Next Article in Journal
Understanding the Allosteric Modulation of PTH1R by a Negative Allosteric Modulator
Next Article in Special Issue
High Intrinsic Oncogenic Potential in the Myc-Box-Deficient Hydra Myc3 Protein
Previous Article in Journal
Rescue of Misfolded Organic Cation Transporter 3 Variants
Previous Article in Special Issue
Tumor Growth Remains Refractory to Myc Ablation in Host Macrophages
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lessons from Using Genetically Engineered Mouse Models of MYC-Induced Lymphoma

Department of Biochemistry, Center for Molecular Biomedicine (CMB), Friedrich Schiller University Jena, 07745 Jena, Germany
*
Author to whom correspondence should be addressed.
Present Address: Josep Carreras Leukaemia Research Institute (IJC), Campus ICO-Germans Trias i Pujol, 08916 Badalona, Spain.
Cells 2023, 12(1), 37; https://doi.org/10.3390/cells12010037
Submission received: 4 November 2022 / Revised: 6 December 2022 / Accepted: 15 December 2022 / Published: 22 December 2022
(This article belongs to the Special Issue MYC Signaling in Cancer)

Abstract

:
Oncogenic overexpression of MYC leads to the fatal deregulation of signaling pathways, cellular metabolism, and cell growth. MYC rearrangements are found frequently among non-Hodgkin B-cell lymphomas enforcing MYC overexpression. Genetically engineered mouse models (GEMMs) were developed to understand MYC-induced B-cell lymphomagenesis. Here, we highlight the advantages of using Eµ-Myc transgenic mice. We thoroughly compiled the available literature to discuss common challenges when using such mouse models. Furthermore, we give an overview of pathways affected by MYC based on knowledge gained from the use of GEMMs. We identified top regulators of MYC-induced lymphomagenesis, including some candidates that are not pharmacologically targeted yet.

1. Introduction

As observed in most human malignancies, overexpression of MYC leads to fatal cellular metabolism, growth, and signaling deregulation, which defines MYC as a classical oncogene [1,2]. MYC rearrangements, copy number amplifications, or mutations are frequently found among non-Hodgkin B-cell lymphomas (B-NHL) and enforce MYC overexpression [3]. More specifically, roughly 80% of Burkitt’s lymphoma (BL), 15% of diffuse large B-cell lymphoma (DLBCL), and 2% of follicular lymphoma (FL) are characterized by MYC translocations [3,4,5].
Despite the progress made in the past decades, cancer heterogeneity, insufficient therapy response, and relapse are still risks for patients. As directly targeting MYC with chemical compounds is challenging [6], exploring genetic vulnerabilities in MYC-induced B-cell lymphoma is expected to uncover new “attack points” for cancer treatment.
This review highlights the advantages of genetically engineered mouse models (GEMMs) to understand MYC-induced lymphomagenesis. We focus on the widely used Eµ-Myc transgenic mouse model and discuss the most common challenges when using this model system. We give here an overview of GEMMs of MYC-induced B-cell lymphoma, revisiting over 170 GEMMs with distinct genetic alterations. By thoroughly compiling these data, we were able to identify top regulators of MYC-induced lymphomagenesis, including some candidates that are not pharmacologically targeted yet. At last, we discuss future directions in using GEMMs in B-cell lymphoma research.

2. MYC and the Origin of the Eµ-Myc Mouse Model

The transcription factor (TF) MYC is one of the most prominent proto-oncogenes and is composed of a basic helix–loop–helix (bHLH) motif, followed by a leucine zipper that enables its DNA binding and five MYC homology boxes (0-IV), which facilitate stability, protein–protein interactions, and transcriptional regulation [7,8]. Moreover, MYC harbors a nuclear localization signal (NLS) to ensure nuclear import [9]. Here, MYC binds to defined areas in the chromatin. The consensus sequence of the enhancer (E)-box region bound by MYC is characterized by a CACGTG motif and can be found in one-third of MYC binding loci in human B-cells [10]. However, non-canonical E-box (CACATG) and E-box-independent mechanisms contribute to the binding of MYC to roughly 5000 promoter sites in B-cell lymphoma [10,11].
In murine naïve B-cells, MYC was associated with regulatory elements pre-existing in a poised transcriptional state, which is crucial for fast responses toward immunological stimuli [12]. Here, MYC and its interaction with key transcription factors of B-cell identity are essential for cell cycle entry of germinal center (GC) B-cells and memory B-cell formation [13,14,15]. To achieve elevated mRNA translation, MYC regulates the expression of genes encoding transfer RNAs, ribosomal RNAs, and components of the spliceosome [12,16,17,18]. In line with this, murine MYC-deficient B-cells fail to amplify RNA synthesis after stimulation with lipopolysaccharide (LPS) due to impaired chromatin reorganization, which impedes TF occupancy at key promoters [19]. Taken together, physiological MYC is a crucial regulator of transcription and translation in normal cells and a potent driver of B-cell lymphomagenesis in mice and humans (Figure 1).
In 1985, Adams and colleagues constructed seven different transgenic mouse strains by inserting the gene encoding MYC into distinct regulatory regions [23]. From these, constitutive expression of Myc under the immunoglobulin enhancers Eκ or Eµ led to lymphoma development originating from B-cells with dramatically increased incidence [23]. Eµ-Myc transgenic mice possess three Myc transgenes on chromosome 19 and two somatic Myc copies on chromosome 15 and are usually bred in a heterozygous manner [24,25]. If all copies are expressed, or some are silenced similar to BL cell lines, is not clear [26,27]. However, the onset of lymphoma is dose-dependent, as homozygous crossings of Eµ-Myc transgenic mice show dramatically reduced survival [28].
Deregulation of Myc transcription in Eµ-Myc transgenic mice occurs stepwise exceeding physiological levels of MYC in pre-tumor cells [12,29,30]. The high abundance of MYC leads to its accumulation at active promoters and enhancers, causing transcriptional amplification by recruitment of factors that phosphorylate RNA polymerase II at serine 2 to speed up transcription, also known as the “general amplifier hypothesis” [16]. Moreover, oncogenic MYC invades low-affinity E-boxes and target sites without E-box motifs [16,31], which can be explained by the biophysical properties of dimeric bHLH domains to non-specifically bind DNA [32].
The “general amplifier hypothesis” is complemented by selective transcriptional regulation of key target genes, generating fatal feedback loops that shape gene expression [12]. These complex pathophysiological mechanisms explain why the expression of an identified core target gene signature for MYC comprising 51 genes could not sufficiently explain how MYC transforms healthy B-cells [33]. In line with the famous “two-hit” hypothesis for tumor development [34], overexpression of MYC in B-cells would generally induce apoptosis or senescence, making the acquisition of secondary mutations necessary for oncogenic transformation [35,36].
At the cellular level, B-cell-restricted overexpression of Myc promotes the accumulation of IgM-negative B-cells in the bone marrow [37]. Moreover, peripheral blood from Eµ-Myc transgenic mice allows tracking of large B220low cells over time, correlating with disease onset and progression [28,38]. Evidence exists that B220low cells are the actual pre-tumorigenic cell type, as they are also found in spleens of Eµ-Myc transgenic mice [29,38]. In wild-type mice, B220low cells are mainly found in the bone marrow but not in the spleen, representing either immature B- or plasma cells [39,40].
The median survival of Eµ-Myc transgenic mice was initially determined to be 11 weeks [23], which enables fast survival analysis of cohorts. Diseased Eµ-Myc transgenic mice develop massively enlarged lymph nodes at inguinal, brachial, and cervical sites, splenomegaly, and sometimes thymoma or bowel obstruction due to tumor masses in the abdomen; this was initially described as “multicentric lymphosarcoma with associated leukemia” [23,41]. Later descriptions of diseased Eµ-Myc transgenic mice included behavioral alterations such as “inactivity, lack of grooming, and/or cachexia” [28], which can now be monitored automatically using body temperature, weight, and food and water intake [42]. Cachexia or “wasting syndrome” can occur in Eµ-Myc transgenic mice without visible lymphoma formation, but with significant weight loss [43]. Taken together, the Eµ-Myc mouse model is a useful system for studying B-cell lymphomagenesis and tumor growth, as molecular changes result in phenotypic alterations.

3. B-Cell Lymphomas from Eµ-Myc Transgenic Mice Arise in a Competent Immune System and Are Highly Heterogeneous

Eµ-Myc transgenic mice possess a competent immune system with all innate and adaptive immune cell types [36,38,44]. The singular deletion of α/β T-cells and natural killer (NK) T-cells or γ/δ T-cells, which are known for their anti-tumor actions, did not affect the survival of Eµ-Myc transgenic mice [45]. This appears to be the case for lymphomas from Eµ-Myc transgenic mice with mutations of the tumor suppressor p53 [36]. In contrast, lymphoma cells overexpressing BCL-2 were immunologically visible and removed by CD8+ T and NK cells [36]. Therefore, this mouse model might be helpful in studying the role of immune cells in counteracting malignant transformation.
The competent immune system enables applications from the growing field of immuno-oncology, including checkpoint inhibitors, cancer vaccines, CAR-T cells, or oncolytic viruses to encounter cancer. Surface expression of PD-1 receptors was found to be increased on cytotoxic T-cells in lymphoma-bearing Eµ-Myc transgenic mice [46], while B-cells from Eµ-Myc transgenic mice showed an upregulation of the inhibitory receptor PD-L1 to prevent T-cell-mediated elimination [25]. The use of chemotherapy and CAR-T-cells against CD19 increased the survival of Eµ-Myc lymphoma cell xenografts, although normal B-cells were diminished as well [47]. Another approach is the “vaccination” of wild-type mice with α-galactosylceramide-loaded (a synthetic CD1d-dependent NK T-cell ligand) and irradiated Eµ-Myc lymphoma cells to lower tumor burden and increase survival after transplantation of malignant cells from Eµ-Myc transgenic mice [48]. A combination of this “vaccine” and an antibody against the immune checkpoint receptor 4–1BB (CD137) resulted in long-term protective effects [49]. Therefore, Eµ-Myc transgenic mice are valuable tools for evaluating potential therapeutic approaches involving the training of the immune system to detect and erase malignant cells.
Lymphomas arising in Eµ-Myc transgenic mice show differences compared to IgM-positive human BL, as constitutive Myc expression occurs from the earliest B-cell progenitor on. Therefore, disease does not necessarily involve GC formation, and all stages of B-cell development can form the tumor cell of origin [36,50]. Mixed phenotypes might be related to multiple events of malignant transformation or cellular de-differentiation rather than “active” differentiation of malignant cells. This is in line with the idea that MYC promotes proliferation but inhibits differentiation [20]. When a specific gene knock-out (KO) blocks or impairs normal B-lymphopoiesis, combination with the Eµ-Myc transgene results in an expansion of the affected B-cell subpopulation. Examples include KO of Msh2 (encoding DNA damage repair protein MSH2) or Prkaa1 (encoding AMP-activated protein kinase AMPKα1) resulting in merely pro-B or pre-B lymphomas [45,51] or, on the contrary, KO of Ube3a (encoding Ubiquitin-Protein-Ligase E3A), which shifts the phenotype towards mature B-cell lymphomas [52].
There is an ongoing debate about using mouse experiments to model human diseases and therapeutic approaches. It became clear in recent years that the experimental design, the mouse model, and the characterization of the patient cohort are critical to obtain transferable data from mouse studies. Eµ-Myc-derived lymphomas are very heterogenous; for example, a study described that individual B-cell lymphomas shared only a quarter of all differentially expressed genes [12]. One would expect that Eµ-Myc lymphomas are similar to human BL due to the Eµ-Myc transgene. However, gene expression signatures close to human DLBCL, including the germinal center-like and activated B-cell-like subtype, were also identified [53]. As DLBCL can be driven by BCL6 translocations or hyperactive RAS signaling [54], the appearance of frequent mutations in Kras or Bcor, which encodes a repressor of BCL-6, as tertiary drivers in Eµ-Myc lymphomas might explain DLBCL-like gene expression signatures [24].
Transgenic mouse models have been widely used to study signaling pathways and test therapeutic agents to understand oncogenic mechanisms. For example, the deregulated mTORC1 signaling in Eμ-Myc lymphomagenesis can be used by treatment with the mTORC inhibitor everolimus, even at the pre-malignant stage, to delay lymphoma onset [55]. These findings from mouse experiments have been validated in human B-cell lymphoma later, including BL and DLBCL, and led to promising clinical trials [56,57,58].
A recent study showed that lymphomas from Eµ-Myc transgenic mice exposed to chemotherapy recapitulate the gene expression profiles of patients suffering from DLBCL. The authors identified a senescence-associated gene signature, termed “SUVARness”, and elevated H3K9me3 marks that predict a favorable outcome in DLBCL patients [59].
These examples clearly show that drug screening in mouse models can be successfully translated into human therapies and that the Eµ-Myc mouse model is a valuable tool for studying human B-cell lymphomas due to similarities in genetic alterations.

4. Critical Genes for MYC-Induced Lymphomagenesis Share Common Pathways

More than 300 peer-reviewed studies involving Eµ-Myc transgenic mice have been published (Figure 2A). We identified 172 different GEMMs presented with survival curves based on the Eµ-Myc transgene (Table 1). These models included additional genetic perturbations, such as complete KOs, conditional KOs, or overexpression by transgene insertion (Figure 2B). A minority of research studies worked with specific (point) mutations or xenografts created with fetal liver or lymphoma cells (Figure 2B).
Roughly 60% of studies analyzed the survival of Eµ-Myc transgenic mice bred to full KOs, resulting in the deletion of the respective gene throughout the entire body. However, this can disturb immuno-surveillance, angiogenesis, or metabolism by affecting other cells. In contrast, only 15% of the studies were based on cell type-specific KOs using B-cell-specific Cre recombinases, such as CD19-cre (active in pro B-cells) or Mb1-cre (active in pre-pro B-cells) to eliminate the target gene flanked by loxP sites [60,61]. This seems favorable for investigating B-cell lymphomagenesis, but a careful interpretation is needed when normal B-cell development is perturbed by the KO, as a decreased B-cell population stochastically lowers malignant transformation events. Moreover, escaping from genetic inactivation can occur under certain conditions during lymphomagenesis, skewing the obtained survival curves [62]. A more sophisticated approach would be the combination of the Eµ-Myc transgene with Cre recombinases that delete in germinal center B-cells, such as AID-cre, CD21-cre, or Cγ1-cre to study mature B-cell lymphomas with the respective genetic deletion [63,64,65].
To rule out strain- or model-dependent effects, when comparing the survival of all these different studies, a percentage normalized to the used control cohort was calculated. By considering all genes where deletion or overexpression had a significant impact on mouse survival (either <50% reduction or >200% increase in life span), we identified 48 critical genes for MYC-induced lymphomagenesis (Figure 2C,D). This also implies that more than two-thirds of all studies did not observe an effect on survival, as defined by our criteria. Common biological functions among the 48 critical genes were identified, being “regulation of transcription by RNA polymerase II”, “chromatin organization”, “histone modification”, and “regulation of catabolic processes” based on the gene ontology (GO) terms (Figure 3). In addition, “signal transduction by p53”, “apoptotic signaling pathway”, and “DNA damage response” were found as common biological pathways among these critical proteins. The contribution of these pathways in MYC-induced B-cell lymphomagenesis will be discussed in detail in this chapter.

4.1. Epigenetic Modifiers Cooperate with MYC to Maintain Uncontrolled Proliferation

Three distinct fail-safe mechanisms prevent cells from becoming cancer: (1) Induction of cell cycle arrest via p21 through activated p53, (2) sequestering of cell cycle promoting factors such as E2F proteins through pRB (retinoblastoma protein), and (3) repression of anti-apoptotic factors or activation of pro-apoptotic factors to induce apoptosis.
GEMMs with B-cell-specific MYC overexpression and deletions or mutations in the p53/ARF/MDM2-axis uniformly showed reduced survival, as expected [69,70,71,72,73,74]. Of note, whole-body KO of Trp53 (encoding p53) can result in T-cell lymphomas, and investigators should be careful to delineate the distinct effects [75]. Interestingly, B-cell-specific KO of Trp53 with Mb1-cre resulted in the development of B-cell lymphomas harboring oncogenic translocations, including Igh/Myc—very similar to Eµ-Myc transgenic mice—and a median survival of 28 days [76]. Lymphomas were also formed when Trp53 was knocked out using CD21-Cre, although these tumors lacked elevated MYC expression [77]. On the contrary, deleting one allele of Rb barely altered the survival properties of Eµ-Myc transgenic mice [71], while a critical role was attributed to levels of E2F in MYC-induced lymphomagenesis [78].
The pRB/E2F-axis is controlled through the interplay of cyclins and cyclin-dependent kinases (CDKs) [79]. MYC directly activates the expression of the gene encoding Cyclin D2, which is essential for driving cell proliferation [80]. RAS, a commonly found tertiary driver in Eµ-Myc lymphomas [24], can activate Cyclin D1 [81]. Cyclin E, for example, is positively regulated by the RNA helicase DDX3, encoded by the sex chromosomes [82]. Tumor formation was heavily impaired in females but not males lacking the X-linked allele Ddx3x using a B-cell-specific KO model crossed to Eµ-Myc transgenic mice [62].
Epigenetic modifiers were shown to be highly involved in impacting the pRB/E2F-axis in Eµ-Myc transgenic mice. This includes the histone acetyltransferase GCN5 (Kat2a), whose homozygous loss resulted in a decreased expression of the genes encoding E2F and Cyclin D and improved survival by almost three-fold [83]. Similarly, heterozygous loss of the gene encoding the histone acetyltransferase MOZ (Kat6a) slowed down the proliferation of malignant B-cells, and pharmacological inhibition of MOZ reduced E2F2 transcription, stopping tumor growth in vivo [84,85].
Cyclin D1 expression was furthermore shown to be regulated by the methyl transferase EZH2 [86]. More specifically, a gain-of-function mutant of EZH2 (Y641F), which is also found among B-cell lymphoma patients [54,87], acted in concert with wild-type EZH2 to elevate activating H3K27 trimethylation marks in Eµ-Myc transgenic mice, resulting in enforced B-cell receptor signaling and lethality [86].
In contrast, the histone methyltransferase SUV39H1 creates repressive histone marks by tri-methylation of H3K9 and directly interacts with pRB, making it crucial for the repression of genes encoding cyclins [88,89]. Therefore, the loss of Suv39h1 alleles reduced the lifespan of Eµ-Myc transgenic mice by 50%, also due to defects in senescence induction in cancer cells [90].
These molecular interactions show how MYC sustains proliferation and overrides cell cycle checkpoints to circumvent cell-intrinsic safeguard programs. Interestingly, the latest subclassification of DLBCL contains a cluster of human B-cell lymphomas, driven by deregulated methyl transferases, such as EZH2 or MLL4 [54,87], that often occur together with MYC and BCL-2 alterations. Many “epi-drugs” exist that target epigenetic regulators in B-NHL [91], which could be further utilized in therapy for MYC-dependent lymphoma or lymphomas with MYC as the secondary driver.

4.2. Impairing Direct MYC Interaction Partners Is Most Effective in Prolonging Survival

MYC has many interaction partners (Figure 4), constantly interchanging through competitive binding. MAX (MYC-associated factor X), for example, dimerizes with MYC under physiological conditions at E-boxes through bHLH domains to stimulate the transcription of target genes [7]. The MYC:MAX interaction is crucial for target gene transcription and high MYC levels, as MAX-deficient B-cells had unstable MYC, and MAX-deficient Eµ-Myc transgenic mice significantly extended survival [92]. Therefore, the small molecule inhibitor 10058-F4, which blocks dimerization between MYC and MAX, was expected to be a breakthrough in targeting MYC-dependent cancer [93]. However, 10058-F4 did not achieve adequate results due to high turnover and the demand for high molar concentrations in vivo [6]. New inhibitors that disrupt the MYC:MAX interaction are MYCi975 and MYCMI-7, and they obtained promising results [94,95].
Binding partners of MYC are functional mediators but also might regulate protein stability (Figure 4). The half-life of MYC is the oncogene’s weak spot, as half of all MYC must be synthesized newly every 30 min, and emphasizes its strong regulation [96]. Phosphorylation of MYC at serine residue 62 is induced by ERK downstream of RAS, positively regulating the protein stability of MYC [97]. In contrast, MYC phosphorylation at threonine residue 58 via GSK3 is destabilizing and frequently mutated in B-cell lymphoma to prevent phosphorylation and decay [98,99]. Furthermore, the isomerase PIN1 alters the configuration of proline residue 63, which sterically shields phosphorylated serine 62 from dephosphorylation, eventually protecting MYC from degradation [100]. This explains why KO of PIN1 significantly increased the survival of Eµ-Myc transgenic mice due to lowered MYC levels [101].
Figure 4. Common effectors of MYC-induced lymphomagenesis physically interact. Critical genes from Figure 2 were clustered based on the physical interaction of the encoding protein (line thickness indicates the strength of data support) of the encoded protein using the STRING database [102]. Additional common interaction partners are shown. Terms describe biological function of some proteins in the clusters.
Figure 4. Common effectors of MYC-induced lymphomagenesis physically interact. Critical genes from Figure 2 were clustered based on the physical interaction of the encoding protein (line thickness indicates the strength of data support) of the encoded protein using the STRING database [102]. Additional common interaction partners are shown. Terms describe biological function of some proteins in the clusters.
Cells 12 00037 g004
Recently, we discovered a novel way to target the stability of MYC by exploiting its constant shuttling from the cytoplasm to the nucleus [44]. MYC is well-known for associating with microtubules [103], and the post-translational acetylation of tubulin is a crucial factor for assembly and stability [104,105]. The histone deacetylase 6 (HDAC6) mediates tubulin deacetylation, and pharmacological inhibition of HDAC6 decreases MYC protein levels, likely by heat-shock protein-mediated proteolysis [44]. Similarly, the broader MYC interaction network could be further utilized for therapeutic interventions by targeting factors that connect central pathways, such as MIZ-1 (Zbtb17), EED, or DDX3X (Figure 4).

4.3. MYC-Induced Apoptosis Might Be Triggered by Transcriptional Stress and DNA Damage

Apoptosis can be seen as the result of an altered balance between anti- and pro-apoptotic effectors. Therefore, KO of genes encoding anti-apoptotic mediators, such as MCL-1 or BCL-w, had a positive effect on the survival of Eµ-Myc transgenic mice because malignant cells could not maintain high levels of anti-apoptotic proteins and initiated apoptosis [106,107,108]. Loss of pro-apoptotic mediators such as BIM or PUMA in lymphoma cells had the opposite effect on survival, as expected [109,110,111]. The interplay of MYC and other TFs was thought to modulate the expression of apoptotic mediators in a stoichiometric way. For example, the promoter of BCL2 is antagonistically controlled by MYC and MIZ-1, implying that this interaction balances apoptosis induction [112,113].
However, a newer hypothesis assumes that “MYC-driven apoptosis results from RNA Polymerase II stalling” and not from direct transcriptional control of apoptotic mediators [114]. Overexpression of MYC leads to replicative stress, which can stall RNA Polymerase II [115,116]. Two phenomena appear to be the primary source of this type of stress: (I) transcription-replication conflicts caused by the crashing of replication and transcription machinery during the S phase due to fast replication and high transcriptional output [117]. (II) R-loops, which are transient DNA-RNA hybrids formed by nascent RNA transcripts and act as a barrier to replication fork progression [118,119]. Both stressors may generate DNA lesions and genomic instability, which trigger the DNA damage repair machinery.
In particular, the DNA damage sensor CHK1 seems crucial in MYC-induced lymphomagenesis, as B-cell-specific heterozygous KO completely abrogated malignant transformation [120]. CHK1-haploinsuffiency correlated with higher γ-H2A.X levels and PARP cleavage, concluding cell death induction due to unrepaired DNA damage [120]. Activated ATM and high γ-H2A.X levels were also found when the acetyltransferase TIP60 (Kat5) was partially knocked out in Eµ-Myc transgenic mice [121]. Interestingly, TIP60 modulates DNA damage pathways and can modify MYC post-translationally, increasing MYC’s protein stability [122].
A decrease in active, phosphorylated CHK1 was found in Eµ-Myc lymphomas lacking functional MIZ-1, indicating an impaired activation of DNA damage response [50]. Other interactors of MYC, such as BRCA1, might be involved in resolving R-loops and, thus, protect from tumorigenesis [123,124]. A functional DNA damage response might be necessary to establish MYC-dependent tumors because apoptosis induction is prevented. However, more experimental evidence might be required to fully understand the connection between MYC, transcriptional stress, and DNA damage repair pathways.

4.4. Why Is Targeting MYC So Effective but Difficult to Realize?

MYC overexpression is related to many hallmarks of cancer, ranging from immunosuppression to metabolic and epigenetic reprogramming [1]. Importantly, MYC-overexpressing cancer cells show an “oncogene addiction” to MYC, which implies that targeted inactivation leads to tumor regression [125]. Selective pressure, however, enforces mechanisms such as post-translational modifications or mutations to sustain high MYC levels, which eventually results in tumor relapse [125]. It appears evident that targeting MYC in cancer cells would reverse all oncogene-induced effects, restore physiological cell state, or induce apoptosis. Even relapse might be circumvented when MYC levels are entirely eradicated. However, all cells need physiological levels of MYC or other members of the MYC family. Here, the factors controlling MYC stability and translocation into the nucleus are the key to finding new pharmacological targets.
The proteasomal degradation of MYC can be initiated, for example, by constant HDAC6 inhibition [44]. A reduction by half was sufficient to induce apoptosis and cell cycle arrest in B-NHL cells but did not result in lymphoma remission in mice [44]. A critical factor for successful therapy is the immunological visibility of the tumor. As MYC-driven cancers suppress the autocrine secretion of factors mediating senescence or immune cell invasion [126,127], combining the insights from targeting MYC at the protein level and activating the immune surveillance could define future therapies. Still, the impact of epigenetic reprogramming in MYC-induced lymphoma is unknown regarding persistent changes in gene expression, even after uncoupling from the oncogene MYC.
Another approach to prevent MYC-induced malignant transformation would be to target essential factors involved in this process. Notably, all 48 critical genes for murine MYC-induced lymphomagenesis are expressed in human lymphoid tissue and most are deregulated in the corresponding tumor (Figure 5A). In addition, some genes are frequently mutated in human B-NHL (Figure 5B), leading to significantly reduced progression-free survival when mutated (Figure 5C). Further research is needed to address how mouse model findings can be used to prevent or treat human disease.

5. Outlook

The mechanisms of how MYC drives malignant transformation are well studied. However, there are still under-investigated areas in MYC-induced lymphomagenesis, such as the role of the three-dimensional chromatin organization or the fatty acid metabolism. Studying the chromatin organization beyond the levels of nucleosomes might further explain the origin of the characteristic MYC translocations found in human B-cell lymphoma [131]. First, chromosome loop anchors are fragile sites for genetic rearrangements in B-cells [132]. Second, removing insulators between strong enhancers and oncogenes might allow the ectopic oncogene expression observed in cancer [133]. Third, the increased frequency of R-loops that occur directly at the MYC locus was also associated with MYC translocations [134].
MYC overexpression in cancer was accompanied by remodeling of the glycerophospholipid metabolism [135,136]. For example, loss of the lipoxygenase ALOX12 dramatically decreased the survival of Eµ-Myc transgenic mice [137]. On the contrary, the loss of one Myc allele extended the lifespan of mice characterized by a healthier lipid metabolism [138]. We recently discovered that MYC-induced lymphomagenesis increased the levels of certain polyunsaturated fatty acids, which was associated with elevated mTORC1 activity and impaired autophagy [30]. It would be interesting to test if interfering with lipid remodeling could prevent MYC-induced lymphomagenesis.
At last, the combination of an established MYC-driven cancer mouse model, such as Eµ-Myc (Table 1), with modern-day technologies, such as spatial single-cell sequencing, might be helpful to resolve intercellular differences in MYC expression tracking the acquisition of secondary mutations in vivo and to clarify the role of MYC in the tumor microenvironment All these efforts will eventually be rewarded with a deeper understanding of MYC-dependent lymphomagenesis, pointing toward future therapies.
Table 1. List of all mouse models analyzed combined with the Eµ-Myc transgene and their respective survival.
Table 1. List of all mouse models analyzed combined with the Eµ-Myc transgene and their respective survival.
NameGeneFunctionModelSurvival[%]Ref.
µMT (IgM heavy chain)IghmReceptorFull KO* CTRL: 120 d,
KO: 80 d
66.67 [45]
4E-BP1Eif4ebp1TranslationDox-inducible KO* CTRL: 90 d,
KO: 145 d
161.1[139]
A1/BFL-1
Bcl2a1aApoptosis(a) Full KO
(b) Transplantation of tamoxifen-inducible KO cells
(c) Constitutive miR-shRNA (KD)
(a) CTRL: 92 d,
KO: 94 d
(b) Vehicle: 17 d, Tamoxifen: 23 d
(c) CTRL: 103 d,
KD: 109 d
(a) 102.2
(b) 135.3
(c) 105.8
[140,141]
AID
Aicda
DNA damage and repairFull KO(a) no effect
(b) CTRL: 112 d,
KO: 130 d
(a) 100
(b) 116.1
[140,141]
ALOX12Alox12MetabolismFull heterozygous KO
CTRL: 220 d,
+/−: 70 d
31.8[137]
AMD1Amd1
MetabolismTransplanted shRNA transduced FL cells * CTRL: 112 d,
KO: 70 d
62.5[142]
EIF5AEif5aTranslationTransplanted shRNA transduced FL cells* CTRL: 112 d,
KO: 56 d
50[142]
AMPKα1Prkaa1SignalingFull KO
CTRL: 10 wks,
KO: 7 wks
70[51]
APAF1Apaf1ApoptosisTransplanted FL cells from Eµ-Myc full KO miceNo effect100[143]
ATF2Atf2Transcription factorCD19-cre, B-cell-specific KONo effect100[144]
ATF4Atf4Transcription factorTamoxifen-inducible* Vehicle: 40 d, Tamoxifen: 80 d200[145]
ATF7Atf7Transcription factorCD19-cre, B-cell-specific KOWT: 105 d,
KO: 135 d
128.6[144]
BADBadApoptosisFull KOCTRL: 138 d,
WT: 78 d
56.5[146]
BAXBaxApoptosisFull KOCTRL: 21.7 wks
WT: 12.6 wks
58.1[146]
BCL-2Bcl2
Apoptosis(a) Full heterozygous KO (b) Transplanted FL cells from Eµ-Myc full KO mice(a) CTRL: 116 d,
KO: 154 d
(b) No effect
(a) 132.8
(b) 100
[147,148]
BCL-WBcl2l2ApoptosisFull KOCTRL: 90 d,
KO: 298.5 s
331.7[106]
BCL-xBcl2l1Apoptosis(a) Full heterozygous KO
(b) Transplantation of tamoxifen-inducible KO cells
(a) CTRL: 116 d,
KO: 174 d
(b) CTRL: 19 d,
KO: 25 d
(a) 150
(b) 131.6
[107,148]
BIF-1Sh3glb1ApoptosisFull KOCTRL: 107 d,
KO: 65 d
60.7[149]
BIKBikApoptosisFull KONo effect100[150]
BIMBcl2l11
Apoptosis(a) Full KO
(b) Mb1-cre, B-cell-specific KO
(a) CTRL: 15 wks,
KO: 8.2 wk
(b) CTRL: 72 d,
KO: 113 d
(a) 54.7
(b) 63.7
[109,110]
BMFBmfApoptosisFull KOCTRL: 138 d,
KO: 87 d
63[151]
BMI1Bmi1
Epigenetic regulator(a,b) Full heterozygous KO
(c) Transplanted overexpressing FL cells
(a) * CTRL: 150 d, KO: >300 d
(b) * CTRL: 100 d, KO: >250 d
(c) CTRL: >300 d,
OE: 74 d
(a) >200
(b) >250
(c) 24.7
[152,153,154]
BOKBokApoptosisFull KOCTRL: 107 d;
KO: 121 d
113.1[155]
BTK/TECBtk
Tec
SignalingFull heterozygous KO: BTK+/− TEC+/−CTRL: 100 d,
KO: 60 d
60[156]
BUB1 Bub1PTMOverexpression of
point mutant (T85)
CTRL: 21 wks,
MUT: 13 wks
61.9[157]
CAMLCamlSignalingSubcutaneous transplant of tamoxifen-inducible full KOVehicle: 7 d, Tamoxifen: >25 d
>357[158]
Caspase 9Casp9ApoptosisFL transplantation of full KO cellsCTRL: 57 wk;
KO: 54 wk
94.7[143]
Caspase 2Casp2ApoptosisFull KOCTRL: 16 wks,
KO: 8 wks
50[159]
CBX7Cbx7Epigenetic regulatorFL cells with overexpressionCTRL: >300 d,
OE: 43 d
<14.3[154]
CD19Cd19ReceptorFull KOCTRL: 13.4 wks,
KO: 24.3 wks
181.3[43]
CDK4Cdk4pRB-axisFull KOCTRL: 18 wks,
KO: 11 wks
61.1[160]
CHK1Chek1DNA damage and repair(a) Full heterozygous KO
(b) Mb1-cre, B-cell-specific KO
(a) CTRL: 106 d,
KO: 205 d
(b) CTRL: 106 d, KO: >350 d
(a) 193
(b) >330
[120]
CKS1Cks1bpRB-axisFull KOCTRL: 91 d,
KO: 268 d
294.5[161]
CREBBPCrebbpEpigenetic regulatorAID-cre + immunization* CTRL: 85 d,
KO: 55 d
64.7[162]
cRELRelTranscription factorFull KOCTRL: 115 d,
KO: 79 d
68.7[163]
CSN6Cops6PTMFull heterozygous KO* CTRL: 100 d,
KO: 190 d
190[164]
CUL9Cul9PTMFull KOCTRL: 126.4 d,
KO: 85.1 d
67.3[165]
DDX3XDdx3xHelicase(a) CD19-cre, B-cell-specific KO
(b) Vav-cre, B-cell-specific KO
(a) ♂: CTRL: 83 d, KO: 105 d;
♀: CTRL: 87 d, KO: 212 d
(b) ♂: CTRL: 98 d, KO: >350 d
♀: CTRL: 110.5 d; KO: 83 d
(a)♂: 126.5;
♀: 243.7
(b)♂: >357.1;
♀: 75.1
[62]
DICER
Dicer1SplicingCD19-cre, B-cell-specific KOWT: 194 d,
KO: 351 d
180.9[166]
DMP1Dmp1p53-axisFull KO* CTRL: 22 wks,
KO: 13 wks
59.1[167]
DNMT3BDnmt3bEpigenetic regulatorFull heterozygous KO * CTRL: 125 d,
KO: 75 d
60[168]
DPY30Dpy30Epigenetic regulatorFull heterozygous KO CTRL: 121 d,
KO: 180.5 d
149.2[169]
E2F1E2f1pRB-axis(a,b) Full KO(a) * CTRL: 24 wks,
KO: 16 wks
(b) No effect
(a) 150
(b) 100
[78,170]
E2F2E2f2pRB-axisFull KOWT: 126 d, KO: 63 d50[78]
E2F3E2f3pRB-axisFull KONo effect100[78]
E2F4E2f4pRB-axisFull KOCTRL: 110 d,
KO: 375 d
340.9[78]
E6APUbe3aPTMFull heterozygous KO CTRL: 103 d,
KO: 153 d
148.5[52]
EZH2Ezh2Epigenetic regulator(a) GOF mutant
(b) Transplanted shRNA transduced FL cells
(a) CTRL: 137.5 d, MUT: 51 d
(b) CTRL: 220 d, KD:55 d
(a) 37.1
(b) 25
[86,153]
FNIP1Fnip1MetabolismFull KO* CTRL: 110 d, KO: >300 d>272.7[171]
FOXOFoxo4TFDominant negative mutant, transplanted transduced FL cells* CTRL: >250 d,
MUT: 50 d
<20[172]
GCN2Eif2ak4TranslationTransplanted tamoxifen-inducible lymphoma cells No effect100[145]
GCN5Kat2aEpigenetic regulatorCD19-cre, B-cell-specific KOCTRL: 21 wks,
KO: 58.4 wks
278.1[83]
H2A.XH2axEpigenetic regulatorFull KONo effect100[25]
HDAC1Hdac1Epigenetic regulatorMb1-cre, B-cell-specific KO CTRL: 161 d,
KO: 170 d
105.6[173]
HDAC2Hdac2Epigenetic regulatorMb1-cre, B-cell-specific KOCTRL: 161 d,
KO: 164 d
101.9[173]
IBTKIbtkSignalingFull KOCTRL: 90 d,
KO: 150 d
166.7[174]
ID2Id2TFFull KO No effect100[175]
IL6R (gp130)Il6raReceptorFL xenograft with CD19-cre deleted cellsCTRL: 277 d,
KO: 20 d
7.2[176]
IL7RIl7r
Receptor(a) LOF (no activation of survival mechanism)
(b) Transplanted cells
(a) CTRL: 15.5 wks, MUT: 66.5 wks
(b) No effect
(a) 429
(b) 100
[177]
INK4A/P16Cdkn2ap53-axisFull heterozygous KO * CTRL: 150 d,
KO: 45 d
30[152]
INK4C/P18Cdkn2cp53-axisFull KONo effect100[175]
KLRK1Klrk1ReceptorFull KOWT: 22 wks,
KO: 15 wks
68.2[178]
KSR1Ksr1SignalingFull KOCTRL: 95 d,
KO: 138 d
145.3[179]
L24Rpl24TranslationFull heterozygous KO * CTRL: 100 d,
KO: 210 d
210[180]
L38Rpl38Epigenetic regulatorFull heterozygous KO * CTRL: 70 d,
KO: 110 d
157.1[180]
LGLLlgl1CytoskeletonFull KONo effect100[181]
MAD2Mad2l1Spindle assemblyTransplanted HSCs with overexpression* CTRL: >350 d,
OE: 60 d
<17.1[182]
MAXMaxTFMb1-cre, B-cell-specific KO CTRL: 97 d,
KO: 300 d
309.3[92]
MCL1Mcl1Apoptosis(a) CD19-cre, B-cell-specific KO
(b) Rag1-cre, heterozygous KO
(c) Transplanted tamoxifen-inducible lymphoma cells
(d) Transgene (H2K promoter)
(e) Transgene (VavP promoter)
(a) CTRL: 91 d,
KO: 123 d
(b) CTRL: 129 d,
KO: 346 d
(c) WT: 19 d,
KO: 35 d
(d) WT: 134 d,
OE: 72 d
(e) WT: 94 d,
OE: 30.5 d
(a) 135.2
(b) 268.2
(c) 184.2
(d) 53.7
(e) 32.4
[107,108,183,184]
MDM2Mdm2p53-axis(a) Full heterozygous KO
(b) Point mutation (LOF) C305F
(a) CTRL: 20.6 wks, KO: 44.3 wks
(b) CTRL: 20.7 wks
MUT: 11.6 wks
(a) 215
(b) 56.0
[69,74]
MDM4Mdm4p53-axis(a) Transgene
(b) Full heterozygous KO
(a) CTRL: 31 wks,
OE: 34 wks
(b) * CTRL: 350 d,
KO: >400 d
109.7

>114.3
[185,186]
MDMXMdmxDeleted in micePoint mutation W201S/W202G* CTRL: 170 d,
MUT: 80 d
47.1[73]
MGAMgaTFCD19-cre, B-cell-specific KOCTRL: 97 d,
KO: 87 d
89.7[187]
MHCIIH2ReceptorFull KO + immunizationNo effect100[162]
MIFMifCytokineFull KOCTRL: 2.67 months,
KO: 3.67 months
137.5[188]
miR146aMir146microRNAFull KOCTRL: 104.5 d,
KO: 82.5 d
78.9[189]
miR-17-92
Mir17hgmicroRNA(a) Transplanted overexpressing FL cells
(b) Transplanted tamoxifen-inducible KO cells
(a) * CTRL: >200 d,
OE: 125 d
(b) * CTRL: 20 d,
KO: 33 d
<62.5

165
[190,191]
MIZ-1Zbtb17TFMb1-cre, B-cell-specific KO* CTRL: 110 d,
KO: 350 d
318.2[50]
MNTMntTF(a) Full heterozygous KO
(b) Rag1-cre, KO
(a) CTRL: 17 wks
KO: 28 wks
(b) CTRL: 86 d,
KO: 463 d
(a) 538.4
(b) 164.7

[192,193]
MOZKat6aEpigenetic regulatorFull heterozygous KO CTRL: 105 d,
KO: 411 d
391.4[84]
MPLMplReceptorFull KOCTRL: 87 d,
KO: 76.5
87.9[194]
MSH2Msh2DNA damage and repair(a) Full KO
(b) Mutation (G674A)
(a) * CTRL: 100 d,
KO: 40 d
(b) * CTRL: 100 d,
MUT: 40 d
(a) 40
(b) 40
[45]
MTAP MtapMetabolismFull heterozygous KOCTRL: 130 d,
KO: 87 d
66.9[195]
MTBP Mtbpp53-axisFull heterozygous KOCTRL: 135 d,
KO: 270 d
200[196]
MYSM1Mysm1PTMTamoxifen-inducible full KO* CTRL: >150 d,
KO: 80 d
<53.3[197]
NFKB1/P105Nfkb1TFFull KONo effect100[198]
NFKB2/P100Nfkb2TFFull KOCTRL: 205 d,
KO: 171 d
83.4[199]
NOXAPmaip1ApoptosisFull KONo effect
100[111]
NucleosteminGnl3SignalingFull heterozygous KO* CTRL: 100 d,
KO: 260 d
260[200]
ODCOdc1MetabolismFull heterozygous KOCTRL: 110 d,
KO: 320 d
290.9[201]
OGG1Ogg1DNA damage and repairFull KONo effect100[202]
p19/ARFCdkn2ap53-axis(a) Full heterozygous KO
(b) Full KO
(c) Full KO
(a) * CTRL:135 d,
KO: 35 d
(b) CTRL: 20.7 wks,
KO: 10.1 wks
(c) CTRL: 89 d,
KO: 73 d
(a) 25.9
(b) 48.8
(c) 82.0
[71,72,74]
p27Cdkn1bp53-axis
Full KOCTRL: 120 d,
KO: 80 d
66.7[203]
p38Mapk14SignalingHeterozygous mutation (T180A, T182F)CTRL: 77 d,
MUT: 85
110.4[204]
p53Trp53p53-axis
(a) Full heterozygous KO
(b) Full heterozygous KO
(c) Full heterozygous KO
(d) Full KO
(e) Full heterozygous KO
(f) Point mutation LOF (G515C)
(a) CTRL: 20.6 wks,
KO: 5.6 wks
(b) * CTRL: 137.5 d,
KO: 37.5 d
(c) * CTRL: 100 d,
KO: 30 d
(d) CTRL: 89 d,
KO: 40 d
(e) CTRL: 138 d,
KO: 35 d
(f) CTRL: 138 d,
MUT: 62 d
(a) 27.2
(b) 27.3
(c) 30
(d) 44.9
(e) 25.4
(f) 44.9
[69,70,71,72,111]
P73Trp73TFFull KONo effect100[205]
PARP1Parp1ApoptosisFull KOCTRL: 127 d,
KO: 90 d
70.1[206]
PARP2Parp2ApoptosisFull KOCTRL: 127 d,
KO: 326 d
257[206]
PARP14Parp14MetabolismFull KO* CTRL: 13 wks,
KO: 20 wks
153.8[207]
PCGF6Pcgf6Epigenetic regulatorCD19-cre, B-cell-specific KOCTRL: 203 d,
KO: 65 d
32.0[187]
PFPPrf1ApoptosisFull KOCTRL: 135 d,
KO: 139 d
103[208]
PIN1Pin1IsomeraseFull KOCTRL: 108 d,
KO: 431 d
399.1[101]
PLCβ3Plcb3SignalingFull heterozygous KO* CTRL: >365 d,
KO: 100 d
<27.4[209]
PLCγ2Plcg2SignalingFull KO* CTRL 20 wks,
KO: 10 wks
50[210]
PMLPmlApoptosis, SignalingFull heterozygous KOCTRL: 103 d,
KO: 153 d
149[52]
PRDM11Prdm11Epigenetic regulatorFull KOCTRL: 113 d,
KO: 94 d
83.2[211]
PRDM15Prdm15TranscriptionTamoxifen-inducible KOCTRL: 107 d,
KO: 332 d
310.3[212]
PREP1Pknox1TFTamoxifen-inducible heterozygous KOCTRL: 58 wks,
KO: 23 wks
39.7[213]
PRMT5Prmt5RNA/SplicingTamoxifen-inducible heterozygous KO* CTRL: 90 d,
KO: 175 d
194.4[17]
PUMABbc3Apoptosis(a) Full KO
(b) Full KO
(a) CTRL: 100 d,
KO: 66 d
(b) * CTRL: 15 wks,
KO: 11 wks
(a) 66
(b) 73.3
[111,214]
RAC1Rac1SignalingTransplanted, transduced cells* CTRL: 18 d,
KD: 28 d
155.6[215]
RAG1Rag1DNA damage and repairFull KO* CTRL 110 d,
KO: 90 d
81.8[141]
RAIDDCraddApoptosisFull KO* CTRL: 120 d,
KO: 110
91.7[216]
RAP1Terf2ipSignalingFull KO* CTRL: 15 wks,
KO: 12 wks
80[217]
RAPTORRptorMetabolismCD2-cre* CTRL: 18 wks,
KO: >55 wks
>305.6[218]
pRBRb1pRB-axisFull heterozygous KO* CTRL: 135 d,
KO: 125 d
92.6[71]
RIPK3Ripk3SignalingFull KOCTRL: 118 d,
KO: 97 d
82.2[219]
RUNX1Runx1TFMx1-cre + pIpC; heterozygous for p53No effect100[220]
SAE2Uba2PTMTransplanted, transduced lymphoma cells* CTRL: 35 d,
KD: >100 d
>285.7[221]
ScribbleScribScaffoldTransplanted FL cells* CTRL: 175 d,
KO: 280 d
160[222]
Septin 4 Septin4CytoskeletonFull KO* CTRL: 270 d,
KO: 100 d
37.0[223]
Sirtuin 4Sirt4Epigenetic regulatorFull KOCTRL: 195 d,
KO: 139 d
71.3[224]
SKP2Skp2PTMFull KOCTRL: 100 d,
KO: 150 d
150[225]
SMARCAL1Smarcal1HelicaseFull KOCTRL: 187 d,
KO: 224 d
119.8[116]
SMYD2Smyd2Epigenetic regulatorCD19-cre, B-cell-specific KO* CTRL: 150 d,
KO: 175 d
116.7[226]
SUV39H1Suv39h1Epigenetic regulatorFull KO* CTRL: 125 d,
KO: 60 d
48[90]
SUZ12Suz12Epigenetic regulatorHeterozygous LOF mutationCTRL: 103 d,
MUT: 72 d
69.9[153]
TCRαTracReceptorFull KONo effect100[45]
TCRΔTrdcReceptorFull KONo effect100[45]
TEL2Etv7Deleted in miceTransplanted, transduced cells (overexpression)CTRL: >16 wks,
OE: 13 wks
81.3[227]
TIP60Kat5Epigenetic regulatorFull heterozygous KO* CTRL: 52 wks,
KO: 12 wks
23.1[121]
TIS11BZfp36l1TranscriptionEµ-Tis11b
(overexpression)
* CTRL: 140 d,
OE: 100 d
71.4[228]
TRAIL-RTnfrsf10bApoptosisFull KOCTRL: 119 d,
KO: 82 d
68.9[229]
Tristetraprolin Zfp36TranscriptionEµ-TTP
(overexpression)
(a) CTRL: 103.5 d,
OE: 194 d
(b) CTRL: 121 d,
OE: 277 d
(a) 187.4
(b) 228.9
[228]
UCH-L1Uchl1Ubiquitin system(a) Transgene
(b) Full KO
(a) * CTRL: >60 wks,
TG: 45 wks
(b) * CTRL: >60 wks,
KO: >60 wks
(a) <75
(b) 100
[230]
UNG1UngRepairFull KO* CTRL: 110 d,
KO: 85 d
77.3[202]
UTXKdm6aEpigenetic regulatorCD19-cre, B-cell-specific KO* ♂: CTRL: 145 d,
KO: 120 d;
♀: CTRL: >200 d,
KO: 70 d
♂: 82.8
♀: <35
[231]
WIP1Ppm1dSignalingFull KOCTRL: 77 d,
KO: 138 d
179.2[204]
WRNWrnHelicaseMutation in helicase domain CTRL: 115 d,
KO: 151 d
131.3[232]
XPO1Xpo1Nuclear exportPoint mutation (E571K), tamoxifen-inducibleCTRL: 35 d,
KO: 28 d
80[233]
ZMAT3Zmat3TranscriptionFull KOCTRL: 125 d,
KO: 93 d
74.4[234]
ZRANB3Zranb3HelicaseFull KOCTRL: 104 d,
KO: 138 d
132.7[116]
Relative survival was calculated based on the control (CTRL) cohort. A symbol (*) designates estimated median survival based on the presented survival curve.

Author Contributions

Writing—original draft preparation, R.W. and C.K.; writing—review and editing, R.W., E.-M.P. and C.K.; visualization, R.W.; supervision, C.K.; funding acquisition, R.W., E.-M.P. and C.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Carl Zeiss Foundation (to R.W.), the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation), GRK 1715/grant number 177710919 (to C.K.) and by a Landesgraduiertenstipendium, Friedrich Schiller University Jena (to E.-M.P.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef] [PubMed]
  2. Dang, C.V. MYC on the Path to Cancer. Cell 2012, 149, 22–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Lin, P.; Medeiros, L.J. The Impact of MYC Rearrangements and “Double Hit” Abnormalities in Diffuse Large B-Cell Lymphoma. Curr. Hematol. Malig. Rep. 2013, 8, 243–252. [Google Scholar] [CrossRef] [PubMed]
  4. Chaudhary, S.; Brown, N.; Song, J.Y.; Yang, L.; Skrabek, P.; Nasr, M.R.; Wong, J.T.; Bedell, V.; Murata-Collins, J.; Kochan, L.; et al. Relative Frequency and Clinicopathologic Characteristics of MYC-Rearranged Follicular Lymphoma. Hum. Pathol. 2021, 114, 19–27. [Google Scholar] [CrossRef]
  5. Molyneux, E.M.; Rochford, R.; Griffin, B.; Newton, R.; Jackson, G.; Menon, G.; Harrison, C.J.; Israels, T.; Bailey, S. Burkitt’s Lymphoma. Lancet 2012, 379, 1234–1244. [Google Scholar] [CrossRef] [Green Version]
  6. Ross, J.; Miron, C.E.; Plescia, J.; Laplante, P.; McBride, K.; Moitessier, N.; Möröy, T. Targeting MYC: From Understanding Its Biology to Drug Discovery. Eur. J. Med. Chem. 2021, 213, 113137. [Google Scholar] [CrossRef]
  7. Baudino, T.A.; Cleveland, J.L. The Max Network Gone Mad. Mol. Cell Biol. 2001, 21, 691–702. [Google Scholar] [CrossRef] [Green Version]
  8. Poole, C.J.; van Riggelen, J. MYC—Master Regulator of the Cancer Epigenome and Transcriptome. Genes 2017, 8, 142. [Google Scholar] [CrossRef]
  9. Dang, C.V.; Lee, W.M. Identification of the Human C-Myc Protein Nuclear Translocation Signal. Mol. Cell Biol. 1988, 8, 4048–4054. [Google Scholar] [CrossRef] [Green Version]
  10. Zeller, K.I.; Zhao, X.; Lee, C.W.H.; Chiu, K.P.; Yao, F.; Yustein, J.T.; Ooi, H.S.; Orlov, Y.L.; Shahab, A.; Yong, H.C.; et al. Global Mapping of C-Myc Binding Sites and Target Gene Networks in Human B Cells. Proc. Natl. Acad. Sci. USA 2006, 103, 17834–17839. [Google Scholar] [CrossRef]
  11. Li, Z.; Van Calcar, S.; Qu, C.; Cavenee, W.K.; Zhang, M.Q.; Ren, B. A Global Transcriptional Regulatory Role for C-Myc in Burkitt’s Lymphoma Cells. Proc. Natl. Acad. Sci. USA 2003, 100, 8164–8169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Sabò, A.; Kress, T.R.; Pelizzola, M.; De Pretis, S.; Gorski, M.M.; Tesi, A.; Morelli, M.J.; Bora, P.; Doni, M.; Verrecchia, A.; et al. Selective Transcriptional Regulation by Myc in Cellular Growth Control and Lymphomagenesis. Nature 2014, 511, 488–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Dominguez-Sola, D.; Victora, G.D.; Ying, C.Y.; Phan, R.T.; Saito, M.; Nussenzweig, M.C.; Dalla-Favera, R. The Proto-Oncogene MYC Is Required for Selection in the Germinal Center and Cyclic Reentry. Nat. Immunol. 2012, 13, 1083–1091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Calado, D.P.; Sasaki, Y.; Godinho, S.A.; Pellerin, A.; Köchert, K.; Sleckman, B.P.; De Alborán, I.M.; Janz, M.; Rodig, S.; Rajewsky, K. The Cell-Cycle Regulator c-Myc Is Essential for the Formation and Maintenance of Germinal Centers. Nat. Immunol. 2012, 13, 1092–1100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Toboso-Navasa, A.; Gunawan, A.; Morlino, G.; Nakagawa, R.; Taddei, A.; Damry, D.; Patel, Y.; Chakravarty, P.; Janz, M.; Kassiotis, G.; et al. Restriction of Memory b Cell Differentiation at the Germinal Center b Cell Positive Selection Stage. J. Exp. Med. 2020, 217, e20191933. [Google Scholar] [CrossRef]
  16. Lin, C.Y.; Lovén, J.; Rahl, P.B.; Paranal, R.M.; Burge, C.B.; Bradner, J.E.; Lee, T.I.; Young, R.A. Transcriptional Amplification in Tumor Cells with Elevated C-Myc. Cell 2012, 151, 56–67. [Google Scholar] [CrossRef] [Green Version]
  17. Koh, C.M.; Bezzi, M.; Low, D.H.P.; Ang, W.X.; Teo, S.X.; Gay, F.P.H.; Al-Haddawi, M.; Tan, S.Y.; Osato, M.; Sabò, A.; et al. MYC Regulates the Core Pre-MRNA Splicing Machinery as an Essential Step in Lymphomagenesis. Nature 2015, 523, 96–100. [Google Scholar] [CrossRef]
  18. Iritani, B.M.; Eisenman, R.N. C-Myc Enhances Protein Synthesis and Cell Size during B Lymphocyte Development. Proc. Natl. Acad. Sci. USA 1999, 96, 13180–13185. [Google Scholar] [CrossRef] [Green Version]
  19. Kieffer-Kwon, K.R.; Nimura, K.; Rao, S.S.P.; Xu, J.; Jung, S.; Pekowska, A.; Dose, M.; Stevens, E.; Mathe, E.; Dong, P.; et al. Myc Regulates Chromatin Decompaction and Nuclear Architecture during B Cell Activation. Mol. Cell 2017, 67, 566–578.e10. [Google Scholar] [CrossRef]
  20. Adhikary, S.; Eilers, M. Transcriptional Regulation and Transformation by Myc Proteins. Nat. Rev. Mol. Cell Biol. 2005, 6, 635–645. [Google Scholar] [CrossRef]
  21. Lourenco, C.; Resetca, D.; Redel, C.; Lin, P.; MacDonald, A.S.; Ciaccio, R.; Kenney, T.M.G.; Wei, Y.; Andrews, D.W.; Sunnerhagen, M.; et al. MYC Protein Interactors in Gene Transcription and Cancer. Nat. Rev. Cancer 2021, 21, 579–591. [Google Scholar] [CrossRef] [PubMed]
  22. Rickert, R.C. New Insights into Pre-BCR and BCR Signalling with Relevance to B Cell Malignancies. Nat. Rev. Immunol. 2013, 13, 578–591. [Google Scholar] [CrossRef] [PubMed]
  23. Adams, J.M.; Harris, A.W.; Pinkert, C.A.; Corcoran, L.M.; Alexander, W.S.; Cory, S.; Palmiter, R.D.; Brinster, R.L. The C-Myc Oncogene Driven by Immunoglobulin Enhancers Induces Lymphoid Malignancy in Transgenic Mice. Nature 1985, 318, 533–538. [Google Scholar] [CrossRef] [PubMed]
  24. Lefebure, M.; Tothill, R.W.; Kruse, E.; Hawkins, E.D.; Shortt, J.; Matthews, G.M.; Gregory, G.P.; Martin, B.P.; Kelly, M.J.; Todorovski, I.; et al. Genomic Characterisation of Eμ-Myc Mouse Lymphomas Identifies Bcor as a Myc Co-Operative Tumour-Suppressor Gene. Nat. Commun. 2017, 8, 14581. [Google Scholar] [CrossRef] [Green Version]
  25. Fusello, A.; Horowitz, J.; Yang-Iott, K.; Brady, B.L.; Yin, B.; Rowh, M.A.W.; Rappaport, E.; Bassing, C.H. Histone H2AX Suppresses Translocations in Lymphomas of Eμ-c-Myc Transgenic Mice That Contain a Germline Amplicon of Tumor-Promoting Genes. Cell Cycle 2013, 12, 2867–2875. [Google Scholar] [CrossRef] [Green Version]
  26. Eick, D.; Bornkamm, G.W. Expression of Normal and Translocated C-Myc Alleles in Burkitt’s Lymphoma Cells: Evidence for Different Regulation. EMBO J. 1989, 8, 1965–1972. [Google Scholar] [CrossRef]
  27. Bemark, M.; Neuberger, M.S. The C-MYC Allele That Is Translocated into the IgH Locus Undergoes Constitutive Hypermutation in a Burkitt’s Lymphoma Line. Oncogene 2000, 19, 3404–3410. [Google Scholar] [CrossRef] [Green Version]
  28. Sidman, C.L.; Denial, T.M.; Marshall, J.D.; Roths, J.B. Multiple Mechanisms of Tumorigenesis in Eµ-Myc Transgenic Mice. Cancer Res. 1993, 53, 1665–1669. [Google Scholar]
  29. Joshi, G.; Eberhardt, A.O.; Lange, L.; Winkler, R.; Hoffmann, S.; Kosan, C.; Bierhoff, H. Dichotomous Impact of Myc on Rrna Gene Activation and Silencing in b Cell Lymphomagenesis. Cancers 2020, 12, 3009. [Google Scholar] [CrossRef]
  30. Thürmer, M.; Gollowitzer, A.; Pein, H.; Neukirch, K.; Gelmez, E.; Waltl, L.; Wielsch, N.; Winkler, R.; Löser, K.; Grander, J.; et al. PI(18:1/18:1) Is a SCD1-Derived Lipokine That Limits Stress Signaling. Nat. Commun. 2022, 13, 2982. [Google Scholar] [CrossRef]
  31. Wiese, K.E.; Haikala, H.M.; von Eyss, B.; Wolf, E.; Esnault, C.; Rosenwald, A.; Treisman, R.; Klefström, J.; Eilers, M. Repression of SRF Target Genes Is Critical for Myc-Dependent Apoptosis of Epithelial Cells. EMBO J. 2015, 34, 1554–1571. [Google Scholar] [CrossRef] [PubMed]
  32. Sauvé, S.; Naud, J.F.; Lavigne, P. The Mechanism of Discrimination between Cognate and Non-Specific DNA by Dimeric b/HLH/LZ Transcription Factors. J. Mol. Biol. 2007, 365, 1163–1175. [Google Scholar] [CrossRef] [PubMed]
  33. Ji, H.; Wu, G.; Zhan, X.; Nolan, A.; Koh, C.; de Marzo, A.; Doan, H.M.; Fan, J.; Cheadle, C.; Fallahi, M.; et al. Cell-Type Independent MYC Target Genes Reveal a Primordial Signature Involved in Biomass Accumulation. PLoS ONE 2011, 6, e26057. [Google Scholar] [CrossRef] [PubMed]
  34. Knudson, A.G. Mutation and Cancer: Statistical Study of Retinoblastoma. Proc. Natl. Acad. Sci. USA 1971, 68, 820–823. [Google Scholar] [CrossRef] [Green Version]
  35. Eischen, C.M.; Weber, J.D.; Roussel, M.F.; Sherr, C.J.; Cleveland, J.L. Disruption of the ARF-Mdm2-P53 Tumor Suppressor Pathway in Myc-Induced Lymphomagenesis. Genes Dev. 1999, 13, 2658–2669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Schuster, C.; Berger, A.; Hoelzl, M.A.; Putz, E.M.; Frenzel, A.; Simma, O.; Moritz, N.; Hoelbl, A.; Kovacic, B.; Freissmuth, M.; et al. The Cooperating Mutation or “Second Hit” Determines the Immunologic Visibility toward MYC-Induced Murine Lymphomas. Blood 2011, 118, 4635–4645. [Google Scholar] [CrossRef] [Green Version]
  37. Vallespinós, M.; Fernández, D.; Rodríguez, L.; Alvaro-Blanco, J.; Baena, E.; Ortiz, M.; Dukovska, D.; Martínez, D.; Rojas, A.; Campanero, M.R.; et al. B Lymphocyte Commitment Program Is Driven by the Proto-Oncogene c-Myc. J. Immunol. 2011, 186, 6726–6736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Croxford, J.L.; Li, M.; Tang, F.; Pan, M.F.; Huang, C.W.; Kamran, N.; Meow, C.; Phua, L.; Chng, W.J.; Ng, S.B.; et al. ATM-Dependent Spontaneous Regression of Early Em—Myc—Induced Murine B-Cell Leukemia Depends on Natural Killer and T Cells. Blood 2013, 121, 2512–2521. [Google Scholar] [CrossRef]
  39. Granato, A.; Hayashi, E.A.; Baptista, B.J.A.; Bellio, M.; Nobrega, A. IL-4 Regulates Bim Expression and Promotes B Cell Maturation in Synergy with BAFF Conferring Resistance to Cell Death at Negative Selection Checkpoints. J. Immunol. 2014, 192, 5761–5775. [Google Scholar] [CrossRef] [Green Version]
  40. Ouk, C.; Roland, L.; Gachard, N.; Poulain, S.; Oblet, C.; Rizzo, D.; Saintamand, A.; Lemasson, Q.; Carrion, C.; Thomas, M.; et al. Continuous MYD88 Activation Is Associated with Expansion and Then Transformation of IgM Differentiating Plasma Cells. Front. Immunol. 2021, 12, 641692. [Google Scholar] [CrossRef]
  41. Piskor, E.M.; Winkler, R.; Kosan, C. Analyzing Lymphoma Development and Progression Using HDACi in Mouse Models. Methods Mol. Biol. 2023, 2589, 3–15. [Google Scholar] [CrossRef]
  42. Hunter, J.E.; Butterworth, J.; Perkins, N.D.; Bateson, M.; Richardson, C.A. Using Body Temperature, Food and Water Consumption as Biomarkers of Disease Progression in Mice with Eμ-Myc Lymphoma. Br. J. Cancer 2014, 110, 928–934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Poe, J.C.; Minard-Colin, V.; Kountikov, E.I.; Haas, K.M.; Tedder, T.F. A C-Myc and Surface CD19 Signaling Amplification Loop Promotes B Cell Lymphoma Development and Progression in Mice. J. Immunol. 2012, 189, 2318–2325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Winkler, R.; Mägdefrau, A.S.; Piskor, E.M.; Kleemann, M.; Beyer, M.; Linke, K.; Hansen, L.; Schaffer, A.M.; Hoffmann, M.E.; Poepsel, S.; et al. Targeting the MYC Interaction Network in B-Cell Lymphoma via Histone Deacetylase 6 Inhibition. Oncogene 2022, 41, 4560–4572. [Google Scholar] [CrossRef] [PubMed]
  45. Nepal, R.M.; Tong, L.; Kolaj, B.; Edelmann, W.; Martin, A. Msh2-Dependent DNA Repair Mitigates a Unique Susceptibility of B Cell Progenitors to c-Myc-Induced Lymphomas. Proc. Natl. Acad. Sci. USA 2009, 106, 18698–18703. [Google Scholar] [CrossRef] [Green Version]
  46. Hilmenyuk, T.; Ruckstuhl, C.A.; Hayoz, M.; Berchtold, C.; Nuoffer, J.M.; Solanki, S.; Keun, H.C.; Beavis, P.A.; Riether, C.; Ochsenbein, A.F. T Cell Inhibitory Mechanisms in a Model of Aggressive Non-Hodgkin’s Lymphoma. Oncoimmunology 2018, 7, e1365997. [Google Scholar] [CrossRef] [Green Version]
  47. Davila, M.L.; Kloss, C.C.; Gunset, G.; Sadelain, M. CD19 CAR-Targeted T Cells Induce Long-Term Remission and B Cell Aplasia in an Immunocompetent Mouse Model of B Cell Acute Lymphoblastic Leukemia. PLoS ONE 2013, 8, e61338. [Google Scholar] [CrossRef] [Green Version]
  48. Mattarollo, S.R.; West, A.C.; Steegh, K.; Duret, H.; Paget, C.; Martin, B.; Matthews, G.M.; Shortt, J.; Chesi, M.; Bergsagel, P.L.; et al. NKT Cell Adjuvant-Based Tumor Vaccine for Treatment of Myc Oncogene-Driven Mouse B-Cell Lymphoma. Blood 2012, 120, 3019–3029. [Google Scholar] [CrossRef] [Green Version]
  49. Kobayashi, T.; Doff, B.L.; Rearden, R.C.; Leggatt, G.R.; Mattarollo, S.R. NKT Cell-Targeted Vaccination plus Anti-4–1BB Antibody Generates Persistent CD8 T Cell Immunity against B Cell Lymphoma. Oncoimmunology 2015, 4, e990793. [Google Scholar] [CrossRef] [Green Version]
  50. Ross, J.; Rashkovan, M.; Fraszczak, J.; Joly-Beauparlant, C.; Vadnais, C.; Winkler, R.; Droit, A.; Kosan, C.; Moroy, T. Deletion of the MIZ-1 POZ Domain Increases Efficacy of Cytarabine Treatment in T- And B-ALL/Lymphoma Mouse Models. Cancer Res. 2019, 79, 4184–4195. [Google Scholar] [CrossRef]
  51. Faubert, B.; Boily, G.; Izreig, S.; Griss, T.; Samborska, B.; Dong, Z.; Dupuy, F.; Chambers, C.; Fuerth, B.J.; Viollet, B.; et al. AMPK Is a Negative Regulator of the Warburg Effect and Suppresses Tumor Growth in Vivo. Cell. Metab. 2013, 17, 113–124. [Google Scholar] [CrossRef] [PubMed]
  52. Wolyniec, K.; Shortt, J.; de Stanchina, E.; Levav-Cohen, Y.; Alsheich-Bartok, O.; Louria-Hayon, I.; Corneille, V.; Kumar, B.; Woods, S.J.; Opat, S.; et al. E6AP Ubiquitin Ligase Regulates PML-Induced Senescence in Myc-Driven Lymphomagenesis. Blood 2012, 120, 822–832. [Google Scholar] [CrossRef] [PubMed]
  53. Mori, S.; Rempel, R.E.; Chang, J.T.; Yao, G.; Lagoo, A.S.; Potti, A.; Bild, A.; Nevins, J.R. Utilization of Pathway Signatures to Reveal Distinct Types of B Lymphoma in the Eμ-Myc Model and Human Diffuse Large B-Cell Lymphoma. Cancer Res. 2008, 68, 8525–8534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Chapuy, B.; Stewart, C.; Dunford, A.J.; Kim, J.; Kamburov, A.; Redd, R.A.; Lawrence, M.S.; Roemer, M.G.M.; Li, A.J.; Ziepert, M.; et al. Molecular Subtypes of Diffuse Large B Cell Lymphoma Are Associated with Distinct Pathogenic Mechanisms and Outcomes. Nat. Med. 2018, 24, 679–690. [Google Scholar] [CrossRef]
  55. Wall, M.; Poortinga, G.; Stanley, K.L.; Lindemann, R.K.; Bots, M.; Chan, C.J.; Bywater, M.J.; Kinross, K.M.; Astle, M.V.; Waldeck, K.; et al. The MTORC1 Inhibitor Everolimus Prevents and Treats Eμ-Myc Lymphoma by Restoring Oncogene-Induced Senescence. Cancer Discov. 2013, 3, 82–95. [Google Scholar] [CrossRef] [Green Version]
  56. Witzig, T.E.; Reeder, C.B.; Laplant, B.R.; Gupta, M.; Johnston, P.B.; Micallef, I.N.; Porrata, L.F.; Ansell, S.M.; Colgan, J.P.; Jacobsen, E.D.; et al. A Phase II Trial of the Oral MTOR Inhibitor Everolimus in Relapsed Aggressive Lymphoma. Leukemia 2011, 25, 341–347. [Google Scholar] [CrossRef]
  57. Hartleben, G.; Müller, C.; Krämer, A.; Schimmel, H.; Zidek, L.M.; Dornblut, C.; Winkler, R.; Eichwald, S.; Kortman, G.; Kosan, C.; et al. Tuberous Sclerosis Complex Is Required for Tumor Maintenance in MYC-driven Burkitt’s Lymphoma. EMBO J. 2018, 37, e98589. [Google Scholar] [CrossRef]
  58. Xu, W.; Berning, P.; Erdmann, T.; Grau, M.; Bettazová, N.; Zapukhlyak, M.; Frontzek, F.; Kosnopfel, C.; Lenz, P.; Grondine, M.; et al. MTOR Inhibition Amplifies the Anti-Lymphoma Effect of PI3Kβ/δ Blockage in Diffuse Large B-Cell Lymphoma. Leukemia 2022. [Google Scholar] [CrossRef]
  59. Schleich, K.; Kase, J.; Dörr, J.R.; Trescher, S.; Bhattacharya, A.; Yu, Y.; Wailes, E.M.; Fan, D.N.Y.; Lohneis, P.; Milanovic, M.; et al. H3K9me3-Mediated Epigenetic Regulation of Senescence in Mice Predicts Outcome of Lymphoma Patients. Nat. Commun. 2020, 11, 3651. [Google Scholar] [CrossRef]
  60. Rickert, R.C.; Roes, J.; Rajewsky, K. B Lymphocyte-Specific, Cre-Mediated Mutagenesis in Mice. Nucleic Acids Res. 1997, 25, 1317–1318. [Google Scholar] [CrossRef]
  61. Hobeika, E.; Thiemann, S.; Storch, B.; Jumaa, H.; Nielsen, P.J.; Pelanda, R.; Reth, M. Testing Gene Function Early in the B Cell Lineage in Mb1-Cre Mice. Proc. Natl. Acad. Sci. USA 2006, 103, 13789–13794. [Google Scholar] [CrossRef] [PubMed]
  62. Lacroix, M.; Beauchemin, H.; Fraszczak, J.; Ross, J.; Shooshtarizadeh, P.; Chen, R.; Möröy, T. The X-Linked Helicase DDX3X Is Required for Lymphoid Differentiation and MYC-Driven Lymphomagenesis. Cancer Res. 2022, 82, 3172–3186. [Google Scholar] [CrossRef] [PubMed]
  63. Crouch, E.E.; Li, Z.; Takizawa, M.; Fichtner-Feigl, S.; Gourzi, P.; Montaño, C.; Feigenbaum, L.; Wilson, P.; Janz, S.; Papavasiliou, F.N.; et al. Regulation of AID Expression in the Immune Response. J. Exp. Med. 2007, 204, 1145–1156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Casola, S.; Cattoretti, G.; Uyttersprot, N.; Koralov, S.B.; Segal, J.; Hao, Z.; Waisman, A.; Egert, A.; Ghitza, D.; Rajewsky, K. Tracking Germinal Center B Cells Expressing Germ-Line Immunoglobulin Γ1 Transcripts by Conditional Gene Targeting. Proc. Natl. Acad. Sci. USA 2006, 103, 7396–7401. [Google Scholar] [CrossRef] [Green Version]
  65. Kraus, M.; Alimzhanov, M.B.; Rajewsky, N.; Rajewsky, K. Survival of Resting Mature B Lymphocytes Depends on BCR Signaling via the Igα/β Heterodimer. Cell 2004, 117, 787–800. [Google Scholar] [CrossRef] [Green Version]
  66. Chen, E.Y.; Tan, C.M.; Kou, Y.; Duan, Q.; Wang, Z.; Meirelles, G.V.; Clark, N.R.; Ma’ayan, A. Enrichr: Interactive and Collaborative HTML5 Gene List Enrichment Analysis Tool. BMC Bioinform. 2013, 14, 128. [Google Scholar] [CrossRef] [Green Version]
  67. Huang, R.; Grishagin, I.; Wang, Y.; Zhao, T.; Greene, J.; Obenauer, J.C.; Ngan, D.; Nguyen, D.T.; Guha, R.; Jadhav, A.; et al. The NCATS BioPlanet—An Integrated Platform for Exploring the Universe of Cellular Signaling Pathways for Toxicology, Systems Biology, and Chemical Genomics. Front. Pharm. 2019, 10, 445. [Google Scholar] [CrossRef] [Green Version]
  68. Supek, F.; Bošnjak, M.; Škunca, N.; Šmuc, T. Revigo Summarizes and Visualizes Long Lists of Gene Ontology Terms. PLoS ONE 2011, 6, e21800. [Google Scholar] [CrossRef] [Green Version]
  69. Alt, J.R.; Greiner, T.C.; Cleveland, J.L.; Eischen, C.M. Mdm2 Haplo-Insufficiency Profoundly Inhibits Myc-Induced Lymphomagenesis. EMBO J. 2003, 22, 1442–1450. [Google Scholar] [CrossRef] [Green Version]
  70. Post, S.M.; Quintás-Cardama, A.; Terzian, T.; Smith, C.; Eischen, C.M.; Lozano, G. P53-Dependent Senescence Delays E-Myc-Induced B-Cell Lymphomagenesis. Oncogene 2010, 29, 1260–1269. [Google Scholar] [CrossRef] [Green Version]
  71. Schmitt, C.A.; McCurrach, M.E.; de Stanchina, E.; Wallace-Brodeur, R.R.; Lowe, S.W. INK4a/ARF Mutations Accelerate Lymphomagenesis and Promote Chemoresistance by Disabling P53. Genes Dev. 1999, 13, 2670–2677. [Google Scholar] [CrossRef] [PubMed]
  72. Yetil, A.; Anchang, B.; Gouw, A.M.; Adam, S.J.; Zabuawala, T.; Parameswaran, R.; van Riggelen, J.; Plevritis, S.; Felsher, D.W. P19ARF Is a Critical Mediator of Both Cellular Senescence and an Innate Immune Response Associated with MYC Inactivation in Mouse Model of Acute Leukemia. Oncotarget 2015, 6, 3563–3577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Huang, Q.; Chen, L.; Yang, L.; Xie, X.; Gan, L.; Cleveland, J.L.; Chen, J. MDMX Acidic Domain Inhibits P53 DNA Binding in Vivo and Regulates Tumorigenesis. Proc. Natl. Acad. Sci. USA 2018, 115, E3368–E3377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Meng, X.; Carlson, N.R.; Dong, J.; Zhang, Y. Oncogenic C-Myc-Induced Lymphomagenesis Is Inhibited Non-Redundantly by the P19Arf-Mdm2-P53 and RP-Mdm2-P53 Pathways. Oncogene 2015, 34, 5709–5717. [Google Scholar] [CrossRef] [Green Version]
  75. Donehower, L.A.; Harvey, M.; Slagle, B.L.; McArthur, M.J.; Montgomery, C.A.; Butel, J.S.; Bradley, A. Mice Deficient for P53 Are Developmentally Normal but Susceptible to Spontaneous Tumours. Nature 1992, 356, 215–221. [Google Scholar] [CrossRef]
  76. Rowh, M.A.W.; Demicco, A.; Horowitz, J.E.; Yin, B.; Yang-Iott, K.S.; Fusello, A.M.; Hobeika, E.; Reth, M.; Bassing, C.H. Tp53 Deletion in B Lineage Cells Predisposes Mice to Lymphomas with Oncogenic Translocations. Oncogene 2011, 30, 4757–4764. [Google Scholar] [CrossRef] [Green Version]
  77. Gostissa, M.; Bianco, J.M.; Malkin, D.J.; Kutok, J.L.; Rodig, S.J.; Morse, H.C.; Bassing, C.H.; Alt, F.W. Conditional Inactivation of P53 in Mature B Cells Promotes Generation of Nongerminal Center-Derived B-Cell Lymphomas. Proc. Natl. Acad. Sci. USA 2013, 110, 2934–2939. [Google Scholar] [CrossRef] [Green Version]
  78. Rempel, R.E.; Mori, S.; Gasparetto, M.; Glozak, M.A.; Andrechek, E.R.; Adler, S.B.; Laakso, N.M.; Lagoo, A.S.; Storms, R.; Smith, C.; et al. A Role for E2F Activities in Determining the Fate of Myc-Induced Lymphomagenesis. PLoS Genet. 2009, 5, e1000640. [Google Scholar] [CrossRef] [Green Version]
  79. Kent, L.N.; Leone, G. The Broken Cycle: E2F Dysfunction in Cancer. Nat. Rev. Cancer 2019, 19, 326–338. [Google Scholar] [CrossRef]
  80. Bouchard, C.; Dittrich, O.; Kiermaier, A.; Dohmann, K.; Menkel, A.; Eilers, M.; Lüscher, B. Regulation of Cyclin D2 Gene Expression by the Myc/Max/Mad Network: Myc-Dependent TRRAP Recruitment and Histone Acetylation at the Cyclin D2 Promoter. Genes Dev. 2001, 15, 2042–2047. [Google Scholar] [CrossRef] [Green Version]
  81. Wang, C.; Lisanti, M.P.; Liao, D.J. Reviewing Once More the C-Myc and Ras Collaboration: Converging at the Cyclin D1-CDK4 Complex and Challenging Basic Concepts of Cancer Biology. Cell Cycle 2011, 10, 57–67. [Google Scholar] [CrossRef] [PubMed]
  82. Lai, M.-C.; Chang, W.-C.; Shieh, S.-Y.; Tarn, W.-Y. DDX3 Regulates Cell Growth through Translational Control of Cyclin E1. Mol. Cell Biol. 2010, 30, 5444–5453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Farria, A.T.; Plummer, J.B.; Salinger, A.P.; Shen, J.; Lin, K.; Lu, Y.; McBride, K.M.; Koutelou, E.; Dent, S.Y.R. Transcriptional Activation of MYC-Induced Genes by GCN5 Promotes B-Cell Lymphomagenesis. Cancer Res. 2020, 80, 5543–5553. [Google Scholar] [CrossRef] [PubMed]
  84. Sheikh, B.N.; Lee, S.C.W.; El-Saafin, F.; Vanyai, H.K.; Hu, Y.; Pang, S.H.M.; Grabow, S.; Strasser, A.; Nutt, S.L.; Alexander, W.S.; et al. MOZ Regulates B-Cell Progenitors and, Consequently, Moz Haploinsufficiency Dramatically Retards MYC-Induced Lymphoma Development. Blood 2015, 125, 1910–1921. [Google Scholar] [CrossRef] [PubMed]
  85. Baell, J.B.; Leaver, D.J.; Hermans, S.J.; Kelly, G.L.; Brennan, M.S.; Downer, N.L.; Nguyen, N.; Wichmann, J.; McRae, H.M.; Yang, Y.; et al. Inhibitors of Histone Acetyltransferases KAT6A/B Induce Senescence and Arrest Tumour Growth. Nature 2018, 560, 253–257. [Google Scholar] [CrossRef] [PubMed]
  86. Berg, T.; Thoene, S.; Yap, D.; Wee, T.; Schoeler, N.; Rosten, P.; Lim, E.; Bilenky, M.; Mungall, A.J.; Oellerich, T.; et al. A Transgenic Mouse Model Demonstrating the Oncogenic Role of Mutations in the Polycomb-Group Gene EZH2 in Lymphomagenesis. Blood 2014, 123, 3914–3924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Schmitz, R.; Wright, G.W.; Huang, D.W.; Johnson, C.A.; Phelan, J.D.; Wang, J.Q.; Roulland, S.; Kasbekar, M.; Young, R.M.; Shaffer, A.L.; et al. Genetics and Pathogenesis of Diffuse Large B-Cell Lymphoma. N. Engl. J. Med. 2018, 378, 1396–1407. [Google Scholar] [CrossRef] [PubMed]
  88. Vandel, L.; Nicolas, E.; Vaute, O.; Ferreira, R.; Ait-Si-Ali, S.; Trouche, D. Transcriptional Repression by the Retinoblastoma Protein through the Recruitment of a Histone Methyltransferase. Mol. Cell Biol. 2001, 21, 6484–6494. [Google Scholar] [CrossRef] [Green Version]
  89. Nielsen, S.J.; Schneider, R.; Bauer, U.M.; Bannister, A.J.; Morrison, A.; O’Carroll, D.; Firestein, R.; Cleary, M.; Jenuwein, T.; Herrera, R.E.; et al. Rb Targets Histone H3 Methylation and HP1 to Promoters. Nature 2001, 412, 561–565. [Google Scholar] [CrossRef]
  90. Reimann, M.; Lee, S.; Loddenkemper, C.; Dörr, J.R.; Tabor, V.; Aichele, P.; Stein, H.; Dörken, B.; Jenuwein, T.; Schmitt, C.A. Tumor Stroma-Derived TGF-β Limits Myc-Driven Lymphomagenesis via Suv39h1-Dependent Senescence. Cancer Cell 2010, 17, 262–272. [Google Scholar] [CrossRef] [Green Version]
  91. Fernández-serrano, M.; Winkler, R.; Santos, J.C.; le Pannérer, M.M.; Buschbeck, M.; Roué, G. Histone Modifications and Their Targeting in Lymphoid Malignancies. Int. J. Mol. Sci. 2022, 23, 253. [Google Scholar] [CrossRef] [PubMed]
  92. Mathsyaraja, H.; Freie, B.; Cheng, P.F.; Babaeva, E.; Catchpole, J.T.; Janssens, D.; Henikoff, S.; Eisenman, R.N. Max Deletion Destabilizes MYC Protein and Abrogates Eµ-Myc Lymphomagenesis. Genes Dev. 2019, 33, 1252–1264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Prochownik, E.V.; Vogt, P.K. Therapeutic Targeting of Myc. Genes Cancer 2010, 1, 650–659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Holmes, A.G.; Parker, J.B.; Sagar, V.; Truica, M.I.; Soni, P.N.; Han, H.; Schiltz, G.E.; Abdulkadir, S.A.; Chakravarti, D. A MYC Inhibitor Selectively Alters the MYC and MAX Cistromes and Modulates the Epigenomic Landscape to Regulate Target Gene Expression. Sci. Adv. 2022, 8, eabh3635. [Google Scholar] [CrossRef] [PubMed]
  95. Castell, A.; Yan, Q.; Fawkner, K.; Bazzar, W.; Zhang, F.; Wickström, M.; Alzrigat, M.; Franco, M.; Krona, C.; Cameron, D.P.; et al. MYCMI-7: A Small MYC-Binding Compound That Inhibits MYC: MAX Interaction and Tumor Growth in a MYC-Dependent Manner. Cancer Res. Commun. 2022, 2, 182–201. [Google Scholar] [CrossRef]
  96. Salghetti, S.E.; Kim, S.Y.; Tansey, W.P. Destruction of Myc by Ubiquitin-Mediated Proteolysis: Cancer-Associated and Transforming Mutations Stabilize Myc. EMBO J. 1999, 18, 717–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Sears, R.; Leone, G.; DeGregori, J.; Nevins, J.R. Ras Enhances Myc Protein Stability. Mol. Cell 1999, 3, 169–179. [Google Scholar] [CrossRef] [PubMed]
  98. Gregory, M.A.; Hann, S.R. C-Myc Proteolysis by the Ubiquitin-Proteasome Pathway: Stabilization of c-Myc in Burkitt’s Lymphoma Cells. Mol Cell Biol. 2000, 20, 2423–2435. [Google Scholar] [CrossRef] [Green Version]
  99. Sears, R.; Nuckolls, F.; Haura, E.; Taya, Y.; Tamai, K.; Nevins, J.R. Multiple Ras-Dependent Phosphorylation Pathways Regulate Myc Protein Stability. Genes Dev. 2000, 14, 2501–2514. [Google Scholar] [CrossRef] [Green Version]
  100. Cohn, G.M.; Liefwalker, D.F.; Langer, E.M.; Sears, R.C. PIN1 Provides Dynamic Control of MYC in Response to Extrinsic Signals. Front. Cell Dev. Biol. 2020, 8, 224. [Google Scholar] [CrossRef]
  101. D’Artista, L.; Bisso, A.; Piontini, A.; Doni, M.; Verrecchia, A.; Kress, T.R.; Morelli, M.J.; del Sal, G.; Amati, B.; Campaner, S. Pin1 Is Required for Sustained B Cell Proliferation upon Oncogenic Activation of Myc. Oncotarget 2016, 7, 21786–21798. [Google Scholar] [CrossRef] [PubMed]
  102. Szklarczyk, D.; Gable, A.L.; Nastou, K.C.; Lyon, D.; Kirsch, R.; Pyysalo, S.; Doncheva, N.T.; Legeay, M.; Fang, T.; Bork, P.; et al. The STRING Database in 2021: Customizable Protein-Protein Networks, and Functional Characterization of User-Uploaded Gene/Measurement Sets. Nucleic Acids Res. 2021, 49, D605–D612. [Google Scholar] [CrossRef] [PubMed]
  103. Alexandrova, N.; Niklinski, J.; Bliskovsky, V.; Otterson, G.A.; Blake, M.; Kaye, F.J.; Zajac-Kaye, M. The N-Terminal Domain of c-Myc Associates with Alpha-Tubulin and Microtubules in Vivo and in Vitro. Mol. Cell Biol. 1995, 15, 5188–5195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Matsuyama, A.; Shimazu, T.; Sumida, Y.; Saito, A.; Yoshimatsu, Y.; Seigneurin-Berny, D.; Osada, H.; Komatsu, Y.; Nishino, N.; Khochbin, S.; et al. In Vivo Destabilization of Dynamic Microtubules by HDAC6-Mediated Deacetylation. EMBO J. 2002, 21, 6820–6831. [Google Scholar] [CrossRef] [Green Version]
  105. Miyake, Y.; Keusch, J.J.; Wang, L.; Saito, M.; Hess, D.; Wang, X.; Melancon, B.J.; Helquist, P.; Gut, H.; Matthias, P. Structural Insights into HDAC6 Tubulin Deacetylation and Its Selective Inhibition. Nat. Chem. Biol. 2016, 12, 748–754. [Google Scholar] [CrossRef]
  106. Adams, C.M.; Kim, A.S.; Mitra, R.; Choi, J.K.; Gong, J.Z.; Eischen, C.M. BCL-W Has a Fundamental Role in B Cell Survival and Lymphomagenesis. J. Clin. Investig. 2017, 127, 635–650. [Google Scholar] [CrossRef] [Green Version]
  107. Kelly, G.L.; Grabow, S.; Glaser, S.P.; Fitzsimmons, L.; Aubrey, B.J.; Okamoto, T.; Valente, L.J.; Robati, M.; Tai, L.; Douglas Fairlie, W.; et al. Targeting of MCL-1 Kills MYC-Driven Mouse and Human Lymphomas Even When They Bear Mutations in P53. Genes Dev. 2014, 28, 58–70. [Google Scholar] [CrossRef] [Green Version]
  108. Grabow, S.; Kelly, G.L.; Delbridge, A.R.D.; Kelly, P.N.; Bouillet, P.; Adams, J.M.; Strasser, A. Critical B-Lymphoid Cell Intrinsic Role of Endogenous MCL-1 in c-MYC-Induced Lymphomagenesis. Cell Death Dis. 2016, 7, e2132. [Google Scholar] [CrossRef]
  109. Egle, A.; Harris, A.W.; Bouillet, P.; Cory, S. Bim Is a Suppressor of Myc-Induced Mouse B Cell Leukemia. Proc. Natl. Acad. Sci. USA 2004, 101, 6164–6169. [Google Scholar] [CrossRef] [Green Version]
  110. Liu, R.; King, A.; Bouillet, P.; Tarlinton, D.M.; Strasser, A.; Heierhorst, J. Proapoptotic BIM Impacts B Lymphoid Homeostasis by Limiting the Survival of Mature B Cells in a Cell-Autonomous Manner. Front. Immunol. 2018, 8, 592. [Google Scholar] [CrossRef]
  111. Michalak, E.M.; Jansen, E.S.; Happo, L.; Cragg, M.S.; Tai, L.; Smyth, G.K.; Strasser, A.; Adams, J.M.; Scott, C.L. Puma and to a Lesser Extent Noxa Are Suppressors of Myc-Induced Lymphomagenesis. Cell Death Differ. 2009, 16, 684–696. [Google Scholar] [CrossRef] [PubMed]
  112. Patel, J.H.; McMahon, S.B. Targeting of Miz-1 Is Essential for Myc-Mediated Apoptosis. J. Biol. Chem. 2006, 281, 3283–3289. [Google Scholar] [CrossRef] [PubMed]
  113. Patel, J.H.; McMahon, S.B. BCL2 Is a Downstream Effector of MIZ-1 Essential for Blocking c-MYC-Induced Apoptosis. J. Biol. Chem. 2007, 282, 5–13. [Google Scholar] [CrossRef] [Green Version]
  114. Baluapuri, A.; Wolf, E.; Eilers, M. Target Gene-Independent Functions of MYC Oncoproteins. Nat. Rev. Mol. Cell Biol. 2020, 21, 255–267. [Google Scholar] [CrossRef] [PubMed]
  115. Murga, M.; Campaner, S.; Lopez-Contreras, A.J.; Toledo, L.I.; Soria, R.; Montaña, M.F.; D’Artista, L.; Schleker, T.; Guerra, C.; Garcia, E.; et al. Exploiting Oncogene-Induced Replicative Stress for the Selective Killing of Myc-Driven Tumors. Nat. Struct. Mol. Biol. 2011, 18, 1331–1335. [Google Scholar] [CrossRef] [PubMed]
  116. Puccetti, M.V.; Adams, C.M.; Kushinsky, S.; Eischen, C.M. SMARCAL1 and ZrAnB3 Protect Replication Forks from MYC-Induced DNA Replication Stress. Cancer Res. 2019, 79, 1612–1623. [Google Scholar] [CrossRef]
  117. Hamperl, S.; Cimprich, K.A. Conflict Resolution in the Genome: How Transcription and Replication Make It Work. Cell 2016, 167, 1455–1467. [Google Scholar] [CrossRef] [Green Version]
  118. Sollier, J.; Cimprich, K.A. Breaking Bad: R-Loops and Genome Integrity. Trends Cell Biol. 2015, 25, 514–522. [Google Scholar] [CrossRef] [Green Version]
  119. Hamperl, S.; Bocek, M.J.; Saldivar, J.C.; Swigut, T.; Cimprich, K.A. Transcription-Replication Conflict Orientation Modulates R-Loop Levels and Activates Distinct DNA Damage Responses. Cell 2017, 170, 774–786.e19. [Google Scholar] [CrossRef] [Green Version]
  120. Schuler, F.; Weiss, J.G.; Lindner, S.E.; Lohmüller, M.; Herzog, S.; Spiegl, S.F.; Menke, P.; Geley, S.; Labi, V.; Villunger, A. Checkpoint Kinase 1 Is Essential for Normal B Cell Development and Lymphomagenesis. Nat. Commun. 2017, 8, 1697. [Google Scholar] [CrossRef] [Green Version]
  121. Gorrini, C.; Squatrito, M.; Luise, C.; Syed, N.; Perna, D.; Wark, L.; Martinato, F.; Sardella, D.; Verrecchia, A.; Bennett, S.; et al. Tip60 Is a Haplo-Insufficient Tumour Suppressor Required for an Oncogene-Induced DNA Damage Response. Nature 2007, 448, 1063–1067. [Google Scholar] [CrossRef] [PubMed]
  122. Patel, J.H.; Du, Y.; Ard, P.G.; Phillips, C.; Carella, B.; Chen, C.-J.; Rakowski, C.; Chatterjee, C.; Lieberman, P.M.; Lane, W.S.; et al. The C-MYC Oncoprotein Is a Substrate of the Acetyltransferases HGCN5/PCAF and TIP60. Mol. Cell Biol. 2004, 24, 10826–10834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Wang, Q.; Zhang, H.; Kajino, K.; Greene, M.I. BRCA1 Binds C-Myc and Inhibits Its Transcriptional and Transforming Activity in Cells. Oncogene 1998, 17, 1939–1948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Herold, S.; Kalb, J.; Büchel, G.; Ade, C.P.; Baluapuri, A.; Xu, J.; Koster, J.; Solvie, D.; Carstensen, A.; Klotz, C.; et al. Recruitment of BRCA1 Limits MYCN-Driven Accumulation of Stalled RNA Polymerase. Nature 2019, 567, 545–549. [Google Scholar] [CrossRef]
  125. Choi, P.S.; Van Riggelen, J.; Gentles, A.J.; Bachireddy, P.; Rakhra, K.; Adam, S.J.; Plevritis, S.K.; Felsher, D.W. Lymphomas That Recur after MYC Suppression Continue to Exhibit Oncogene Addiction. Proc. Natl. Acad. Sci. USA 2011, 108, 17432–17437. [Google Scholar] [CrossRef] [Green Version]
  126. Van Riggelen, J.; Müller, J.; Otto, T.; Beuger, V.; Yetil, A.; Choi, P.S.; Kosan, C.; Möröy, T.; Felsher, D.W.; Eilers, M. The Interaction between Myc and Miz1 Is Required to Antagonize TGFβ-Dependent Autocrine Signaling during Lymphoma Formation and Maintenance. Genes Dev. 2010, 24, 1281–1294. [Google Scholar] [CrossRef] [Green Version]
  127. Muthalagu, N.; Monteverde, T.; Raffo-Iraolagoitia, X.; Wiesheu, R.; Whyte, D.; Hedley, A.; Laing, S.; Kruspig, B.; Upstill-Goddard, R.; Shaw, R.; et al. Repression of the Type i Interferon Pathway Underlies MYC-and KRAS-Dependent Evasion of NK and B Cells in Pancreatic Ductal Adenocarcinoma. Cancer Discov. 2020, 10, 872–887. [Google Scholar] [CrossRef] [Green Version]
  128. Bartha, Á.; Győrffy, B. TNMplot.Com: A Web Tool for the Comparison of Gene Expression in Normal, Tumor and Metastatic Tissues. Int. J. Mol. Sci. 2021, 22, 2622. [Google Scholar] [CrossRef]
  129. Gao, J.; Aksoy, B.A.; Dogrusoz, U.; Dresdner, G.; Gross, B.; Sumer, S.O.; Sun, Y.; Jacobsen, A.; Sinha, R.; Larsson, E.; et al. Integrative Analysis of Complex Cancer Genomics and Clinical Profiles Using the CBioPortal. Sci. Signal. 2013, 6. [Google Scholar] [CrossRef] [Green Version]
  130. Cerami, E.; Gao, J.; Dogrusoz, U.; Gross, B.E.; Sumer, S.O.; Aksoy, B.A.; Jacobsen, A.; Byrne, C.J.; Heuer, M.L.; Larsson, E.; et al. The CBio Cancer Genomics Portal: An Open Platform for Exploring Multidimensional Cancer Genomics Data. Cancer Discov. 2012, 2, 401–404. [Google Scholar] [CrossRef] [Green Version]
  131. Rivas, M.A.; Melnick, A.M. Role of Chromosomal Architecture in Germinal Center B Cells and Lymphomagenesis. Curr. Opin. Hematol. 2019, 26, 294–302. [Google Scholar] [CrossRef] [PubMed]
  132. Canela, A.; Maman, Y.; Jung, S.; Wong, N.; Callen, E.; Day, A.; Kieffer-Kwon, K.R.; Pekowska, A.; Zhang, H.; Rao, S.S.P.; et al. Genome Organization Drives Chromosome Fragility. Cell 2017, 170, 507–521.e18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Hnisz, D.; Weintrau, A.S.; Day, D.S.; Valton, A.L.; Bak, R.O.; Li, C.H.; Goldmann, J.; Lajoie, B.R.; Fan, Z.P.; Sigova, A.A.; et al. Activation of Proto-Oncogenes by Disruption of Chromosome Neighborhoods. Science (1979) 2016, 351, 1454–1458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Yang, Y.; McBride, K.M.; Hensley, S.; Lu, Y.; Chedin, F.; Bedford, M.T. Arginine Methylation Facilitates the Recruitment of TOP3B to Chromatin to Prevent R Loop Accumulation. Mol. Cell 2014, 53, 484–497. [Google Scholar] [CrossRef] [Green Version]
  135. Hall, Z.; Ament, Z.; Wilson, C.H.; Burkhart, D.L.; Ashmore, T.; Koulman, A.; Littlewood, T.; Evan, G.I.; Griffin, J.L. MYC Expression Drives Aberrant Lipid Metabolism in Lung Cancer. Cancer Res. 2016, 76, 4608–4618. [Google Scholar] [CrossRef] [Green Version]
  136. Eberlin, L.S.; Gabay, M.; Fan, A.C.; Gouw, A.M.; Tibshirani, R.J.; Felsher, D.W.; Zare, R.N. Alteration of the Lipid Profile in Lymphomas Induced by MYC Overexpression. Proc. Natl. Acad. Sci. USA 2014, 111, 10450–10455. [Google Scholar] [CrossRef] [Green Version]
  137. Chu, B.; Kon, N.; Chen, D.; Li, T.; Liu, T.; Jiang, L.; Song, S.; Tavana, O.; Gu, W. ALOX12 Is Required for P53-Mediated Tumour Suppression through a Distinct Ferroptosis Pathway. Nat. Cell Biol. 2019, 21, 579–591. [Google Scholar] [CrossRef]
  138. Hofmann, J.W.; Zhao, X.; De Cecco, M.; Peterson, A.L.; Pagliaroli, L.; Manivannan, J.; Hubbard, G.B.; Ikeno, Y.; Zhang, Y.; Feng, B.; et al. Reduced Expression of MYC Increases Longevity and Enhances Healthspan. Cell 2015, 160, 477–488. [Google Scholar] [CrossRef] [Green Version]
  139. Pourdehnad, M.; Truitt, M.L.; Siddiqi, I.N.; Ducker, G.S.; Shokat, K.M.; Ruggero, D. Myc and MTOR Converge on a Common Node in Protein Synthesis Control That Confers Synthetic Lethality in Myc-Driven Cancers. Proc. Natl. Acad. Sci. USA 2013, 110, 11988–11993. [Google Scholar] [CrossRef] [Green Version]
  140. Kotani, A.; Kakazu, N.; Tsuruyama, T.; Okazaki, I.M.; Muramatsu, M.; Kinoshita, K.; Nagaoka, H.; Yabe, D.; Honjo, T. Activation-Induced Cytidine Deaminase (AID) Promotes B Cell Lymphomagenesis in Emu-Cmyc Transgenic Mice. Proc. Natl. Acad. Sci. USA 2007, 104, 1616–1620. [Google Scholar] [CrossRef] [Green Version]
  141. Nepal, R.M.; Zaheen, A.; Basit, W.; Li, L.; Berger, S.A.; Martin, A. AID and RAG1 Do Not Contribute to Lymphomagenesis in Eμ C-Myc Transgenic Mice. Oncogene 2008, 27, 4752–4756. [Google Scholar] [CrossRef] [PubMed]
  142. Scuoppo, C.; Miething, C.; Lindqvist, L.; Reyes, J.; Ruse, C.; Appelmann, I.; Yoon, S.; Krasnitz, A.; Teruya-Feldstein, J.; Pappin, D.; et al. A Tumour Suppressor Network Relying on the Polyamine-Hypusine Axis. Nature 2012, 487, 244–248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Scott, C.L.; Schuler, M.; Marsden, V.S.; Egle, A.; Pellegrini, M.; Nesic, D.; Gerondakis, S.; Nutt, S.L.; Green, D.R.; Strasser, A. Apaf-1 and Caspase-9 Do Not Act as Tumor Suppressors in Myc-Induced Lymphomagenesis or Mouse Embryo Fibroblast Transformation. J. Cell Biol. 2004, 164, 89–96. [Google Scholar] [CrossRef] [PubMed]
  144. Walczynski, J.; Lyons, S.; Jones, N.; Breitwieser, W. Sensitisation of C-MYC-Induced B-Lymphoma Cells to Apoptosis by ATF2. Oncogene 2014, 33, 1027–1036. [Google Scholar] [CrossRef]
  145. Tameire, F.; Verginadis, I.I.; Leli, N.M.; Polte, C.; Conn, C.S.; Ojha, R.; Salas Salinas, C.; Chinga, F.; Monroy, A.M.; Fu, W.; et al. ATF4 Couples MYC-Dependent Translational Activity to Bioenergetic Demands during Tumour Progression. Nat. Cell Biol. 2019, 21, 889–899. [Google Scholar] [CrossRef]
  146. Eischen, C.M.; Roussel, M.F.; Korsmeyer, S.J.; Cleveland, J.L. Bax Loss Impairs Myc-Induced Apoptosis and Circumvents the Selection of P53 Mutations during Myc-Mediated Lymphomagenesis. Mol. Cell Biol. 2001, 21, 7653–7662. [Google Scholar] [CrossRef] [Green Version]
  147. Kelly, P.N.; Puthalakath, H.; Adams, J.M.; Strasser, A. Endogenous Bcl-2 Is Not Required for the Development of Eμ-Myc-Induced B-Cell Lymphoma. Blood 2007, 109, 4907–4913. [Google Scholar] [CrossRef] [Green Version]
  148. Kelly, P.N.; Grabow, S.; Delbridge, A.R.D.; Strasser, A.; Adams, J.M. Endogenous Bcl-XL Is Essential for Myc-Driven Lymphomagenesis in Mice. Blood 2011, 118, 6380–6386. [Google Scholar] [CrossRef] [Green Version]
  149. Takahashi, Y.; Hori, T.; Cooper, T.K.; Liao, J.; Desai, N.; Serfass, J.M.; Young, M.M.; Park, S.; Izu, Y.; Wang, H.G. Bif-1 Haploinsufficiency Promotes Chromosomal Instability and Accelerates Myc-Driven Lymphomagenesis via Suppression of Mitophagy. Blood 2013, 121, 1622–1632. [Google Scholar] [CrossRef] [Green Version]
  150. Happo, L.; Phipson, B.; Smyth, G.K.; Strasser, A.; Scott, C.L. Neither Loss of Bik Alone, nor Combined Loss of Bik and Noxa, Accelerate Murine Lymphoma Development or Render Lymphoma Cells Resistant to DNA Damaging Drugs. Cell Death Dis. 2012, 3, e306. [Google Scholar] [CrossRef] [Green Version]
  151. Frenzel, A.; Labi, V.; Chmelewskij, W.; Ploner, C.; Geley, S.; Fiegl, H.; Tzankov, A.; Villunger, A. Suppression of B-Cell Lymphomagenesis by the BH3-Only Proteins Bmf and Bad. Blood 2010, 115, 995–1005. [Google Scholar] [CrossRef] [PubMed]
  152. Jacobs, J.J.L.; Scheijen, B.; Voncken, J.W.; Kieboom, K.; Berns, A.; van Lohuizen, M. Bmi-1 Collaborates with c-Myc in Tumorigenesis by Inhibiting c-Myc- Induced Apoptosis via INK4a/ARF. Genes Dev. 1999, 13, 2678–2690. [Google Scholar] [CrossRef] [PubMed]
  153. Lee, S.C.W.; Phipson, B.; Hyland, C.D.; Leong, H.S.; Allan, R.S.; Lun, A.; Hilton, D.J.; Nutt, S.L.; Blewitt, M.E.; Smyth, G.K.; et al. Polycomb Repressive Complex 2 (PRC2) Suppresses Em-Myc Lymphoma. Blood 2013, 122, 2654–2664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Scott, C.L.; Gil, J.; Hernando, E.; Teruya-Feldstein, J.; Narita, M.; Martínez, D.; Visakorpi, T.; Mu, D.; Cordon-Cardo, C.; Peters, G.; et al. Role of the Chromobox Protein CBX7 in Lymphomagenesis. Proc. Natl. Acad. Sci. USA 2007, 104, 5389–5394. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Ke, F.; Voss, A.; Kerr, J.B.; O’Reilly, L.A.; Tai, L.; Echeverry, N.; Bouillet, P.; Strasser, A.; Kaufmann, T. BCL-2 Family Member BOK Is Widely Expressed but Its Loss Has Only Minimal Impact in Mice. Cell Death Differ. 2012, 19, 915–925. [Google Scholar] [CrossRef] [PubMed]
  156. Habib, T.; Park, H.; Tsang, M.; de Alborán, I.M.; Nicks, A.; Wilson, L.; Knoepfler, P.S.; Andrews, S.; Rawlings, D.J.; Eisenman, R.N.; et al. Myc Stimulates B Lymphocyte Differentiation and Amplifies Calcium Signaling. J. Cell Biol. 2007, 179, 717–731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Ricke, R.M.; Jeganathan, K.B.; van Deursen, J.M. Bub1 Overexpression Induces Aneuploidy and Tumor Formation through Aurora B Kinase Hyperactivation. J. Cell Biol. 2011, 193, 1049–1064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Shing, J.C.; Lindquist, L.D.; Borgese, N.; Bram, R.J. CAML Mediates Survival of Myc-Induced Lymphoma Cells Independent of Tail-Anchored Protein Insertion. Cell Death Discov. 2017, 3, 16098. [Google Scholar] [CrossRef] [Green Version]
  159. Ho, L.H.; Taylor, R.; Dorstyn, L.; Cakouros, D.; Bouillet, P.; Kumar, S. A Tumor Suppressor Function for Caspase-2. Proc. Natl. Acad. Sci. USA 2009, 106, 5336–5341. [Google Scholar] [CrossRef] [Green Version]
  160. Lu, Y.; Wu, Y.; Feng, X.; Shen, R.; Wang, J.H.; Fallahi, M.; Li, W.; Yang, C.; Hankey, W.; Zhao, W.; et al. CDK4 Deficiency Promotes Genomic Instability and Enhances Myc-Driven Lymphomagenesis. J. Clin. Investig. 2014, 124, 1672–1684. [Google Scholar] [CrossRef] [Green Version]
  161. Keller, U.B.; Old, J.B.; Dorsey, F.C.; Nilsson, J.A.; Nilsson, L.; MacLean, K.H.; Chung, L.; Yang, C.; Spruck, C.; Boyd, K.; et al. Myc Targets Cks1 to Provoke the Suppression of P27Kip1, Proliferation and Lymphomagenesis. EMBO J. 2007, 26, 2562–2574. [Google Scholar] [CrossRef] [PubMed]
  162. Hashwah, H.; Schmid, C.A.; Kasser, S.; Bertram, K.; Stelling, A.; Manz, M.G.; Müller, A. Inactivation of CREBBP Expands the Germinal Center B Cell Compartment, down-Regulates MHCII Expression and Promotes DLBCL Growth. Proc. Natl. Acad. Sci. USA 2017, 114, 9701–9706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Hunter, J.E.; Butterworth, J.A.; Zhao, B.; Sellier, H.; Campbell, K.J.; Thomas, H.D.; Bacon, C.M.; Cockell, S.J.; Gewurz, B.E.; Perkins, N.D. The NF-ΚB Subunit c-Rel Regulates Bach2 Tumour Suppressor Expression in B-Cell Lymphoma. Oncogene 2016, 35, 3476–3484. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Chen, J.; Shin, J.H.; Zhao, R.; Phan, L.; Wang, H.; Xue, Y.; Post, S.M.; Ho Choi, H.; Chen, J.S.; Wang, E.; et al. CSN6 Drives Carcinogenesis by Positively Regulating Myc Stability. Nat. Commun. 2014, 5, 5384. [Google Scholar] [CrossRef] [Green Version]
  165. Pei, X.H.; Bai, F.; Li, Z.; Smith, M.D.; Whitewolf, G.; Jin, R.; Xiong, Y. Cytoplasmic CUL9/PARC Ubiquitin Ligase Is a Tumor Suppressor and Promotes P53-Dependent Apoptosis. Cancer Res. 2011, 71, 2969–2977. [Google Scholar] [CrossRef] [Green Version]
  166. Arrate, M.P.; Vincent, T.; Odvody, J.; Kar, R.; Jones, S.N.; Eischen, C.M. MicroRNA Biogenesis Is Required for Myc-Induced b-Cell Lymphoma Development and Survival. Cancer Res. 2010, 70, 6083–6092. [Google Scholar] [CrossRef] [Green Version]
  167. Inoue, K.; Zindy, F.; Randle, D.H.; Rehg, J.E.; Sherr, C.J. Dmp1 Is Haplo-Insufficient for Tumor Suppression and Modifies the Frequencies of Arf and P53 Mutations in Myc-Induced Lymphomas. Genes Dev. 2001, 15, 2934–2939. [Google Scholar] [CrossRef] [Green Version]
  168. Vasanthakumar, A.; Lepore, J.B.; Zegarek, M.H.; Kocherginsky, M.; Singh, M.; Davis, E.M.; Link, P.A.; Anastasi, J.; le Beau, M.M.; Karpf, A.R.; et al. Dnmt3b Is a Haploinsufficient Tumor Suppressor Gene in Myc-Induced Lymphomagenesis. Blood 2013, 121, 2059–2063. [Google Scholar] [CrossRef] [Green Version]
  169. Yang, Z.; Shah, K.; Busby, T.; Giles, K.; Khodadadi-Jamayran, A.; Li, W.; Jiang, H. Hijacking a Key Chromatin Modulator Creates Epigenetic Vulnerability for MYC-Driven Cancer. J. Clin. Investig. 2018, 128, 3605–3618. [Google Scholar] [CrossRef] [Green Version]
  170. Baudino, T.A.; Maclean, K.H.; Brennan, J.; Parganas, E.; Yang, C.; Aslanian, A.; Lees, J.A.; Sherr, C.J.; Roussel, M.F.; Cleveland, J.L. Myc-Mediated Proliferation and Lymphomagenesis, but Not Apoptosis, Are Compromised by E2f1 Loss. Mol. Cell 2003, 11, 905–914. [Google Scholar] [CrossRef]
  171. Park, H.; Staehling, K.; Tsang, M.; Appleby, M.W.; Brunkow, M.E.; Margineantu, D.; Hockenbery, D.M.; Habib, T.; Liggitt, H.D.; Carlson, G.; et al. Disruption of Fnip1 Reveals a Metabolic Checkpoint Controlling B Lymphocyte Development. Immunity 2012, 36, 769–781. [Google Scholar] [CrossRef] [PubMed]
  172. Bouchard, C.; Lee, S.; Paulus-Hock, V.; Loddenkemper, C.; Eilers, M.; Schmitt, C.A. FoxO Transcription Factors Suppress Myc-Driven Lymphomagenesis via Direct Activation of Arf. Genes Dev. 2007, 21, 2775–2787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Pillonel, V.; Reichert, N.; Cao, C.; Heideman, M.R.; Yamaguchi, T.; Matthias, G.; Tzankov, A.; Matthias, P. Histone Deacetylase 1 Plays a Predominant Pro-Oncogenic Role in Eμ-Myc Driven B Cell Lymphoma. Sci. Rep. 2016, 6, 37772. [Google Scholar] [CrossRef] [Green Version]
  174. Vecchio, E.; Golino, G.; Pisano, A.; Albano, F.; Falcone, C.; Ceglia, S.; Iaccino, E.; Mimmi, S.; Fiume, G.; Giurato, G.; et al. IBTK Contributes to B-Cell Lymphomagenesis in Eμ-Myc Transgenic Mice Conferring Resistance to Apoptosis. Cell Death Dis. 2019, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Nilsson, L.M.; Keller, U.B.; Yang, C.; Nilsson, J.A.; Cleveland, J.L.; Roussel, M.F. Ink4c Is Dispensable for Tumor Suppression in Myc-Induced B-Cell Lymphomagenesis. Oncogene 2007, 26, 2833–2839. [Google Scholar] [CrossRef] [Green Version]
  176. Scherger, A.K.; Al-Maarri, M.; Maurer, H.C.; Schick, M.; Maurer, S.; Öllinger, R.; Gonzalez-Menendez, I.; Martella, M.; Thaler, M.; Pechloff, K.; et al. Activated Gp130 Signaling Selectively Targets B Cell Differentiation to Induce Mature Lymphoma and Plasmacytoma. JCI Insight 2019, 4, e128435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Osborne, L.C.; Duthie, K.A.; Seo, J.H.; Gascoyne, R.D.; Abraham, N. Selective Ablation of the YxxM Motif of IL-7Rα Suppresses Lymphomagenesis but Maintains Lymphocyte Development. Oncogene 2010, 29, 3854–3864. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Guerra, N.; Tan, Y.X.; Joncker, N.T.; Choy, A.; Gallardo, F.; Xiong, N.; Knoblaugh, S.; Cado, D.; Greenberg, N.R.; Raulet, D.H. NKG2D-Deficient Mice Are Defective in Tumor Surveillance in Models of Spontaneous Malignancy. Immunity 2008, 28, 571–580. [Google Scholar] [CrossRef] [Green Version]
  179. Gramling, M.W.; Eischen, C.M. Suppression of Ras/Mapk Pathway Signaling Inhibits Myc-Induced Lymphomagenesis. Cell Death Differ. 2012, 19, 1220–1227. [Google Scholar] [CrossRef] [Green Version]
  180. Barna, M.; Pusic, A.; Zollo, O.; Costa, M.; Kondrashov, N.; Rego, E.; Rao, P.H.; Ruggero, D. Suppression of Myc Oncogenic Activity by Ribosomal Protein Haploinsufficiency. Nature 2008, 456, 971–975. [Google Scholar] [CrossRef] [Green Version]
  181. Hawkins, E.D.; Oliaro, J.; Ramsbottom, K.M.; Ting, S.B.; Sacirbegovic, F.; Harvey, M.; Kinwell, T.; Ghysdael, J.; Johnstone, R.W.; Humbert, P.O.; et al. Lethal Giant Larvae 1 Tumour Suppressor Activity Is Not Conserved in Models of Mammalian T and B Cell Leukaemia. PLoS ONE 2014, 9, e87376. [Google Scholar] [CrossRef]
  182. Sotillo, R.; Hernando, E.; Díaz-Rodríguez, E.; Teruya-Feldstein, J.; Cordón-Cardo, C.; Lowe, S.W.; Benezra, R. Mad2 Overexpression Promotes Aneuploidy and Tumorigenesis in Mice. Cancer Cell 2007, 11, 9–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Brunelle, J.K.; Ryan, J.; Yecies, D.; Opferman, J.T.; Letai, A. MCL-1-Dependent Leukemia Cells Are More Sensitive to Chemotherapy than BCL-2-Dependent Counterparts. J. Cell Biol. 2009, 187, 429–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Campbell, K.J.; Bath, M.L.; Turner, M.L.; Vandenberg, C.J.; Bouillet, P.; Metcalf, D.; Scott, C.L.; Cory, S. Elevated Mcl-1 Perturbs Lymphopoiesis, Promotes Transformation of Hematopoietic Stem/Progenitor Cells, and Enhances Drug Resistance. Blood 2010, 116, 3197–3207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. de Clercq, S.; Gembarska, A.; Denecker, G.; Maetens, M.; Naessens, M.; Haigh, K.; Haigh, J.J.; Marine, J.-C. Widespread Overexpression of Epitope-Tagged Mdm4 Does Not Accelerate Tumor Formation In vivo. Mol. Cell Biol. 2010, 30, 5394–5405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Terzian, T.; Wang, Y.; van Pelt, C.S.; Box, N.F.; Travis, E.L.; Lozano, G. Haploinsufficiency of Mdm2 and Mdm4 in Tumorigenesis and Development. Mol. Cell Biol. 2007, 27, 5479–5485. [Google Scholar] [CrossRef] [Green Version]
  187. Tanaskovic, N.; Dalsass, M.; Filipuzzi, M.; Ceccotti, G.; Verrecchia, A.; Nicoli, P.; Doni, M.; Olivero, D.; Pasini, D.; Koseki, H.; et al. Polycomb Group Ring Finger Protein 6 Suppresses Myc-Induced Lymphomagenesis. Life Sci. Alliance 2022, 5, e202101344. [Google Scholar] [CrossRef] [PubMed]
  188. Talos, F.; Mena, P.; Fingerle-Rowson, G.; Moll, U.; Petrenko, O. MIF Loss Impairs Myc-Induced Lymphomagenesis. Cell Death Differ. 2005, 12, 1319–1328. [Google Scholar] [CrossRef] [Green Version]
  189. Contreras, J.R.; Palanichamy, J.K.; Tran, T.M.; Fernando, T.R.; Rodriguez-Malave, N.I.; Goswami, N.; Arboleda, V.A.; Casero, D.; Rao, D.S. MicroRNA-146a Modulates B-Cell Oncogenesis by Regulating Egr1. Oncotarget 2015, 6, 11023–11037. [Google Scholar] [CrossRef] [Green Version]
  190. Mu, P.; Han, Y.C.; Betel, D.; Yao, E.; Squatrito, M.; Ogrodowski, P.; de Stanchina, E.; D’Andrea, A.; Sander, C.; Ventura, A. Genetic Dissection of the MiR-17-92 Cluster of MicroRNAs in Myc-Induced B-Cell Lymphomas. Genes Dev. 2009, 23, 2806–2811. [Google Scholar] [CrossRef] [Green Version]
  191. Olive, V.; Sabio, E.; Bennett, M.J.; de Jong, C.S.; Biton, A.; McGann, J.C.; Greaney, S.K.; Sodir, N.M.; Zhou, A.Y.; Balakrishnan, A.; et al. A Component of the Mir-17-92 Polycistronic Oncomir Promotes Oncogene-Dependent Apoptosis. Elife 2013, 2, e00822. [Google Scholar] [CrossRef] [PubMed]
  192. Campbell, K.J.; Vandenberg, C.J.; Anstee, N.S.; Hurlin, P.J.; Cory, S. Mnt Modulates Myc-Driven Lymphomagenesis. Cell Death Differ. 2017, 24, 2117–2126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Nguyen, H.V.; Vandenberg, C.J.; Ng, A.P.; Robati, M.R.; Anstee, N.S.; Rimes, J.; Hawkins, E.D.; Cory, S. Development and Survival of MYC-Driven Lymphomas Require the MYC Antagonist MNT to Curb MYC-Induced Apoptosis. Blood 2020, 135, 1019–1031. [Google Scholar] [CrossRef] [PubMed]
  194. Au, A.E.; Lebois, M.; Sim, S.A.; Cannon, P.; Corbin, J.; Gangatirkar, P.; Hyland, C.D.; Moujalled, D.; Rutgersson, A.; Yassinson, F.; et al. Altered B-Lymphopoiesis in Mice with Deregulated Thrombopoietin Signaling. Sci. Rep. 2017, 7, 14953. [Google Scholar] [CrossRef] [Green Version]
  195. Kadariya, Y.; Tang, B.; Wang, L.; Al-Saleem, T.; Hayakawa, K.; Slifker, M.J.; Kruger, W.D. Germline Mutations in Mtap Cooperate with Myc to Accelerate Tumorigenesis in Mice. PLoS ONE 2013, 8, e67635. [Google Scholar] [CrossRef]
  196. Odvody, J.; Vincent, T.; Arrate, M.P.; Grieb, B.; Wang, S.; Garriga, J.; Lozano, G.; Iwakuma, T.; Haines, D.S.; Eischen, C.M. A Deficiency in Mdm2 Binding Protein Inhibits Myc-Induced B-Cell Proliferation and Lymphomagenesis. Oncogene 2010, 29, 3287–3296. [Google Scholar] [CrossRef] [Green Version]
  197. Lin, Y.H.; Wang, H.C.; Fiore, A.; Förster, M.; Tung, L.T.; Belle, J.I.; Robert, F.; Pelletier, J.; Langlais, D.; Nijnik, A. Loss of MYSM1 Inhibits the Oncogenic Activity of CMYC in B Cell Lymphoma. J. Cell Mol. Med. 2021, 25, 7089–7094. [Google Scholar] [CrossRef]
  198. Keller, U.; Nilsson, J.A.; Maclean, K.H.; Old, J.B.; Cleveland, J.L. Nfkb1 Is Dispensable for Myc-Induced Lymphomagenesis. Oncogene 2005, 24, 6231–6240. [Google Scholar] [CrossRef] [Green Version]
  199. Keller, U.; Huber, J.; Nilsson, J.A.; Fallahi, M.; Hall, M.A.; Peschel, C.; Cleveland, J.L. Myc Suppression of Nfkb2 Accelerates Lymphomagenesis. BMC Cancer 2010, 10, 348. [Google Scholar] [CrossRef] [Green Version]
  200. Zwolinska, A.K.; Heagle Whiting, A.; Beekman, C.; Sedivy, J.M.; Marine, J.C. Suppression of Myc Oncogenic Activity by Nucleostemin Haploinsufficiency. Oncogene 2012, 31, 3311–3321. [Google Scholar] [CrossRef] [Green Version]
  201. Nilsson, J.A.; Keller, U.B.; Baudino, T.A.; Yang, C.; Norton, S.; Old, J.A.; Nilsson, L.M.; Neale, G.; Kramer, D.L.; Porter, C.W.; et al. Targeting Ornithine Decarboxylase in Myc-Induced Lymphomagenesis Prevents Tumor Formation. Cancer Cell 2005, 7, 433–444. [Google Scholar] [CrossRef] [PubMed]
  202. Green, B.; Martin, A.; Belcheva, A. Deficiency in the DNA Glycosylases UNG1 and OGG1 Does Not Potentiate C-Myc-Induced B-Cell Lymphomagenesis. Exp. Hematol. 2018, 61, 52–58. [Google Scholar] [CrossRef] [PubMed]
  203. Martins, C.P.; Berns, A. Loss of P27Kip1 but Not P21Cip1 Decreases Survival and Synergizes with MYC in Murine Lymphomagenesis. EMBO J. 2002, 21, 3739–3748. [Google Scholar] [CrossRef] [PubMed]
  204. Shreeram, S.; Weng, K.H.; Demidov, O.N.; Kek, C.; Yamaguchi, H.; Fornace, A.J.; Anderson, C.W.; Appella, E.; Bulavin, D.V. Regulation of ATM/P53-Dependent Suppression of Myc-Induced Lymphomas by Wip1 Phosphatase. J. Exp. Med. 2006, 203, 2793–2799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Nemajerova, A.; Petrenko, O.; Trümper, L.; Palacios, G.; Moll, U.M. Loss of P73 Promotes Dissemination of Myc-Induced B Cell Lymphomas in Mice. J. Clin. Investig. 2010, 120, 2070–2080. [Google Scholar] [CrossRef] [PubMed]
  206. Galindo-Campos, M.A.; Lutfi, N.; Bonnin, S.; Martínez, C.; Velasco-Hernandez, T.; García-Hernández, V.; Martín-Caballero, J.; Ampurdanés, C.; Gimeno, R.; Colomo, L.; et al. Distinct Roles for PARP-1 and PARP-2 in c-Myc–Driven B-Cell Lymphoma in Mice. Blood 2022, 139, 228–239. [Google Scholar] [CrossRef]
  207. Cho, S.H.; Ahn, A.K.; Bhargava, P.; Lee, C.H.; Eischen, C.M.; McGuinness, O.; Boothby, M. Glycolytic Rate and Lymphomagenesis Depend on PARP14, an ADP Ribosyltransferase of the B Aggressive Lymphoma (BAL) Family. Proc. Natl. Acad. Sci. USA 2011, 108, 15972–15977. [Google Scholar] [CrossRef] [Green Version]
  208. Bolitho, P.; Street, S.E.A.; Westwood, J.A.; Edelmann, W.; MacGregor, D.; Waring, P.; Murray, W.K.; Godfrey, D.I.; Trapani, J.A.; Johnstone, R.W.; et al. Perforin-Mediated Suppression of B-Cell Lymphoma. Proc. Natl. Acad. Sci. USA 2009, 106, 2723–2728. [Google Scholar] [CrossRef] [Green Version]
  209. Xiao, W.; Hong, H.; Kawakami, Y.; Kato, Y.; Wu, D.; Yasudo, H.; Kimura, A.; Kubagawa, H.; Bertoli, L.F.; Davis, R.S.; et al. Tumor Suppression by Phospholipase C-Β3 via SHP-1-Mediated Dephosphorylation of Stat5. Cancer Cell 2009, 16, 161–171. [Google Scholar] [CrossRef] [Green Version]
  210. Wen, R.; Chen, Y.; Bai, L.; Fu, G.; Schuman, J.; Dai, X.; Zeng, H.; Yang, C.; Stephan, R.P.; Cleveland, J.L.; et al. Essential Role of Phospholipase Cγ2 in Early B-Cell Development and Myc-Mediated Lymphomagenesis. Mol. Cell Biol. 2006, 26, 9364–9376. [Google Scholar] [CrossRef] [Green Version]
  211. Fog, C.K.; Asmar, F.; Côme, C.; Jensen, K.T.; Johansen, J.V.; Kheir, T.B.; Jacobsen, L.; Friis, C.; Louw, A.; Rosgaard, L.; et al. Loss of PRDM11 Promotes MYC-Driven Lymphomagenesis. Blood 2015, 125, 1272–1281. [Google Scholar] [CrossRef] [PubMed]
  212. Mzoughi, S.; Fong, J.Y.; Papadopoli, D.; Koh, C.M.; Hulea, L.; Pigini, P.; di Tullio, F.; Andreacchio, G.; Hoppe, M.M.; Wollmann, H.; et al. PRDM15 Is a Key Regulator of Metabolism Critical to Sustain B-Cell Lymphomagenesis. Nat. Commun. 2020, 11, 3520. [Google Scholar] [CrossRef] [PubMed]
  213. Iotti, G.; Mejetta, S.; Modica, L.; Penkov, D.; Ponzoni, M.; Blasi, F. Reduction of Prep1 Levels Affects Differentiation of Normal and Malignant B Cells and Accelerates Myc Driven Lymphomagenesis. PLoS ONE 2012, 7, e48353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Garrison, S.P.; Jeffers, J.R.; Yang, C.; Nilsson, J.A.; Hall, M.A.; Rehg, J.E.; Yue, W.; Yu, J.; Zhang, L.; Onciu, M.; et al. Selection against PUMA Gene Expression in Myc-Driven B-Cell Lymphomagenesis. Mol. Cell Biol. 2008, 28, 5391–5402. [Google Scholar] [CrossRef] [Green Version]
  215. Meacham, C.E.; Ho, E.E.; Dubrovsky, E.; Gertler, F.B.; Hemann, M.T. In Vivo RNAi Screening Identifies Regulators of Actin Dynamics as Key Determinants of Lymphoma Progression. Nat. Genet. 2009, 41, 1133–1137. [Google Scholar] [CrossRef] [Green Version]
  216. Peintner, L.; Dorstyn, L.; Kumar, S.; Aneichyk, T.; Villunger, A.; Manzl, C. The Tumor-Modulatory Effects of Caspase-2 and Pidd1 Do Not Require the Scaffold Protein Raidd. Cell Death Differ. 2015, 22, 1803–1811. [Google Scholar] [CrossRef] [Green Version]
  217. Khattar, E.; Maung, K.Z.Y.; Chew, C.L.; Ghosh, A.; Mok, M.M.H.; Lee, P.; Zhang, J.; Chor, W.H.J.; Cildir, G.; Wang, C.Q.; et al. Rap1 Regulates Hematopoietic Stem Cell Survival and Affects Oncogenesis and Response to Chemotherapy. Nat. Commun. 2019, 10, 5349. [Google Scholar] [CrossRef] [Green Version]
  218. Zeng, H.; Yu, M.; Tan, H.; Li, Y.; Su, W.; Shi, H.; Dhungana, Y.; Guy, C.; Neale, G.; Cloer, C.; et al. Discrete Roles and Bifurcation of PTEN Signaling and MTORC1-Mediated Anabolic Metabolism Underlie IL-7-Driven B Lymphopoiesis. Sci. Adv. 2018, 4, eaar5701. [Google Scholar] [CrossRef] [Green Version]
  219. Thijssen, R.; Alvarez-Diaz, S.; Grace, C.; Gao, M.Y.; Segal, D.H.; Xu, Z.; Strasser, A.; Huang, D.C.S. Loss of RIPK3 Does Not Impact MYC-Driven Lymphomagenesis or Chemotherapeutic Drug-Induced Killing of Malignant Lymphoma Cells. Cell Death Differ. 2020, 27, 2531–2533. [Google Scholar] [CrossRef]
  220. Borland, G.; Kilbey, A.; Hay, J.; Gilroy, K.; Terry, A.; Mackay, N.; Bell, M.; McDonald, A.; Mills, K.; Cameron, E.; et al. Addiction to Runx1 Is Partially Attenuated by Loss of P53 in the Eμ-Myc Lymphoma Model. Oncotarget 2016, 7, 22973–22987. [Google Scholar] [CrossRef] [Green Version]
  221. Hoellein, A.; Fallahi, M.; Schoeffmann, S.; Steidle, S.; Schaub, F.X.; Rudelius, M.; Laitinen, I.; Nilsson, L.; Goga, A.; Peschel, C.; et al. Myc-Induced SUMOylation Is a Therapeutic Vulnerability for B-Cell Lymphoma. Blood 2014, 124, 2081–2090. [Google Scholar] [CrossRef] [PubMed]
  222. Hawkins, E.D.; Oliaro, J.; Ramsbottom, K.M.; Newbold, A.; Humbert, P.O.; Johnstone, R.W.; Russell, S.M. Scribble Acts as an Oncogene in Eμ-Myc-Driven Lymphoma. Oncogene 2016, 35, 1193–1197. [Google Scholar] [CrossRef] [PubMed]
  223. García-Fernández, M.; Kissel, H.; Brown, S.; Gorenc, T.; Schile, A.J.; Rafii, S.; Larisch, S.; Steller, H. Sept4/ARTS Is Required for Stem Cell Apoptosis and Tumor Suppression. Genes Dev. 2010, 24, 2282–2293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Jeong, S.M.; Lee, A.; Lee, J.; Haigis, M.C. SIRT4 Protein Suppresses Tumor Formation in Genetic Models of Myc-Induced B Cell Lymphoma. J. Biol. Chem. 2014, 289, 4135–4144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Old, J.B.; Kratzat, S.; Hoellein, A.; Graf, S.; Nilsson, J.A.; Nilsson, L.; Nakayama, K.I.; Peschel, C.; Cleveland, J.L.; Keller, U.B. Skp2 Directs Myc-Mediated Suppression of P27Kip1 yet Has Modest Effects on Myc-Driven Lymphomagenesis. Mol. Cancer Res. 2010, 8, 353–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Bagislar, S.; Sabò, A.; Kress, T.R.; Doni, M.; Nicoli, P.; Campaner, S.; Amati, B. Smyd2 Is a Myc-Regulated Gene Critical for MLL-AF9 Induced Leukemogenesis. Oncotarget 2016, 7, 66398–66415. [Google Scholar] [CrossRef] [Green Version]
  227. Cardone, M.; Kandilci, A.; Carella, C.; Nilsson, J.A.; Brennan, J.A.; Sirma, S.; Ozbek, U.; Boyd, K.; Cleveland, J.L.; Grosveld, G.C. The Novel ETS Factor TEL2 Cooperates with Myc in B Lymphomagenesis. Mol. Cell Biol. 2005, 25, 2395–2405. [Google Scholar] [CrossRef] [Green Version]
  228. Rounbehler, R.J.; Fallahi, M.; Yang, C.; Steeves, M.A.; Li, W.; Doherty, J.R.; Schaub, F.X.; Sanduja, S.; Dixon, D.A.; Blackshear, P.J.; et al. Tristetraprolin Impairs Myc-Induced Lymphoma and Abolishes the Malignant State. Cell 2012, 150, 563–574. [Google Scholar] [CrossRef] [Green Version]
  229. Finnberg, N.; Klein-Szanto, A.J.P.; El-Deiry, W.S. TRAIL-R Deficiency in Mice Promotes Susceptibility to Chronic Inflammation and Tumorigenesis. J. Clin. Investig. 2008, 118, 111–123. [Google Scholar] [CrossRef]
  230. Hussain, S.; Bedekovics, T.; Liu, Q.; Hu, W.; Jeon, H.; Johnson, S.H.; Vasmatzis, G.; May, D.G.; Roux, K.J.; Galardy, P.J. UCH-L1 Bypasses MTOR to Promote Protein Biosynthesis and Is Required for MYC-Driven Lymphomagenesis in Mice. Blood 2018, 132, 2564–2574. [Google Scholar] [CrossRef] [Green Version]
  231. Li, X.; Zhang, Y.; Zheng, L.; Liu, M.; Chen, C.D.; Jiang, H. UTX Is an Escape from X-Inactivation Tumor-Suppressor in B Cell Lymphoma. Nat. Commun. 2018, 9, 2720. [Google Scholar] [CrossRef] [PubMed]
  232. Moser, R.; Toyoshima, M.; Robinson, K.; Gurley, K.E.; Howie, H.L.; Davison, J.; Morgan, M.; Kemp, C.J.; Grandori, C. MYC-Driven Tumorigenesis Is Inhibited by WRN Syndrome Gene Deficiency. Mol. Cancer Res. 2012, 10, 535–545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Taylor, J.; Sendino, M.; Gorelick, A.N.; Pastore, A.; Chang, M.T.; Penson, A.V.; Gavrila, E.I.; Stewart, C.; Melnik, E.M.; Chavez, F.H.; et al. Altered Nuclear Export Signal Recognition as a Driver of Oncogenesis. Cancer Discov. 2019, 9, 1452–1467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Best, S.A.; Vandenberg, C.J.; Abad, E.; Whitehead, L.; Guiu, L.; Ding, S.; Brennan, M.S.; Strasser, A.; Herold, M.J.; Sutherland, K.D.; et al. Consequences of Zmat3 Loss in C-MYC- and Mutant KRAS-Driven Tumorigenesis. Cell Death Dis. 2020, 11, 877. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Regulation of the MYC gene and functions of MYC. Shown are pathways that activate or repress MYC transcription. Oncogenic translocations, such as (t8;14) found in human BL, juxtapose MYC and potent enhancers, resulting in high MYC expression. MYC has many interaction partners, leading to context-dependent regulation of biological processes. This scheme summarizes findings from various sources [2,8,20,21,22]. BCR—B-cell receptor.
Figure 1. Regulation of the MYC gene and functions of MYC. Shown are pathways that activate or repress MYC transcription. Oncogenic translocations, such as (t8;14) found in human BL, juxtapose MYC and potent enhancers, resulting in high MYC expression. MYC has many interaction partners, leading to context-dependent regulation of biological processes. This scheme summarizes findings from various sources [2,8,20,21,22]. BCR—B-cell receptor.
Cells 12 00037 g001
Figure 2. A meta-analysis of GEMMs of MYC-induced lymphoma. (A) Timeline of published articles using Eµ-Myc transgenic mice found on Pubmed with search terms: “Emu-myc”, “Eµ-myc”, or “Myc-induced lymphoma”. (B) Pie chart showing the type of mouse model used. Transplants do not fall under the term “GEMM”. (C,D) Ranked list of genes that decrease (C) or increase (D) survival, normalized to the control cohort for each study. Please refer to Table 1 for a complete list of genes.
Figure 2. A meta-analysis of GEMMs of MYC-induced lymphoma. (A) Timeline of published articles using Eµ-Myc transgenic mice found on Pubmed with search terms: “Emu-myc”, “Eµ-myc”, or “Myc-induced lymphoma”. (B) Pie chart showing the type of mouse model used. Transplants do not fall under the term “GEMM”. (C,D) Ranked list of genes that decrease (C) or increase (D) survival, normalized to the control cohort for each study. Please refer to Table 1 for a complete list of genes.
Cells 12 00037 g002
Figure 3. Common biological functions among the critical genes from Figure 2 for Eµ-Myc lymphomas were analyzed using Enrichr and the corresponding GO (gene ontology) terms [66,67]. Visualization was performed using the Revigo tool [68]. The color of the bubble corresponds to the adjusted p value (the redder, the lower the p value) and the size to the genes under the GO Term.
Figure 3. Common biological functions among the critical genes from Figure 2 for Eµ-Myc lymphomas were analyzed using Enrichr and the corresponding GO (gene ontology) terms [66,67]. Visualization was performed using the Revigo tool [68]. The color of the bubble corresponds to the adjusted p value (the redder, the lower the p value) and the size to the genes under the GO Term.
Cells 12 00037 g003
Figure 5. Critical genes for murine MYC-induced lymphoma play a role in human B-NHL. (A) Gene expression of the critical genes from Figure 2 is shown for normal and malignant human lymphoid tissue using the TNMplot tool [128]. (B) The mutation rate of the critical genes was assessed using cBioPortal [129,130]. A total of 41 out of 48 genes showed mutations in B-NHL. A total of 2117 samples from eight studies were included in the analysis. (C) Progression-free survival analysis of B-NHL patients with mutations (n = 81) or no mutations (n = 91) in the 48 critical genes derived from Eµ-Myc mice is shown. Overall survival was unaltered. Data and visualization are derived from cBioPortal [129,130].
Figure 5. Critical genes for murine MYC-induced lymphoma play a role in human B-NHL. (A) Gene expression of the critical genes from Figure 2 is shown for normal and malignant human lymphoid tissue using the TNMplot tool [128]. (B) The mutation rate of the critical genes was assessed using cBioPortal [129,130]. A total of 41 out of 48 genes showed mutations in B-NHL. A total of 2117 samples from eight studies were included in the analysis. (C) Progression-free survival analysis of B-NHL patients with mutations (n = 81) or no mutations (n = 91) in the 48 critical genes derived from Eµ-Myc mice is shown. Overall survival was unaltered. Data and visualization are derived from cBioPortal [129,130].
Cells 12 00037 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Winkler, R.; Piskor, E.-M.; Kosan, C. Lessons from Using Genetically Engineered Mouse Models of MYC-Induced Lymphoma. Cells 2023, 12, 37. https://doi.org/10.3390/cells12010037

AMA Style

Winkler R, Piskor E-M, Kosan C. Lessons from Using Genetically Engineered Mouse Models of MYC-Induced Lymphoma. Cells. 2023; 12(1):37. https://doi.org/10.3390/cells12010037

Chicago/Turabian Style

Winkler, René, Eva-Maria Piskor, and Christian Kosan. 2023. "Lessons from Using Genetically Engineered Mouse Models of MYC-Induced Lymphoma" Cells 12, no. 1: 37. https://doi.org/10.3390/cells12010037

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop