residual lymphocyte production was attributable to cells that had escaped recombination of the Suz12 fl allele (supplemental Figure 4B). Analysis of BM showed that the numbers of HSC-enriched LSK cells, lymphoid-primed multipotent progenitors, common lymphoid progenitors (CLPs), and myeloid progenitors were normal in Suz12fl/fl/Rag1CreKI mice (Figure 4A). However, B-lymphopoiesis was arrested at the pre-pro–B to pro-B transition in Suz12fl/fl/Rag1CreKI mice as shown by the absence of CD19-expressing B-lymphocytes (Figure 4B). Similarly, Suz12fl/fl/Rag1CreKI mice had defective T-lymphopoiesis (Figure 4C). The lymphoid defects were cell-intrinsic because in competitive transplantation assays production of lymphoid lineages from Suz12fl/fl/Rag1CreKI, BM was outcompeted by WT marrow with effects on both B- and T-lymphopoiesis during the transition from the CLPs to pre-pro–B cells in BM, and the DN1/ETP cells in the thymus, respectively (Figure 4D). As expected, Suz12fl/fl/Rag1CreKI BM contributed robustly to the myeloid lineage (Figure 4D).

Figure 4

Suz12 is required for B- and T-lymphopoiesis. Representative plots of flow cytometric analysis of BM from approximately 8- to 10-weeks old Suz12+/+/Rag1CreKI, Suz12fl/+/Rag1CreKI, and Suz12fl/fl/Rag1CreKI mice. Numbers represent the average proportions of (A) hematopoietic stem and progenitor cells (lineage Sca-1+ c-Kit+ [LSK]), lymphoid-primed multipotent progenitors ([LMPP], LSK Flt3hi), and common lymphoid progenitors ([CLPs], lineage Sca-1+ c-Kitint Flt3+ IL7Rα+) populations, and (B) total B cells (B220+ CD19+), pre-B + pro-B (B220+ IgM-), immature B ([Imm. b] B220lo IgM+), recirculating B ([Recirc. B] B220hi IgM+), and pre-proB (NK1.1 CD11c CD19 B220+ CD43+) populations. Data represent mean ± standard deviation, from approximately 7-12 mice per genotype. (C) Flow cytometric analysis of thymus from approximately 8- to 10-week-old Suz12+/+/Rag1CreKI, Suz12fl/+/Rag1CreKI, and Suz12fl/fl/Rag1CreKI mice. The numbers represent the proportions of cells gated from the indicated populations. Data represent mean ± standard deviation from approximately 7-12 mice per genotype. (D) Competitive BM transplantation with equal numbers of BM cells from 8-week-old Suz12+/+/Rag1CreKI or Suz12fl/fl/Rag1CreKI mice (test; CD45.2) and Suz12+/+ mice (competitor; CD45.1) transplanted into lethally irradiated CD45.1 recipient mice. Test donor-derived (CD45.2+) contribution to specific cell types was assessed in the BM (top panel), thymus (middle panel), and peripheral blood (bottom panel) 8 weeks after transplantation. Gating strategies for BM: LSK, lymphoid-primed multipotent progenitors (LMPP), common lymphoid progenitor (CLP), pre-proB, immature, and recirculating B-cell populations are defined as described above, pro-B (B220+ CD19+ c-Kit+ IgM), pre-B (B220+ CD19+ CD25+ c-Kit IgM); thymus: double negative [DN] (CD4 CD8), DN1 (DN CD44+ CD25), DN2 (DN CD44+ CD25+), DN3 (DN CD44 CD25+), DN4 (DN CD44 CD25), DP (CD4+ CD8+); blood: total test donor-derived (CD45.2+), B cells (CD19+), T cells (CD4+ and CD8+), myeloid cells ([M] Mac-1+). Data represent mean ± standard deviation from (n = 3 donor) mice. Each donor marrow is transplanted into 3 recipients per test, and the averaged mean from each test is used to calculate the final mean. A two-tailed Student t test was performed to test statistical significance between genotypes. *P < .05; **P < .005; ***P < .001.

Figure 4

Suz12 is required for B- and T-lymphopoiesis. Representative plots of flow cytometric analysis of BM from approximately 8- to 10-weeks old Suz12+/+/Rag1CreKI, Suz12fl/+/Rag1CreKI, and Suz12fl/fl/Rag1CreKI mice. Numbers represent the average proportions of (A) hematopoietic stem and progenitor cells (lineage Sca-1+ c-Kit+ [LSK]), lymphoid-primed multipotent progenitors ([LMPP], LSK Flt3hi), and common lymphoid progenitors ([CLPs], lineage Sca-1+ c-Kitint Flt3+ IL7Rα+) populations, and (B) total B cells (B220+ CD19+), pre-B + pro-B (B220+ IgM-), immature B ([Imm. b] B220lo IgM+), recirculating B ([Recirc. B] B220hi IgM+), and pre-proB (NK1.1 CD11c CD19 B220+ CD43+) populations. Data represent mean ± standard deviation, from approximately 7-12 mice per genotype. (C) Flow cytometric analysis of thymus from approximately 8- to 10-week-old Suz12+/+/Rag1CreKI, Suz12fl/+/Rag1CreKI, and Suz12fl/fl/Rag1CreKI mice. The numbers represent the proportions of cells gated from the indicated populations. Data represent mean ± standard deviation from approximately 7-12 mice per genotype. (D) Competitive BM transplantation with equal numbers of BM cells from 8-week-old Suz12+/+/Rag1CreKI or Suz12fl/fl/Rag1CreKI mice (test; CD45.2) and Suz12+/+ mice (competitor; CD45.1) transplanted into lethally irradiated CD45.1 recipient mice. Test donor-derived (CD45.2+) contribution to specific cell types was assessed in the BM (top panel), thymus (middle panel), and peripheral blood (bottom panel) 8 weeks after transplantation. Gating strategies for BM: LSK, lymphoid-primed multipotent progenitors (LMPP), common lymphoid progenitor (CLP), pre-proB, immature, and recirculating B-cell populations are defined as described above, pro-B (B220+ CD19+ c-Kit+ IgM), pre-B (B220+ CD19+ CD25+ c-Kit IgM); thymus: double negative [DN] (CD4 CD8), DN1 (DN CD44+ CD25), DN2 (DN CD44+ CD25+), DN3 (DN CD44 CD25+), DN4 (DN CD44 CD25), DP (CD4+ CD8+); blood: total test donor-derived (CD45.2+), B cells (CD19+), T cells (CD4+ and CD8+), myeloid cells ([M] Mac-1+). Data represent mean ± standard deviation from (n = 3 donor) mice. Each donor marrow is transplanted into 3 recipients per test, and the averaged mean from each test is used to calculate the final mean. A two-tailed Student t test was performed to test statistical significance between genotypes. *P < .05; **P < .005; ***P < .001.

Close modal

Suz12 is dispensable for the differentiation of myeloid and megakaryocytic lineages

We have previously shown that partial loss of Suz12 was able to ameliorate thrombocytopenia in c-Mpl−/− mice.2  Although it was not possible in these previous studies to dissect contributions from stem cell and platelet-specific effects, here we used Pf4CreT transgenic mice39  to specifically delete Suz12 in megakaryocytes and platelets. Both Suz12fl/+/Pf4CreT and Suz12fl/fl/Pf4CreT mice displayed normal numbers of platelets (supplemental Table 5). The number of BM megakaryocytes was also unchanged in Suz12fl/fl/Pf4CreT mice, and platelet turnover, as assessed by thiazole orange staining, was normal (supplemental Figure 5). In addition, Lin progenitors isolated from the BM of Suz12fl/fl/Pf4CreT mice were able to efficiently differentiate into mature megakaryocytes, in which near absolute deletion of Suz12 was evident, using a short-term in vitro culture system (supplemental Figure 6A). Suz12 deficient megakaryocytes had significantly reduced levels of H3K27me3 and displayed modest activation of PRC2 target genes (supplemental Figure 6B-D). Similarly, in Suz12fl/fl/LysMCreT mice, in which Suz12 was deleted specifically in granulocytes and macrophages,38  no differences in the numbers of mature neutrophils, monocytes, or eosinophils were evident relative to control littermates (supplementary Table 6). Suz12fl/fl/LysMCreT BM cells were also able to differentiate into mature macrophages in vitro in the absence of Suz12. Genomic PCR analysis from BM-derived macrophages showed efficient recombination of the fl allele (supplemental Figure 7A), which was further supported by immunoblot analysis showing decreased levels of Suz12, Ezh2, and global H3K27me3 (supplemental Figure 7B).

The recent discoveries of recurring mutations in PRC2 core components EZH2, EED, and SUZ12 in various hematologic malignancies30-35  have underscored the need for more precise understanding of PRC2 function in normal and malignant hematopoiesis. Although a number of studies have documented roles for PRC2 in murine hematopoiesis, these reports paint a complex picture with seemingly diverse effects of manipulating expression of different PRC2 components, cell type-specific effects and profound biological consequences of apparently subtle alterations in PRC2 complex activity. Contributing to this complexity is the functional redundancy between Ezh2-PRC2 and Ezh1-PRC2. Loss of Ezh2-PRC2 can be compensated by Ezh1-PRC2, and vice versa, in HSCs and in other stem cells.14,17,18,24,45 Ezh2 expression in HSCs declines with age, whereas Ezh1 expression increases, and homozygous deletion of Ezh2 in adult BM appears to have minimal impact on HSC function mainly due to compensation by Ezh1-PRC2,18,20  which appears to be the dominant enzymatic component for PRC2 in adult HSCs. Therefore, it is not surprising that Ezh1-deficient HSCs are unable to engraft in lethally irradiated mice, and mice with hematopoietic-specific deletion in Ezh1 eventually develop progressive BM failure.24  To circumvent the complexities of Ezh1/Ezh2 redundancy, we generated Suz12 conditional knockout mice for the inactivation of PRC2. In adult chimeric mice with a mixture of Suz12-conditional and WT hematopoiesis, induced deletion of Suz12 resulted in a near complete absence of Suz12-null hematopoietic cells in the BM and peripheral organs accompanied by a precipitous of loss of Suz12-deleted HSCs, establishing a critical requirement for Suz12 in the maintenance of adult HSCs and their capacity to generate blood. Suz12-deficient HSCs activate expression of classic PRC2 target genes, including Cdkn2a (p16 and p19) and Cdkn1a, which likely explains their marked functional impairment. Our data are consistent with that of a recent study using Eed conditional knockout mice,20  emphasizing a critical role for PRC2 in adult HSCs.

Although our results establish an important consensus with previous studies on the role of PRC2 in adult HSCs, differences remain regarding the role of PRC2 in fetal hematopoiesis. Although Ezh1 appears to be dispensable for fetal hematopoiesis,24  Mochizuki-Kashio et al18  found that deletion of Ezh2 (driven by Tie2Cre) led to compromised HSC function and defective erythroid differentiation at E12.5. However, a later study by Xie et al20  suggested that deletion of Ezh2 via VavCre conditional targeting had no obvious fetal hematopoietic defect, and the authors concluded that Ezh2 is dispensable for both adult and fetal hematopoiesis. The use of different cre-recombinases and Ezh2 conditional alleles could be the simplest explanation for the contradictory findings. Xie et al20  also found that fetal hematopoiesis in embryos with VavCre-mediated loss of Eed was undisturbed, but that Eed was critical after birth, with neonatal Eed−/− mice displaying pancytopenia, anemia, and BM failure. In contrast, we show here that VavCre-mediated deletion of Suz12 during fetal hematopoiesis significantly reduced the number of long-term HSCs and myeloid progenitors, as well as severely compromised erythroid development. These data prompt the surprising conclusion that fetal HSC are highly dependent on Suz12 without the requirement for Eed. New evidence has recently emerged that supports the existence of a noncanonical complex that contains Ezh1 and Suz12 but lacks Eed.46  A second, and more intriguing hypothesis is that Suz12 may regulate fetal hematopoiesis independent of Ezh1/Ezh2. Support for PRC2-independent actions of core PRC2 components has emerged with a recent report that Ezh2 modulates transcription in a non–PRC2-dependent fashion in prostate cancer depending on its phosphorylation state.47 

Increased blood lymphocyte numbers were observed in Suz12 heterozygous mice generated by ubiquitous (DelCre) or hematopoietic-specific (VavCre) cre-recombinase–mediated gene deletion, consistent with our previous study in mice carrying the hypomorphic Suz12Plt8 allele.44  Although previous studies have clearly demonstrated impaired lymphopoiesis as a result of complete loss of PRC2 core components, we have extended this finding by showing that this effect is due to PRC2 loss in specific progenitor cells, independent of actions in HSCs. Although PRC2 inactivation in maturing lymphocytes had little impact on B- or T-cell numbers,19,23  we demonstrate here that Rag1Cre-mediated inactivation of Suz12 in CLPs results in severe defects in both B- and T-lymphoid development. The contrasting increase in self-renewal of B-lymphoid precursors in mice with partial loss of PRC2 function44  is consistent with a specific role for PRC2 in lymphoid progenitor cells and further suggests that the influence of PRC2 on lymphoid development is highly sensitive to gene dosage, as it is in HSCs.

PRC2 appears to be dispensable in maturing granulocytes and macrophages, as well as in megakaryocytes and platelets. Although we have not identified the precise stage in myelopoiesis at which dependency on Suz12/PRC2 is lost, this insensitivity to PRC2 loss in the myeloid lineages is intriguing, and is in stark contrast to the absolute requirement for PRC2 during lymphopoiesis. The inability of lymphoid cells to tolerate deletion of Suz12 is likely due to the involvement of PRC2 at a number of key points during differentiation. For example, Ezh2 is required for effective TCR-signaling in developing thymocytes23  and for rearrangement of the immunoglobulin heavy chain (IgH) gene in B-lymphoid progenitors.19  Indeed, committed B-lymphocytes that have completed immunoglobulin rearrangement tolerate loss of Ezh2.19  It may be that repressive marks established at key target genes by PRC2 in early hematopoietic stem and progenitor cells are sufficient to enable myeloid differentiation, but not lymphoid differentiation, or it may be that there is compensation by other epigenetic factors that are specific to the myeloid lineage. Thus, in addition to a critical requirement for PRC2 in HSCs, our data establish additional lineage-specific roles in committed hematopoietic cells.

Although a profound loss of PRC2 activity results in defective adult HSC function, we and others have previously shown that a partial reduction in the activity of PRC2 via genetic manipulation of Eed, Ezh2, or Suz12 results in enhanced HSC self-renewal in serial transplantation assays,2,20,27,28  which are findings we have confirmed here in mice with induced heterozygous deletion of Suz12 in adult mice. Interestingly, Ezh2-overexpressing murine HSCs display increased proliferation and self-renewal, and eventually caused a myeloproliferative disorder similar to that found in humans.25,26  Collectively, these studies establish that PRC2 dosage needs to be tightly regulated for hematopoietic homeostasis, with slight perturbations resulting in pathological consequences. Indeed, in myeloid malignancies, the observation that heterozygous loss-of-function mutations in EZH2, EED, and SUZ12 are more frequent than homozygous deletion30,32,34  implies that PRC2 depletion favors the initial clonal expansion of leukemic stem cells, whereas complete PRC2 loss abrogates self-renewal and proliferation, as was observed in MLL-AF9–driven leukemic stem cells.48  With PRC2 increasingly being considered as a potential target for cancer therapy, and EZH2 inhibitors currently under evaluation,49,50  the biological consequences of altered PRC2 activity, as well as the cell and developmental stage-specific effects evident from mouse genetic studies, such as those presented here, will need to be carefully considered in any therapeutic context.

The online version of this article contains a data supplement.

The publication costs of this article were defrayed in part by page charge payment. Therefore, and solely to indicate this fact, this article is hereby marked “advertisement” in accordance with 18 USC section 1734.

The authors thank Stefan Glaser for Rosa26CreERT2 mice and Stephen Nutt for Rag1Cre mice; Takashi Ushiki, Ashley Ng, and Rhys Allan for helpful discussions; Jason Corbin, Jasmine McManus, Janelle Lochland, Lauren Wilkins, Keti Stoev, Elizabeth Kyran, Hayley Backhaus, Carolina Alvarado, and Liana Mackiewicz for skilled technical assistance and animal husbandry; and Dr Lachlan Whitehead for assistance with immunofluorescence image analysis.

This work was supported by program (1016647 and 575500) and project (1011663) grants, fellowships (W.S.A., I.J.M., M.E.B.), and Independent Research Institutes Infrastructure Support Scheme grant (361646) from the Australian National Health and Medical Research Council, fellowships from the Australia Research Council (M.E.B.), an Australian Postgraduate Award (S.C.W.L.), the Carden Fellowship (D.M.) of the Cancer Council, Victoria, the Australian Cancer Research Fund, and Victorian State Government Operational Infrastructure Support.

Contribution: S.C.W.L., I.J.M., and W.S.A. designed experiments and wrote the manuscript; S.C.W.L., S.M., C.H., M.K., M.L., E.C.J., L.D.R., D.M., and I.J.M. performed experiments and analyzed data; M.E.B. and S.A.K. provided critical conceptual input and feedback on data interpretation and the manuscript; and I.J.M. and W.S.A. supervised research.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Ian J. Majewski, Cancer and Haematology Division, The Walter and Eliza Hall Institute of Medical Research, 1G Royal Parade, Parkville, VIC 3052, Australia; e-mail: majewski@wehi.edu.au.

1
Doulatov
 
S
Notta
 
F
Laurenti
 
E
Dick
 
JE
Hematopoiesis: a human perspective.
Cell Stem Cell
2012
, vol. 
10
 
2
(pg. 
120
-
136
)
2
Majewski
 
IJ
Blewitt
 
ME
de Graaf
 
CA
, et al. 
Polycomb repressive complex 2 (PRC2) restricts hematopoietic stem cell activity.
PLoS Biol
2008
, vol. 
6
 
4
pg. 
e93
 
3
Adams
 
GB
Deconstructing the hematopoietic stem cell niche: revealing the therapeutic potential.
Regen Med
2008
, vol. 
3
 
4
(pg. 
523
-
530
)
4
Butler
 
JS
Dent
 
SY
The role of chromatin modifiers in normal and malignant hematopoiesis.
Blood
2013
, vol. 
121
 
16
(pg. 
3076
-
3084
)
5
Paralkar
 
VR
Weiss
 
MJ
Long noncoding RNAs in biology and hematopoiesis.
Blood
2013
, vol. 
121
 
24
(pg. 
4842
-
4846
)
6
Wilkinson
 
AC
Göttgens
 
B
Transcriptional regulation of haematopoietic stem cells.
Adv Exp Med Biol
2013
, vol. 
786
 (pg. 
187
-
212
)
7
Sauvageau
 
M
Sauvageau
 
G
Polycomb group proteins: multi-faceted regulators of somatic stem cells and cancer.
Cell Stem Cell
2010
, vol. 
7
 
3
(pg. 
299
-
313
)
8
Margueron
 
R
Reinberg
 
D
The Polycomb complex PRC2 and its mark in life.
Nature
2011
, vol. 
469
 
7330
(pg. 
343
-
349
)
9
Faust
 
C
Lawson
 
KA
Schork
 
NJ
Thiel
 
B
Magnuson
 
T
The Polycomb-group gene eed is required for normal morphogenetic movements during gastrulation in the mouse embryo.
Development
1998
, vol. 
125
 
22
(pg. 
4495
-
4506
)
10
O’Carroll
 
D
Erhardt
 
S
Pagani
 
M
Barton
 
SC
Surani
 
MA
Jenuwein
 
T
The polycomb-group gene Ezh2 is required for early mouse development.
Mol Cell Biol
2001
, vol. 
21
 
13
(pg. 
4330
-
4336
)
11
Pasini
 
D
Bracken
 
AP
Jensen
 
MR
Lazzerini Denchi
 
E
Helin
 
K
Suz12 is essential for mouse development and for EZH2 histone methyltransferase activity.
EMBO J
2004
, vol. 
23
 
20
(pg. 
4061
-
4071
)
12
Aoki
 
R
Chiba
 
T
Miyagi
 
S
, et al. 
The polycomb group gene product Ezh2 regulates proliferation and differentiation of murine hepatic stem/progenitor cells.
J Hepatol
2010
, vol. 
52
 
6
(pg. 
854
-
863
)
13
Chamberlain
 
SJ
Yee
 
D
Magnuson
 
T
Polycomb repressive complex 2 is dispensable for maintenance of embryonic stem cell pluripotency.
Stem Cells
2008
, vol. 
26
 
6
(pg. 
1496
-
1505
)
14
Ezhkova
 
E
Pasolli
 
HA
Parker
 
JS
, et al. 
Ezh2 orchestrates gene expression for the stepwise differentiation of tissue-specific stem cells.
Cell
2009
, vol. 
136
 
6
(pg. 
1122
-
1135
)
15
Pasini
 
D
Bracken
 
AP
Hansen
 
JB
Capillo
 
M
Helin
 
K
The polycomb group protein Suz12 is required for embryonic stem cell differentiation.
Mol Cell Biol
2007
, vol. 
27
 
10
(pg. 
3769
-
3779
)
16
Pereira
 
JD
Sansom
 
SN
Smith
 
J
Dobenecker
 
MW
Tarakhovsky
 
A
Livesey
 
FJ
Ezh2, the histone methyltransferase of PRC2, regulates the balance between self-renewal and differentiation in the cerebral cortex.
Proc Natl Acad Sci USA
2010
, vol. 
107
 
36
(pg. 
15957
-
15962
)
17
Shen
 
X
Liu
 
Y
Hsu
 
YJ
, et al. 
EZH1 mediates methylation on histone H3 lysine 27 and complements EZH2 in maintaining stem cell identity and executing pluripotency.
Mol Cell
2008
, vol. 
32
 
4
(pg. 
491
-
502
)
18
Mochizuki-Kashio
 
M
Mishima
 
Y
Miyagi
 
S
, et al. 
Dependency on the polycomb gene Ezh2 distinguishes fetal from adult hematopoietic stem cells.
Blood
2011
, vol. 
118
 
25
(pg. 
6553
-
6561
)
19
Su
 
IH
Basavaraj
 
A
Krutchinsky
 
AN
, et al. 
Ezh2 controls B cell development through histone H3 methylation and Igh rearrangement.
Nat Immunol
2003
, vol. 
4
 
2
(pg. 
124
-
131
)
20
Xie
 
H
Xu
 
J
Hsu
 
JH
, et al. 
Polycomb repressive complex 2 regulates normal hematopoietic stem cell function in a developmental-stage-specific manner.
Cell Stem Cell
2014
, vol. 
14
 
1
(pg. 
68
-
80
)
21
Béguelin
 
W
Popovic
 
R
Teater
 
M
, et al. 
EZH2 is required for germinal center formation and somatic EZH2 mutations promote lymphoid transformation.
Cancer Cell
2013
, vol. 
23
 
5
(pg. 
677
-
692
)
22
Zhang
 
Y
Kinkel
 
S
Maksimovic
 
J
, et al. 
The polycomb repressive complex 2 governs life and death of peripheral T cells.
Blood
2014
, vol. 
124
 
5
(pg. 
737
-
749
)
23
Su
 
IH
Dobenecker
 
MW
Dickinson
 
E
, et al. 
Polycomb group protein ezh2 controls actin polymerization and cell signaling.
Cell
2005
, vol. 
121
 
3
(pg. 
425
-
436
)
24
Hidalgo
 
I
Herrera-Merchan
 
A
Ligos
 
JM
, et al. 
Ezh1 is required for hematopoietic stem cell maintenance and prevents senescence-like cell cycle arrest.
Cell Stem Cell
2012
, vol. 
11
 
5
(pg. 
649
-
662
)
25
Herrera-Merchan
 
A
Arranz
 
L
Ligos
 
JM
de Molina
 
A
Dominguez
 
O
Gonzalez
 
S
Ectopic expression of the histone methyltransferase Ezh2 in haematopoietic stem cells causes myeloproliferative disease.
Nat Commun
2012
, vol. 
3
 pg. 
623
 
26
Kamminga
 
LM
Bystrykh
 
LV
de Boer
 
A
, et al. 
The Polycomb group gene Ezh2 prevents hematopoietic stem cell exhaustion.
Blood
2006
, vol. 
107
 
5
(pg. 
2170
-
2179
)
27
Majewski
 
IJ
Ritchie
 
ME
Phipson
 
B
, et al. 
Opposing roles of polycomb repressive complexes in hematopoietic stem and progenitor cells.
Blood
2010
, vol. 
116
 
5
(pg. 
731
-
739
)
28
Lessard
 
J
Schumacher
 
A
Thorsteinsdottir
 
U
van Lohuizen
 
M
Magnuson
 
T
Sauvageau
 
G
Functional antagonism of the Polycomb-Group genes eed and Bmi1 in hemopoietic cell proliferation.
Genes Dev
1999
, vol. 
13
 
20
(pg. 
2691
-
2703
)
29
Richie
 
ER
Schumacher
 
A
Angel
 
JM
Holloway
 
M
Rinchik
 
EM
Magnuson
 
T
The Polycomb-group gene eed regulates thymocyte differentiation and suppresses the development of carcinogen-induced T-cell lymphomas.
Oncogene
2002
, vol. 
21
 
2
(pg. 
299
-
306
)
30
Ernst
 
T
Chase
 
AJ
Score
 
J
, et al. 
Inactivating mutations of the histone methyltransferase gene EZH2 in myeloid disorders.
Nat Genet
2010
, vol. 
42
 
8
(pg. 
722
-
726
)
31
Morin
 
RD
Johnson
 
NA
Severson
 
TM
, et al. 
Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin.
Nat Genet
2010
, vol. 
42
 
2
(pg. 
181
-
185
)
32
Nikoloski
 
G
Langemeijer
 
SM
Kuiper
 
RP
, et al. 
Somatic mutations of the histone methyltransferase gene EZH2 in myelodysplastic syndromes.
Nat Genet
2010
, vol. 
42
 
8
(pg. 
665
-
667
)
33
Ntziachristos
 
P
Tsirigos
 
A
Van Vlierberghe
 
P
, et al. 
Genetic inactivation of the polycomb repressive complex 2 in T cell acute lymphoblastic leukemia.
Nat Med
2012
, vol. 
18
 
2
(pg. 
298
-
301
)
34
Score
 
J
Hidalgo-Curtis
 
C
Jones
 
AV
, et al. 
Inactivation of polycomb repressive complex 2 components in myeloproliferative and myelodysplastic/myeloproliferative neoplasms.
Blood
2012
, vol. 
119
 
5
(pg. 
1208
-
1213
)
35
Simon
 
C
Chagraoui
 
J
Krosl
 
J
, et al. 
A key role for EZH2 and associated genes in mouse and human adult T-cell acute leukemia.
Genes Dev
2012
, vol. 
26
 
7
(pg. 
651
-
656
)
36
Takeuchi
 
T
Nomura
 
T
Tsujita
 
M
, et al. 
Flp recombinase transgenic mice of C57BL/6 strain for conditional gene targeting.
Biochem Biophys Res Commun
2002
, vol. 
293
 
3
(pg. 
953
-
957
)
37
Schwenk
 
F
Baron
 
U
Rajewsky
 
K
A cre-transgenic mouse strain for the ubiquitous deletion of loxP-flanked gene segments including deletion in germ cells.
Nucleic Acids Res
1995
, vol. 
23
 
24
(pg. 
5080
-
5081
)
38
Clausen
 
BE
Burkhardt
 
C
Reith
 
W
Renkawitz
 
R
Förster
 
I
Conditional gene targeting in macrophages and granulocytes using LysMcre mice.
Transgenic Res
1999
, vol. 
8
 
4
(pg. 
265
-
277
)
39
Tiedt
 
R
Schomber
 
T
Hao-Shen
 
H
Skoda
 
RC
Pf4-Cre transgenic mice allow the generation of lineage-restricted gene knockouts for studying megakaryocyte and platelet function in vivo.
Blood
2007
, vol. 
109
 
4
(pg. 
1503
-
1506
)
40
McCormack
 
MP
Forster
 
A
Drynan
 
L
Pannell
 
R
Rabbitts
 
TH
The LMO2 T-cell oncogene is activated via chromosomal translocations or retroviral insertion during gene therapy but has no mandatory role in normal T-cell development.
Mol Cell Biol
2003
, vol. 
23
 
24
(pg. 
9003
-
9013
)
41
Anastassiadis
 
K
Glaser
 
S
Kranz
 
A
Berhardt
 
K
Stewart
 
AF
A practical summary of site-specific recombination, conditional mutagenesis, and tamoxifen induction of CreERT2.
Methods Enzymol
2010
, vol. 
477
 (pg. 
109
-
123
)
42
Croker
 
BA
Metcalf
 
D
Robb
 
L
, et al. 
SOCS3 is a critical physiological negative regulator of G-CSF signaling and emergency granulopoiesis.
Immunity
2004
, vol. 
20
 
2
(pg. 
153
-
165
)
43
Lee
 
TI
Jenner
 
RG
Boyer
 
LA
, et al. 
Control of developmental regulators by Polycomb in human embryonic stem cells.
Cell
2006
, vol. 
125
 
2
(pg. 
301
-
313
)
44
Lee
 
SC
Phipson
 
B
Hyland
 
CD
, et al. 
Polycomb repressive complex 2 (PRC2) suppresses Eμ-myc lymphoma.
Blood
2013
, vol. 
122
 
15
(pg. 
2654
-
2663
)
45
Ezhkova
 
E
Lien
 
WH
Stokes
 
N
Pasolli
 
HA
Silva
 
JM
Fuchs
 
E
EZH1 and EZH2 cogovern histone H3K27 trimethylation and are essential for hair follicle homeostasis and wound repair.
Genes Dev
2011
, vol. 
25
 
5
(pg. 
485
-
498
)
46
Xu
 
J
Shao
 
Z
Li
 
D
, et al. 
Developmental control of polycomb subunit composition by GATA factors mediates a switch to non-canonical functions.
Mol Cell
2015
, vol. 
57
 
2
(pg. 
304
-
316
)
47
Xu
 
K
Wu
 
ZJ
Groner
 
AC
, et al. 
EZH2 oncogenic activity in castration-resistant prostate cancer cells is Polycomb-independent.
Science
2012
, vol. 
338
 
6113
(pg. 
1465
-
1469
)
48
Neff
 
T
Sinha
 
AU
Kluk
 
MJ
, et al. 
Polycomb repressive complex 2 is required for MLL-AF9 leukemia.
Proc Natl Acad Sci USA
2012
, vol. 
109
 
13
(pg. 
5028
-
5033
)
49
McCabe
 
MT
Ott
 
HM
Ganji
 
G
, et al. 
EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations.
Nature
2012
, vol. 
492
 
7427
(pg. 
108
-
112
)
50
Qi
 
W
Chan
 
H
Teng
 
L
, et al. 
Selective inhibition of Ezh2 by a small molecule inhibitor blocks tumor cells proliferation.
Proc Natl Acad Sci USA
2012
, vol. 
109
 
52
(pg. 
21360
-
21365
)

Author notes

I.J.M. and W.S.A. contributed equally to this study.

Sign in via your Institution