Next Article in Journal
Credibility of the Neutrophil-to-Lymphocyte Count Ratio in Severe Traumatic Brain Injury
Next Article in Special Issue
Spinal Cord Involvement in Adult Mitochondrial Diseases: A Cohort Study
Previous Article in Journal
Association of Metabolically Healthy and Unhealthy Obesity Phenotype with Markers Related to Obesity, Diabetes among Young, Healthy Adult Men. Analysis of MAGNETIC Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondria and Their Relationship with Common Genetic Abnormalities in Hematologic Malignancies

1
Department of Internal Medicine and Haematology, Semmelweis University, H-1085 Budapest, Hungary
2
Department of Pathology and Laboratory Medicine, Loma Linda University Health, Loma Linda, CA 92354, USA
3
Independent Researcher, Diamond Bar, CA 91765, USA
*
Author to whom correspondence should be addressed.
Life 2021, 11(12), 1351; https://doi.org/10.3390/life11121351
Submission received: 1 November 2021 / Revised: 29 November 2021 / Accepted: 29 November 2021 / Published: 7 December 2021
(This article belongs to the Special Issue Mitochondrial Medicine: From Rare to Common Diseases)

Abstract

:
Hematologic malignancies are known to be associated with numerous cytogenetic and molecular genetic changes. In addition to morphology, immunophenotype, cytochemistry and clinical characteristics, these genetic alterations are typically required to diagnose myeloid, lymphoid, and plasma cell neoplasms. According to the current World Health Organization (WHO) Classification of Tumors of Hematopoietic and Lymphoid Tissues, numerous genetic changes are highlighted, often defining a distinct subtype of a disease, or providing prognostic information. This review highlights how these molecular changes can alter mitochondrial bioenergetics, cell death pathways, mitochondrial dynamics and potentially be related to mitochondrial genetic changes. A better understanding of these processes emphasizes potential novel therapies.

1. Introduction

The contribution of mitochondria in hematologic malignancies as possible targets for therapy-resistant malignancies is emerging and has been discussed in some excellent recent review articles, such as the one by Barbato and co-workers [1]. The attention is not surprising, given the need for new therapeutical targets in therapy-resistant cases and due to the diverse role of mitochondria in normal and tumor tissue. Mitochondria are not just essential in ATP production via oxidative phosphorylation (OXPHOS) (see all the Abbreviations in Table S1); it is also vital in other biosynthetic and bioenergetic pathways, apoptosis regulation, intracellular calcium and reactive oxygen species (ROS) signaling, and iron storage, metabolism, and heme biosynthesis [2,3,4,5]. Furthermore, in the last two decades, our understanding of mitochondrial fission, fusion, mitophagy, and mitochondrial trafficking has largely evolved, giving us more insight into their function in health and disease [2,3].
In this review article, we will focus on our current understanding of the role of mitochondria in health and disease, with a focus on their role in carcinogenesis. This will be followed by their role in hematologic malignancies, organized and discussed by the different diseases or disease groups. In each disease or disease group, we will highlight how their common genetic abnormalities are related to mitochondria and what additional therapeutic potential they hold. These genetic changes or chromosomal abnormalities are either driver mutations or have prognostic value, often related to poor prognosis or therapy resistance in a certain disease. The genes we discuss are chosen using the current literature and the latest (2017) WHO Classification of Hematopoietic and Lymphoid Tissue Malignancies book as a guide [6], latter which we use in practice for the diagnoses of these diseases.

2. Mitochondria in Healthy and Tumor Cells

Many mitochondrial changes are advantageous for cellular adaptation and proliferation, making them essential elements of cancer cell survival. These alterations, therefore, serve as potential therapeutic targets in various tumors. In this chapter, we discuss the main mitochondria-related changes concerning carcinogenesis.

2.1. Oxidative Phosphorylation (OXPHOS) and Reactive Oxygen Species Production

2.1.1. Mitochondrial Metabolism

Mitochondria have a complex role in cellular metabolism, producing adenosine triphosphate (ATP) during OXPHOS; or from using NADH and FADH2, generated via cytoplasmic β-oxidation of fatty acids or the tricarboxylic acid cycle (TCA), latter also known as Krebs cycle. In addition, mitochondria are responsible for synthetizing lipids, amino acids, pyrimidine, and various other metabolic intermediates necessary for a functioning cell [4].

2.1.2. OXPHOS and Anaerobic Glycolysis

During OXPHOS, electrons are delivered through the mitochondrial respiratory complexes, making a proton gradient across the inner mitochondrial membrane, which is the source of ATP production during this process. OXPHOS requires the presence of oxygen and often generates ROS, especially at complexes I and II [3]. On the other hand, anerobic glycolysis takes place in the cytosol, not requiring oxygen. It produces lactate while generating less ATP than OXPHOS. Lactate can also be converted to pyruvate by lactate dehydrogenase (LDH), which can then enter the mitochondria to generate ATP through OXPHOS [7,8].

2.1.3. Mitochondrial Metabolism Adaptation

Mitochondrial metabolism adapts to different environmental stresses, making the cell capable of surviving under ever-changing conditions. Tumor cells, in general, have unique ways to maintain their high energy demand, with mitochondria in the center of reprogramming their metabolism [7]. Interestingly, tumor cell ATP production shifts to primarily increased glycolysis with ongoing OXPHOS, which can be reduced, normal, or increased [7,9]. The original observation that cancer cells use significantly more glucose and produce a large amount of lactate was made by Otto Warburg in the 1920s [10]. This finding is since referred to as the “Warburg effect” [7,11,12]. The reason why these changes are advantageous to tumors is diverse. First, anaerobic glycolysis can be 10–100 times faster than OXPHOS. In addition, glycolysis can be advantageous when the tumor cells compete for energy sources and oxygen. Glycolysis also can support amino acid and nucleic acid synthesis, essential for cell proliferation [11]. In addition, tumor cells can increase the efficacy of their glucose transporter, such as GLUT1, to import more glucose to the cytoplasm [13]. However, OXPHOS is still important in tumor cell metabolism, with many tumor cells having normal or increased OXPHOS along with accelerated glycolysis [12]. Increased OXPHOS has been linked to more aggressive tumor behavior [8].

2.1.4. Mitochondrial Metabolism as a Therapeutic Target

Based on these findings, mitochondrial metabolism has been proposed as potential tumor therapy in addition to inhibiting glycolysis. Several inhibitors are available in targeting mitochondrial metabolic pathways or enzymes, such as the ones in the electron transport chain (ETC) (metformin, ME344, IACS10759), enzymes in the TCA cycle (devimistat, enasidenib, ivosidenib), or glutaminase (converting glutamine to glutamate) (CB839, compound 27) [7,14]. As discussed later in this article, several of these pathways are affected by genetic alterations seen in hematologic malignancies. The ETC inhibitor ME344 has been shown to reduce acute myeloid leukemia (AML) cell viability and cell growth with no effect on normal hematopoietic cells [7,15]. In addition, inhibitors of a defective TCA cycle enzyme, the isocitrate dehydrogenase (AG120/ivosidenib, and AG221/enasidenib), have been FDA-approved for treating IDH-mutated relapsed/refractory AML [7,8,16].

2.1.5. ROS Generation and Function

As previously mentioned, OXPHOS is important in the production of ROS. Low levels of ROS are constantly produced in normal cells when electrons escape from the ETC, typically from complexes I and II, and react with molecular oxygen. ROS are important in intracellular signaling in low levels, such as in the response to hypoxia, growth factor-induced cell proliferation, and inflammatory response generation [3,9]. In addition, a moderate amount of ROS can activate intracellular signaling pathways critical in tumor cell survival, such as mitogen-activated protein kinase/extracellular signal-regulated protein kinases 1/2 (MAPK/ERK1/2), p38, c-Jun N-terminal kinase (JNK), and phosphoinositide-3-kinase/protein kinase B (PI3K/Akt) [17]. Given that higher levels of ROS would result in severe damage and eventually in apoptotic or necrotic cell death, a well-developed anti-oxidant system is in place to remove these highly reactive molecules [18]. Other than cell death induction, increased ROS production often leads to DNA mutations, especially in the mitochondrial DNA (mtDNA), given its proximity to the ETC proteins, and the minimal amount of mtDNA repair mechanisms existing. These mutations eventually can lead to carcinogenesis and tumor progression [19].

2.1.6. ROS in Tumor Therapy

Most tumor cells, including leukemia cells, produce an increased amount of ROS due to their increased metabolism and, therefore, can be targeted with ROS-generating agents to induce apoptosis [18]. Drugs using this mechanism of action are some widely used cytotoxic chemotherapies, such as cisplatin, 5-fluorouracil and paclitaxel. In addition, the ROS-producing anti-cancer drug procarbazine, has been approved for the treatment of stage III and IV Hodgkin lymphoma [7,14,20] and have been shown to be effective in other kinds of tumors as well [7]. The mitochondrial adaptation in carcinogenesis is summarized in Figure 1.

2.2. Mitochondrial DNA (mtDNA)

2.2.1. mtDNA

The human mtDNA are multi-copied and have circular genome, approximately 16 kb in size, encoding 13 mRNA, 22 transfer RNAs (tRNAs), and two ribosomal RNAs (rRNAs). The coding regions in the mtDNA are in close connection with partial overlaps. One longer non-coding region is present in the mtDNA, referred to as the control region, with transcription promoters for both strands. The 13 encoded proteins are essential subunits of the ETC I, II, IV, and V, with the majority of the mitochondrial proteins encoded by the nuclear DNA (nDNA). The nuclear-encoded mitochondrial proteins are imported from the cytosol by two translocase complexes: the translocase of the outer membrane (TOM) and the translocase of the inner membrane (TIM) [4,21,22,23,24].

2.2.2. mtDNA Transcription

Mitochondrial biogenesis requires coordinated nuclear and mitochondrial gene expressions for effective protein synthesis, regulated by various transcription factors. Several different proteins can directly or indirectly regulate mtDNA transcription. Most of those directly binding to the mtDNA bind to the D-loop area in the control region. The direct regulators include the mitochondrial RNA polymerase (POLRMT), transcription factors mitochondrial transcription factor A (TFAM), and mitochondrial transcription factor B1/2 (TFB1M/2M), a transcription elongation factor (TEFM), and one termination factor (mTERF1). TFAM can stabilize the mtDNA and initiate transcription, and is regulated by many intracellular regulatory proteins, such as the protein kinase A (PKA), extracellular regulated protein kinases 1/2 (ERK), or the peroxisome proliferator-activated receptor gamma coactivator 1-α (PGC-1α). A small change in PGC-1α’s concentration strongly affects mtDNA replication and transcription, and it is involved in mtDNA copy number alteration, mitochondrial dynamics, and OXPHOS regulation. Other molecules, such as signal transducers and activators of transcription 3 (STAT3) and cAMP response element-binding protein (CREB) can also bind to the D-loop area, promoting TFAM-independent gene expression. Other transcription enhancers, such as c-Jun and CCAAT enhancer-binding protein (CEBP)β, bind to different mtDNA regions [21,24,25,26]. In addition, POLG and POLG2 genes, encoding mtDNA polymerse γ, are involved in mtDNA replication, whereas TWNK encodes twinkle mtDNA helicase [27].

2.2.3. mtDNA Damage and Repair

mtDNA is more vulnerable to insults, such as ROS production, than nDNA due to its proximity to the ROS production site, smaller amount of repair mechanisms existing, and the lack of protection by histones [23]. Replication errors, however, also contribute to a large number of mtDNA mutations [25]. mtDNA repair mechanisms are not well characterized, and the damaged mtDNA can also be degraded. Repair mechanisms include base excision repair, and direct reversal, with evidence of mismatch repair and double-strand break repair possibly existing in the human mitochondria [23,28]. Importantly, mtDNA mutations can be heteroplasmic or homoplasmic, indicating the presence of a mixture or identical mtDNA genotypes in the mitochondria, respectively [7].

2.2.4. mtDNA Translation as a Therapeutic Target

Inhibitors of mitochondrial transcription at the POLRMT level have been developed, such as IMT1 and IMT1B. They successfully decrease OXPHOS and ATP production via interfering with the transcription of ETC proteins. In addition, these inhibitors reduced tumor cell growth and cell viability in ovarian and colon cancer models [29]. Mitochondrial protein translation can also be inhibited by drugs, such as Tigecycline, which could promote cell death in AML cells, and can improve the effectiveness of tyrosine kinase inhibitor Imatinib in chronic myeloid leukemia (CML) models [7,30,31].

2.2.5. mtDNA Changes in Cancer

Not only are mtDNA mutations more frequent in cancer cells, but altered mtDNA copy number has also been associated with carcinogenesis [7,32,33,34]. While some types of cancer cells exhibit decreased mtDNA copy number, lymphoma and certain leukemia cells are typically associated with increased mtDNA copy number [32,35,36]. In pediatric AML cells increased TFAM and POLG expression have been described along with increased mtDNA copy number, reversible with PGC-1α inhibition [35]. In addition, mtDNA copy number was found to be increased in pediatric acute lymphoblastic leukemia (ALL) samples, which significantly decreased after treatment [37]. Additionally, there is growing evidence that mtDNA polymorphisms can influence drug response in various cancers [38]. The regulation of mtDNA replication, transcription and translation are summarized in Figure 2.

2.3. Apoptosis and Necrosis Regulation

Apoptotic pathways and their role in carcinogenesis have been well studied. It can be initiated via the extrinsic or intrinsic pathway leading to caspase activation, the central effectors of apoptotic cell death [22]. The intrinsic pathway is activated by intracellular stress signals, which trigger the homo-oligomerization of bak and bax proteins, creating pores on the mitochondrial outer membrane (MOM) and releasing cytochrome c from the mitochondrial intermembrane space. Bak and bax, along with PUMA and NOXA are pro-apoptotic members of the B-cell lymphoma 2 (Bcl 2) protein family, which are inhibited under normal circumstances. Anti-apoptotic members of the Bcl-2 protein family, such as bcl-2, bcl-Xl, and mcl-1, can bind and inhibit bak and bax in their inactivated or activated forms. Cytochrome c induces a so-called apoptosome formation, able to activate caspase 9, which in turn can activate the executioner caspase, caspase 3 [39,40,41]. Interestingly, the most famous anti-apoptotic protein in cancers, p53 has a much more complex effect on mitochondria than just interfering with the apoptotic pathways, described in more details in Section 3.5.2. The extrinsic pathway of apoptosis is initiated via the activation of a death receptor, which activates caspase 8 and then caspase 3 [22,39].

2.3.1. Necrosis and Necroptosis

Unlike apoptosis, necrosis is an unregulated process of cell death caused by severe injury. However, necroptosis, a controlled form of necrosis was described in 2005 by Degterev and co-workers [42], which has been further characterized since [39]. The main control proteins of necroptosis are the receptor-interacting proteins 1 (RIP1) and 3 (RIP3), activated in an apoptosis-deficient environment. Interestingly, RIP1 can induce apoptosis via caspase 8 activation, but it also initiates necroptosis via recruiting RIP3. RIP3 can interact with the mixed lineage kinase domain-like pseudokinase (MLKL) protein, which becomes an oligomer after the interaction and migrates to the cell membrane, increasing its permeability [39].

2.3.2. Apoptosis as a Therapeutic Target

The connection between the pro- and anti-apoptotic genes and carcinogenesis is well described in the literature. Increased expression of anti-apoptotic proteins, such as bcl-2 and bcl-XL are seen in untreated AML samples and advanced MDS, with higher expression associated with worse prognosis or therapy resistance. A Bcl-2 inhibitor, venetoclax (ABT-199), has been used to treat AML and relapsed Chronic lymphocytic leukemia (CLL) with 17p deletion. Similarly, altered anti-apoptotic protein mcl-1 expression was described in hematologic malignancies, such as AML, multiple myeloma (MM), and B-cell acute lymphoblastic leukemia (B-ALL). Conversely, bax level can be high with either therapy sensitivity or therapy resistance in AML. Their localization is, however, important. The cells are more sensitive to chemotherapy when bax is mainly in the mitochondria, and not in the cytosol. Interestingly, BTSA1, a bax activator, effectively suppress apoptosis in human AML xenografts [41,43,44,45].

2.3.3. Necroptosis as a Therapeutic Target

Necroptosis is also a potentially viable treatment for cancers that have acquired resistance to apoptotic pathways. The serine-threonine kinases RIP1 and RIP3 have been shown to induce mitochondrial fission through activation of dynamin-related protein1 (Drp1) via serine phosphorylation. Subsequently, Drp1 can interact with another regulator protein, Fis1, to induce mitochondrial fission. Ultimately, RIP1-RIP3 induced mitochondrial fission will lead to the generation of reactive oxygen species and eventually cell death via necroptosis [46]. Necroptosis, however, can also occur independently of mitochondrial fission, which might be cell-type dependent [47].

2.4. Mitochondrial Fission and Fusion, Mitophagy

2.4.1. Mitochondrial Dynamics

The changes in mitochondrial number and morphology are described as mitochondrial dynamics [41]. Mitochondria are highly dynamic structures that form a cytosol network, responsive to intracellular changes, such as altered metabolic needs. The adaptation of the mitochondrial network is via changes in mitochondrial dynamics, including fission, fusion, and “mitophagy”. These processes regulate the number of mitochondria in the cells, and they are also important in the redistribution or elimination of mtDNA. In addition, autophagic degradation of the mitochondria, known as “mitophagy”, is responsible for removing healthy mitochondria when less is needed or damaged mitochondria to maintain a healthy pool in the cell [2,23,48,49].

2.4.2. Mitochondrial Fission

Mitochondrial fission results in an increased number of mitochondria, with dynamin-related protein 1 (Drp1), a large GTPase, being the central effector protein [50,51]. Drp1 is primarily found in the cytosol and is translocated to the MOM at the time of fission [49,51] to any of its mitochondrial receptors: fission protein 1 (Fis1), mitochondrial fission factor (Mff), or mitochondrial elongation factors (MIEFs) 1 and 2. Amongst the Drp1 receptors, Fis1 is the least significant in mammals [49]. When Drp1 is recruited to the mitochondrial surface, it polymerizes and forms a ring-like structure around the outer membrane of the mitochondria, where fission occurs via its GTPase activity [49]. Mitochondrial fission results in changes in mitochondrial metabolism, most notably in decreased OXPHOS capacity [52]. An increase in Drp1 and fission occurs due to different stress mechanisms, and as a part of cell adaptation mechanisms. A significant decrease or loss of mitochondrial outer membrane potential (MOMP) of damaged mitochondria for example results in mitochondrial fragmentation. Interestingly this can be due to increased fission but can also result in shortened mitochondria due to the collapse of the mitochondrial structure [5]. Fission is also an important part of removing aberrant mitochondria via selective mitophagy, such as following uneven fission, when one daughter mitochondrion with low MOMP would be eliminated [25].

2.4.3. Mitochondrial Fusion

On the other hand, mitochondrial fusion results in a decreased number, and more interconnected mitochondria, with increased oxidative capacity and MOMP [25,52]. Key proteins involved in mitochondrial fusion are mitofusin 1 (Mfn1) and 2(Mfn2) on the MOM and optic atrophy 1 (OPA1) on the inner mitochondrial membrane (IMM). Both Mfn1 and 2 are transmembrane proteins with GTPase activities that can form homotypic (Mfn1-Mfn1 or Mfn2-Mfn2) or heterotypic (Mfn1-Mfn2) connections to fuse the outer mitochondrial membranes of two mitochondria, while the homotypic Opa1-Opa1 complexes fuse the internal mitochondrial membranes [25,48,53]. Additionally, Mfn2 is involved in the endoplasmic reticulum (ER)—mitochondria tethering; in mitophagy, via ubiquitination (see later in this chapter); and in apoptosis regulation (see later in this chapter) [22]. Damaged mitochondria with reduced MOMP are unlikely to undergo fusion, partially due to decreased OPA1 levels, which helps to eliminate and isolate those mitochondria from the healthy pool. In addition, higher OPA1 expression is linked to decreased mitophagy [25,52].

2.4.4. Mitophagy

Besides clearing dysfunctional mitochondria, mitophagy is important in various adaptive responses, where reduced overall mitochondrial mass is desirable [25], and is also involved in controlling inflammatory responses in immune cells [54]. The major pathways of selective mitophagy are the ubiquitin-mediated and the receptor-mediated mitophagies, with evidence of significant crosstalk between them [54,55,56]. Ubiquitin-mediated mitophagy can be parkin-dependent or parkin-independent. The parkin-dependent PINK1-parkin pathway is the classic and most studied pathway of mitophagy. It eliminates defective organelles, which have for example reduced MOMP. Reduced MOMP results in an increased number of the phosphatase and tensin homolog (PTEN)-induced putative kinase 1 (PINK1) protein on the outer mitochondrial membrane (OMM), as the unfunctional mitochondria cannot transport PINK1 to the IMM to be cleaved [55,56]. Therefore, in mitochondrial membrane depolarization, mitochondrial complex dysfunction, mtDNA mutations, or proteotoxicity, PINK1 accumulates at the OMM rapidly and recruits parkin from the cytosol via phosphorylation. Activated parkin ubiquitinates OMM proteins, such as VDAC1 and Mfn2, serving as an “eat me” signal for damaged organelles. The autophagy adaptor proteins, such as p62, Optn, and Ndp52 recognize these signals and initiate autophagosome formation. PINK1 can amplify the autophagy signals via phosphorylation of ubiquitin and poly-ubiquitin chains. PINK1 also increases Drp1 activity, resulting in the fragmentation of the damaged mitochondria, which can be cleared easier [25,53,55,56]. In addition to parkin, other ubiquitin E3 ligases can also trigger mitophagy via a PINK1-independent manner. These proteins include Gp78, SMURF1, SIAH1, MUL1, and Arih1, which recruit autophagy adaptor proteins to complete the mitophagy process [56]. The receptor-mediated mitophagy pathway on the other hand is initiated by one of the mitophagy receptors, such as the Nip3-like protein X or Bnip3L (Nix), Bcl-2 family adenovirus E1B 19 kDa-interacting protein 3 (Bnip3), activating molecule in BECLIN1-regulated autophagy (ambra1) or FUN14 domain containing 1 (fundc1). These proteins can directly bind to LC3 in the phagophore, unlike in the non-receptor mediated pathway, where autophagy receptors are also needed for LC3 binding [25,54,55,57].

2.4.5. Mitochondrial Dynamics in Different Cell Cycles

Mitochondrial fusion predominates G1/and early S-phases during the cell cycle, producing more ATP for cell growth. On the other hand, during S/G2/and M phases, mitochondrial fission is increased to provide equal distribution of mitochondria between daughter cells and to reduce OXPHOS and subsequent ROS generation and mtDNA mutations [52,58,59]. Blocking Drp-1 in vitro resulted in fewer cancer cells in S-phase and increased apoptosis [60,61].

2.4.6. Mitochondrial Dynamics and OXPHOS

Both bioenergetic changes and apoptosis regulation via altered mitochondrial dynamics are important in carcinogenesis and thus hold tremendous therapeutic potential for various malignancies. As for the bioenergetic changes, fission generally reduces, whereas fusion increases OXPHOS efficacy [9,52]. Although both disrupted fission and fusion have been linked to impaired mitochondrial energy production in some cases, unopposed fusion has been shown to produce more ATP through enhanced cristae density and ATP synthase dimerization [62,63,64]. Therefore, the observation that mitochondrial dynamics are shifted towards fission in most tumors aligns with the decreased OXPHOS previously described in various cancer models [65].

2.4.7. Mitochondrial Dynamics and Apoptosis

The relationship between apoptosis and mitochondrial fission/fusion is complex and bi-directional. First, both Drp1 and Mfn2 are related to the pro-apoptotic protein Bax. Both proteins can be co-localized with Bax at the MOM, with Drp1 being recruited by Bax after apoptosis induction, promoting fission; and Mfn2 binding to Bax, leading to mitochondrial fusion inhibition [52,53,66,67]. Furthermore, Drp1 plays an essential role in cytochrome c release and in pro-apoptotic Bak activation [51,68]. Downregulation or blocking of Drp1 has been shown to have anti-apoptotic effect in different in vitro models [53,68,69,70]. Of note, apoptosis can occur without mitochondrial fragmentation, with some studies showing that blocking fission only slows down apoptosis rather than entirely preventing it [53,68,71,72]. Fusion protein OPA1 on the other hand can inhibit cytochrome c release and resultant apoptosis [73,74]. Additionally, bak activation can induce mitochondrial fragmentation, whereas bcl-2 and bcl-Xl shift mitochondria to fusion [75]. Nevertheless, mitochondrial fission can induce apoptosis via increased ROS production [76].
Despite many models proving the pro-apoptotic effect of Drp-1 and mitochondrial fission, in most tumors, Drp1-mediated fission is anti-apoptotic. Drp1 expression is high in various tumors, which is associated with increased cell proliferation, metastatic potential, and poor outcome in many tumors, including AML [47,50,61,65,77,78,79,80]. Loss or inhibition of Drp1 was shown to inhibit tumor growth in multiple models [47]. Interestingly, the pro-apoptotic effect of the Drp1 receptor Fis1 was described, which seems to be unrelated to its fission function [58]. Blocking Mff has also been shown to inhibit tumor growth and cell proliferation in an in vivo model [81]. In addition, Mfn2 expression is decreased in several tumors, whereas increased Mfn2 expression has been linked to tumor suppression. In line with these findings, low Mfn2 expression in breast cancer patients is associated with poor outcome, and overexpression of Mfn2 was shown to induce apoptosis in a hepatocellular carcinoma model [65,82,83,84]. The explanation for the discrepancy in tumor cells versus other models is likely due to that increased mitochondrial fission in tumor cells can protect them from Ca2+-dependent apoptosis by limiting mitochondrial Ca2+ overload [58,85]. In addition, increased MOM surface impairs bax insertion and resultant bax-induced apoptosis [65].

2.4.8. Mitophagy and Apoptosis

Mitophagy is also associated with apoptosis in varied ways. Excessive mitophagy induces apoptotic cell death [55]. The PINK1-parkin pathway can protect the cell from apoptosis via VDAC1 monoubiquitination, whereas it induces mitophagy via polyubiquitination of the same protein [86]. Mitophagy receptors Nix and Bnip3 have a BH3 domain similar to the Bcl-2 family proteins and can induce apoptosis [55,87].

2.4.9. Mitochondrial Dynamics in Cancer

In sum, mitochondrial dynamics is a key regulator of apoptosis and cellular metabolism in cancer cells. Therefore, blocking Drp1-dependent mitochondrial fission, MFF, or increasing Mfn2 levels, increasing mitophagy receptor density on the mitochondria, or decreasing VDAC1 monoubiquitination are viable therapeutic options for several types of malignancies, including hematologic malignancies, as discussed later.

2.4.10. Mitophagy in Cancer

Mitophagy is important both in normal hematopoietic cell development and in various hematologic malignancies, discussed in detail in a recent review by Stergiou and Kapsogeorgou [88]. In AML, decreased mitophagy was shown to reduce cell proliferation. Erythroblasts in low-risk MDS demonstrate increased mitophagy, whereas in the high-risk group, accumulation of enlarged mitochondria was present. In addition, mitophagy was found to be prominent in the myeloid lineage in MDS. Interestingly, mitophagy receptor, BNIP3 and NIX, expressions were decreased in MDS, latter in the high-risk group [88,89,90,91,92,93]. Mitophagy inhibitors and inducers are listed in Section 4, Table 3. Mitochondrial dynamics are summarized in Figure 3.

2.5. Mitochondrial Trafficking

2.5.1. Mechanisms of Mitochondrial Trafficking

Mitochondria travel in the cytosol via a microtubular network of β-tubulin or actin filaments. Mitochondria’s microtubular transport primarily uses Rho GTPases, such as Mitochondrial receptor protein 1 and 2 (Miro1/2), encoded by RHOT1 and RHOT2, respectively. Miro1/2 binds to the kinesin-1/3 motor via the trafficking kinesin protein 2 (Trak2), encoded by the TRAK2 gene, moving the mitochondria towards the + end of the microtubule, called anterograde transport. The KIF5B gene encodes the kinesin-1 heavy chain. Besides Miro, additional kinesin-binding proteins, such as syntabulin (encoded by SYBU), fasciculation and elongation protein zeta 1 (encoded by FEZ1) and Ran-binding protein 2 (encoded by RANBP2), have been described, mostly in neurons and in some tumors of non-nervous system origin. For retrograde transport, Miro is forming a complex with a dynein motor and the bicaudal 2 (BiCD2) adaptor protein. There are four ways described in the literature on how mitochondrial detachment can occur from the microtubule when there is a mitochondrium-Miro-Trak2-kinesin complex present: 1. Myosin motor (MYO)-dependent binding to actin filaments 2. Mitochondrial anchoring to the microtubules via syntaphilin (SNPH), disrupting kinesin-Trak2 binding 3. Calcium-induced detachment resulting in conformational change in Miro and in the kinesin protein, and 4. The irreversible proteasomal degradation of the kinesin-1/Trak complex. All these can inhibit mitochondrial transfer along the microtubules. In addition, mitochondrial transfer along the actin filaments is typically short-range and is executed by MYOs, such as MYO19, MYO6, and MYO5. MYOs can move mitochondria both anterograde and retrograde, with the exact mechanism of mitochondrial binding unknown [50,52,94,95].

2.5.2. Mitochondrial Trafficking in Cancer

Excellent recent review articles are available on the changes in mitochondrial trafficking in cancer [94,95,96], which we discuss here briefly. In recent years, changes in mitochondrial trafficking and mitochondrial localization have been described concerning cancer progression. Interestingly, the intracellular ATP:ADP ratio changes not only with a change in the mitochondrial mass but also with their position inside the cells [97]. Increased mitochondrial density, often with evidence of intense glycolysis at the plasma membrane has been described in invasive cells [96]. These changes likely help to have the energy source available right at the plasma membrane, supporting signal transduction and cell invasion [94]. Increased expression or mutation of genes involved in microtubule-based movements, such as TRAK1, MIRO1/2, KIF5B, RANBP2, or FEZ1, are associated with enhanced cell proliferation and invasion and chemotherapy resistance in some solid tumors. SNPH is typically downregulated in cancers, with its depletion or loss resulting in a worse prognosis and enhanced cell invasion. On the other hand, we have minimal knowledge on how the actin-MYO mitochondrial trafficking is altered in cancer cells. Furthermore, more research is needed to clarify how mitochondrial trafficking is affected in hematologic malignancies [94]. As discussed in later chapters, scattered data exist on their role in multiple myeloma (MM) and in MYC-related malignancies.

3. Mitochondrial Changes in Relation to Driver Mutations, Genetic and Chromosome Abnormalities in Hematologic Malignancies

3.1. Mitochondria in Hematologic Malignancies

Primary mitochondrial issues in hematologic diseases, such as in Pearson syndrome or some congenital neutropenias, are rare [98,99]. In addition, according to a small study, primary mitochondrial disorders are not typically associated with hematologic malignancies but with anemia, thrombocytopenia, thrombocytosis, leukopenia, or eosinophilia [100]. Despite the fact that mitochondrial changes are not the root cause of hematologic malignancies, tumor cells heavily rely on them for their survival and therefore are great potential targets in their treatments [101]. Here we discuss these changes and treatment options by diseases and related genetic changes.

3.2. Myelodysplastic Syndrome (MDS) and Acute Myeloid Leukemia (AML)

3.2.1. Myelodysplastic Syndrome (MDS)

MDSs are clonal hematopoietic malignancies predominating in people above 70 years. It is characterized by ineffective hematopoiesis leading to cytopenias and can transform to AML in approximately 30–40% of the cases. The etiology of MDS is unknown in 85% of the cases. Numerous cytogenetic alterations, hereditary or acquired mutations in somatic and mitochondrial genes, hereditary genetic syndromes, or environmental factors are responsible for developing the disease [102].

3.2.2. Acute Myeloid Leukemia (AML)

AML is a phenotypically and genetically heterogeneous group of hematological malignancies characterized by clonal expansion of myeloid precursors with diminished capacity for differentiation [103]. AML represents 15–20% of acute leukemia cases in children and 80% in adults [104]. Population aging contributes to a significant increase in AML in Europe, as its incidence rises markedly in patients above 60 years. With advanced age, there is a decreased incidence of AML with recurrent genetic abnormalities [105]. In contrast, the incidence of other AML categories, such as AML with myelodysplasia-related changes (MRC-AML), or therapy-related AML (tAML), increases with age, comprising about 19% and 7% of AML cases, respectively [106,107,108,109]. According to the current World Health Organization (WHO) Classification of Tumors of Hematopoietic and Lymphoid Tissues, numerous genetic changes are highlighted, defining a distinct subtype of diseases (AML with recurrent genetic abnormalities), such as the ones involving RUNX1, CBFB-MYH11, PML-RARA, BCR-ABL1 etc. Also, there is a separate category for myeloid neoplasms with germline mutations, including DDX41 and CEBPA, amongst others. In addition, several genes are highlighted that provide prognostic information both in AML and MDS [6]. Additionally, various publications report further genetic abnormalities in correlation with prognosis.

3.2.3. Cytogenetic and Molecular Genetic Alterations in MDS and MDS/AML

The most common genetic alterations in MDS/AML are clonal chromosomal abnormalities, found in approximately 30–80% of patients. In the remaining 20–70% of patients with normal karyotype, point mutations, microdeletions, amplifications, epigenetic changes, or copy number neutral loss, such as uniparental disomy provide the genetic basis of the disease [102]. The complexity of cytogenetic abnormalities is related to the clinical severity of the disease: the more complex cytogenetic alteration is associated with poorer disease outcomes. Complex abnormalities can be further divided by the presence or absence of TP53 mutation/alteration. Acquired somatic molecular mutations are seen in 80–90% of MDS patients, which can be grouped by gene function as follows [102]:
(1)
Epigenetic regulators and chromatin remodeling factors: methylcytosine dioxygenases of the ten eleven translocated family (TET2), additional sex comb-like genes (ASXL1), DNA-methyltransferase family (DNMT3A), isocitrate dehydrogenases (IDH 1/2), enhancer of zeste homolog 2 (EZH2)
(2)
mRNA splicing factors: SF3B1, SRSF2, U2AF1
(3)
Transcriptional factors: RUNX1, TP53
(4)
Signaling molecules: NRAS, KRAS
(5)
Cohesin complex: STAG2
The most common mutations present in more than 5–10% of the patients include TET2, SF3B1, ASXL1, SRSF2, DNMT3A, TP53, EZH2, IDH1/2, NRAS, BCOR, and RUNX1; and approximately further 30 genes are altered in about 1% of patients. Most of these mutations carry a poor prognosis, except for SF3B1. It is to note that some elderly (70–80 years old) people carry one or more of these mutations, typically DNMT3A, TET2, or ASXL1, and less often JAK2, PPM1D, SF3B1, SRSF2, or TP53, with no hematologic malignancy, and the term “clonal hematopoiesis of indeterminate potential” was established to describe this condition [6,102,110].

3.2.4. Genetic Alterations in De Novo AML

Besides the mutations mentioned above in the context of MDS and post MDS AML, some additional mutations and chromosomal rearrangements have been described in the pathogenesis of AML [111,112]. These are often classified into two broad classes of mutations: Class I mutations that confer proliferative and survival advantages and Class II mutations affecting cell differentiation and apoptosis. Recent studies, however, have identified further mutations that do not conform to any of the two classes, many promoting epigenetic modifications (Table 1). There is a collaboration between the three classes of gene alterations, which is associated with the appearance of AML phenotype [6,113,114,115,116].

3.2.5. MDS, AML and Mitochondrial Metabolism

Metabolic changes are numerous in AML and have been reviewed in excellent recent articles from Barbato and co-workers and Panuzzo and co-workers [1,117]. Here we are going to discuss these metabolic changes in relation to genetic alterations in AML and MDS. In short, in AML, increased SIRT3-mediated OXPHOS has been described along with increased mitochondrial fitness and ROS levels, leading to drug resistance. In addition, PDH and α-KGD inhibitors, IACS-010759 and Venolex, could re-sensitize AML cells to chemotherapy [1]. Also, a previous study showed that 8% of AML patients had mutations in their ETC complex genes, with a majority affecting Complex IV [117].

3.2.6. MDS, AML and Mitochondrial Dynamics, Mitochondrial Transfer

Overexpression of mitochondrial fission receptor Fis1-mediated mitophagy was found to be essential for AML cell proliferation and differentiation in vitro. In the same model, depleting Fis1 via shRNA-mediated FIS1 knock-down resulted in cell cycle arrest, loss of self-renewal and attenuated myeloid cell differentiation [118]. In addition, increased Drp1-dependent mitophagy has been implicated in the mechanism of FLT3-internal tandem duplication (ITD) inhibition therapy in AML [119]. In a previous study, AML patients with high FIS1 expression likely to be chemotherapy-resistant and were more frequently M0/M1 FAB subtypes [77]. Similarly, mitochondrial fragmentation and increased Drp1 is also associated with MDS with CBL exon deletion and RUNX1 mutation, promoting dysplasia and impaired granulopoiesis [120]. Additionally, Drp-1-dependent mitochondrial fragmentation is seen in the mesenchymal stromal cells in MDS with iron overload, contributing to cell damage [121]. Interestingly, bone marrow stromal cells can transfer mitochondria to AML cells [122,123].

3.2.7. Genes Discussed

Here we aim to summarize the relationship between different genetic mutations, mitochondrial metabolism, mitochondrial dynamics, and the development of MDS/AML in selected genes. It is to note that there is a significant overlap in genetic alterations in these two diseases and that some genetic alterations are also present in other hematologic malignancies.

3.3. Epigenetic Regulators

Epigenetic regulation or modification means different mechanisms that alter gene expression without alterations in nucleotide sequence. The two main mechanisms of epigenetic regulation are DNA methylation and histone modifications. The interaction between DNAm, mitochondria, and transcriptional factors in hematologic malignancies are summarized in Figure 4.

3.3.1. DNA Methylation (DNAm)

DNAm mechanisms. DNAm entails the conversion of cytosine to 5-methylcytosine (5mC) to the 5th carbon of the ring of cytosine, predominantly due to the DNA methyltransferase (DNMT) enzyme transferring a methyl group from S-adenosylmethionine (SAM). This conversion is usually found within CpG dinucleotide sites, regions of the DNA in which a cytosine nucleotide is immediately followed by a guanine nucleotide in a linear sequence in the 5′ to 3′ direction [124]. TET enzymes catalyze the conversion of 5-methylcytosine (5-mC) to 5-hydroxymethylcytosine (5-hmC) and promote DNA demethylation. Besides DNA demethylation, it also plays a role in histone modification: the mechanism that may account for TET-mediated gene activation is the recruitment of O-linked β-D-N-acetylglucosamine (O-GlcNAc) transferase (OGT) to chromatin [125]. DNAm and related genes. Such epigenetic DNA modifications occur in intronic, exonic and intergenic regions [126]. They are involved in the regulation of gene expression, either via interaction with promoters, enhancers, transcription factors, and gene bodies, or via stimulating transcriptional elongation and gene splicing [126,127]. With age, genome-wide methylation levels are reported to generally decrease overall across multiple tissues, referred to as hypomethylation [128,129]. This hypomethylation is more pronounced in certain tissues, such as blood and brain [130], particularly in repetitive elements of intergenic regions [131]. Dysregulated DNA methylation through mutations in DNA-methyltransferase family (DNMT3A), isocitrate dehydrogenase 1/2 (IDH1/IDH2), and ten-eleven translocation superfamily 2 (TET2) play an important role in the pathogenesis of a large proportion of MDSs and AMLs, with possible common targeted therapies in the future [110].
DNAm genes and mitochondrial genomics. Cells harboring distinct mitochondrial haplogroups, meaning that they possess identical nuclei with different mtDNA or mtDNA polymorphisms (such as peripheral blood, articular cartilage, and human retinal cell cybrids), have been shown to have different nuclear DNA (nDNA) methylation patterns [132,133,134,135,136]. Moreover, as many metabolites are produced in the mitochondria but transit to the nucleus, alterations in their levels can influence the efficiency of methylation enzymes and affect the production of substrates required for methylation [137,138]. These include but are not restricted to serine biosynthesis, the folate cycle, the methionine cycle, the transsulfuration pathway, and the TCA/Krebs cycle [132,139,140]. Some of the produced metabolites influence the activity of TET enzymes [137]: succinate and fumarate, as seen in multiple human cancer cell lines, are inhibitors of TET enzymes, and in the case of fumarate, it is capable of modulating TET via reducing mRNA expression of TET1 and TET2, though increasing mRNA expression of TET3 enzymes [138]. Additionally, the activity of TET enzymes also changes in response to adenosine monophosphate-activated protein kinase (AMPK)-mediated phosphorylation. AMPK can enhance the expression of TET enzymes directly or via increased IDH2 expression, a mitochondrial enzyme found in the TCA/Krebs cycle involved in the production of α-ketoglutarate (α-KG), and hence activate TETs to decrease DNAm ultimately. Nuclear DNAm can impact the expression level of nuclear-encoded genes, including nuclear-encoded mitochondrial genes, such as genes translated into proteins and enzymes required for mitochondrial transcription and replication, and structural proteins or proteins of the mitochondrial respiratory chain complex [141]. Additionally, alterations in DNAm can also be found in OXPHOS genes associated with various diseases and aging [142].
The methylation status of the mtDNA. For a long time after the 1970s it was a debated whether methylation occurs in the mtDNA or not [143,144,145] Using different methods and cell types, mtDNA methylation was found to range between 1 and 20%. Many authors highlighted that using the current, standard techniques of measuring DNAm, results in very low mtDNA methylation detectability and occurs differently than nDNA methylation. Interestingly, DNMTs and TETs, typically found in the nucleus, have been spotted in the mitochondria, affecting mtDNA methylation [142,146].
Mitochondrial respiratory complex dysfunction and DNAm. Notable data are available from studies, where rotenone-induced complex I dysfunction resulted in global changes in DNAm levels in rats, human cybrid cells, and the change was also seen in the offspring of a mouse model [147,148,149,150]. Together, these studies enforce that mtDNA alterations can be signaled to the nucleus, affecting the nDNA, and are consequently associated with nDNA methylation [132,151].

3.3.2. Histone Acetylation/Deacetylation

Histone. Both DNAm and histone modifications, such as deacetylation, acetylation, and methylation are associated with regulation of chromatin structure and gene expression. Histone deacetylases (HDAC), which remove acetyl groups from the histones, result in more closed chromatin structure, inhibiting or decreasing gene transcription. HDAC enzymes have been found over-expressed in various malignancies, including AML. However, HDAC inhibitors have limited efficacy as single agents in some studies [110,152,153]. Interestingly, the HDAC inhibitor suberoylanilide hydroxamic acid (SAHA, Vorinostat), was found to induce PGC-1α-mediated mitochondrial biogenesis in an in vitro model of cardiac ischemia [154]. The mtDNA, as previously mentioned, lacks histones.

3.3.3. Isocitrate Dehydrogenases (IDH)

IDH. The human genome has five IDH genes that encode three distinct IDH enzymes, whose activities are either NADP+ (IDH1 and IDH2) or NAD+ (IDH3) dependent. In MDS/AML, only IDH1 and 2 have a pathogenetic role. IDH1 is cytoplasmic, whereas IDH2 is a mitochondrial enzyme. The two enzymes’ main biological functions are related to the biosynthesis of essential metabolites of the TCA/Krebs cycle that, in connection with the pentose phosphate pathway, is mainly responsible for the generation of NADPH, the main component of the cellular redox homeostasis [155].
IDH and mitochondria and the pathogenetic role of IDH2 mutation. IDHs contribute to mitochondrial metabolism, although they have an indirect but crucial effect on DNA methylation via influencing enzyme activities through the products of their activity. In mitochondria, the reductive carboxylation of α-ketoglutarate to isocitrate by IDH2 consumes mitochondrial NADPH, with citrate/isocitrate transported to the cytoplasm, where these metabolites can be oxidized to produce cytosolic NADPH. The reversed process can be used to produce mitochondrial NADPH. Mutant IDH2, on the other hand, catalyzes the reductive conversion of α-KG to 2-dihydroxyglutarate with concomitant oxidation of NADPH to NADP+. This oxidative microenvironment plays a crucial role in the altered metabolism of clonal hematopoietic cells. It is suggested that 2-dihydroxyglutarate (2-DHG) could represent the oncogenic mediator in leukaemogenetic processes. α-KG is a cofactor of many deoxygenases involved in regulating key biological processes, such as nucleic acid repair, hypoxic response, chromatin modification, and fatty acid metabolism. In contrast, 2-DHG has the opposite effect as it acts as an inhibitor of these processes. 2-DHG is a competitive inhibitor of histone demethylases, and it has an influence on hypoxia-inducible factor (HIF) prolyl hydroxylase (Figure 5) [41,155].
The mechanism of IDH-induced leukemogenesis. IDH mutants exert their pro-oncogenic effect by interfering with the differentiation program of hematopoietic cells. Cell culture studies showed that the expression of an IDH-mutant enzyme induced an increase in stem cell markers and impaired myeloid cell differentiation [169]. Recently a knock-in mouse model was established, in which the IDH1-R132H mutation was inserted into the murine IDH1 locus and expressed in myeloid cells. These mutant mice displayed an increased number of early hematopoietic cell progenitors, impaired myeloid cell differentiation, anemia, splenomegaly and extramedullary hematopoiesis. The hematopoietic cells of these animals displayed hypermethylated histones and changes to DNAm, similar to those observed in IDH-mutant AMLs [170].
IDH mutations as therapeutic targets. IDH1 and IDH2-mutant mouse and human leukemia models suggest that they are sensitive to all-trans retinoic acid (ATRA) and the proapoptotic effect of arsenic trioxide (ATO) by themselves or in combination with each other [170]. Besides ATRA, IDH mutations can act as therapeutic targets for bcl-2 inhibitors. IDH1 and 2-mutant primary human AML cells were more sensitive than IDH1/IDH2-WT AML cells to ABT-199, a specific BCL-2 inhibitor, by inhibiting cytochrome c oxidase in the mitochondrial ETC [171]. In addition, IDH-inhibitors have been proven to be valuable in models when using IDH-mutant cells by inhibiting the development of the leukemic phenotype, by decreasing the production of oncogenic metabolites responsible for inhibiting cell differentiation and changes in gene expression. These inhibitors include the mutant IDH2 inhibitor AG-221 (enasidenib), and mutant IDH1 inhibitorAG-120 (ivosidenib), which have been extensively investigated for the treatment of patients with AML or MDS with a susceptible IDH mutation [142].
IDH2 inhibition. At the cellular level, the main effect of enasidenib is to induce cell differentiation without apoptosis induction. In addition, it reduces the 2-DHG level markedly. Co-occurrence of mutations at the MAPK and RAS pathways level was associated with reduced clinical response to enasidenib. Furthermore, enasidenib improved outcomes in an in vivo human AML xenograft model [155]. These observations strongly supported the clinical use of enasidenib. Several preclinical trials proved the efficacy of IDH2 inhibition in combination therapy. A study using a mouse model with combined mutation of TET2, FLT3-ITD, and IDH2-R140Q; thus, triple-transformed leukemia, showed tumor sensitivity both to 5-azacytidine and to IDH2 inhibitor enasidenib. The combined treatment with these two drugs resulted in a marked potentiation of the antileukemic effect, with a pronounced decrease of leukemic blasts and with their differentiation and, particularly, with a decrease of mutant allele burden and progressive recovery of normal hematopoiesis from non-mutant stem-progenitor cells [155,172].
IDH2 inhibition in clinical trials. Both monotherapy with enasidenib and combination therapy with enasidenib and 5-azacytidine vs. 5-azacytidine alone showed high response rates in IDH2 mutant AML patients [158,173,174,175]. In addition, enasidenib is a promising treatment option for relapsed/refractory MDS patients harboring IDH2 mutation following allogenic stem cell transplantation, who had overall good response rate and an improved median of survival [176]. Efficacy and tolerability of enasidenib alone and combined with azacytidine were examined in another trial, involving patients with high-risk IDH2-mutated MDS: the best result, which was a 100% response, was achieved in hypomethylating-agent naive patients. Interestingly, the clearance of IDH2 mutation was observed in some patients [177].
IDH2 and mitochondrial dynamics. Besides its already complex role in cellular metabolism, loss of IDH2 was associated with increased mitochondrial motility and fission, the latter via Drp1 activation and expression in in-vitro and in vivo tumor models. These changes resulted in greater tumor cell movements. The exaggerated mitochondrial trafficking was induced by ROS and subsequent hypoxia-inducible factor-1α stabilization [178]. Targeting mitochondrial fission and trafficking could be valuable therapeutic options, just as altered metabolism in tumors with IDH2-mutations.

3.3.4. Ten-Eleven Translocation (TET2) Enzyme

The TET enzyme superfamily. The three enzymes of the TET family (TET1, TET2 and TET3) identified in humans are evolutionarily conserved dioxygenases. Different TET enzymes exhibit distinct expression patterns in vivo, with TET1 being mainly expressed in embryonic stem cells. TET2 and TET3 are more ubiquitous, with TET2 expression predominating in various differentiated tissues, especially in hematopoietic and neuronal lineages [179].
The physiological role of TET2 and its contribution in hematologic malignancies. TET enzymes are one of the homeostatic links between intracellular metabolism and epigenetic gene regulation [180]. TET dioxygenases require α-KG, oxygen and Fe(II) for their activity, which is enhanced in the presence of ascorbic acid [181,182]. As mentioned above, mutant cytosolic IDH1 or mutant mitochondrial IDH2 produce 2-dihydroxyglutarate, an ‘oncometabolite’ that inhibits 2-oxoglutarate-dependent enzymes, including TET dioxygenases. TET enzymes may also be sensitive to changes in oxygen availability and susceptible to reactive oxygen species and carcinogenic metals that displace iron such as arsenic, nickel, or chromium [183].
TET isoforms in mitochondria. Despite unknown mechanisms of translocation to the mitochondria, TET1 and TET2 enzymes have been found to be present in mouse neuronal mitochondria, for example in the cerebellum and Purkinje cells of aged animals [184]. Moreover, the expression of TET2 and TET3, once increased, are associated with increased levels of 5hmC (5-hydroxymethylcytosine) found both in the nDNA and mtDNA [184,185]. TET enzymes could have a role in mtDNA demethylation, with the exact functioning of such enzymes yet to be elucidated. It is still debated whether potential DNMT and TET-modulated mtDNA methylation/demethylation could influence the expression levels of mitochondrial-encoded genes and their respective functionalities or whether this phenomenon loops back into affecting the nDNA.
Animal models. Several studies have examined TET2 inactivation in mice: its deletion leads to hematopoietic defects, including enhanced HSC self-renewal and myeloid expansion, correlating with global loss of 5-hmC in primitive hematopoietic populations [186,187,188].
Factors influencing TET2 function. It was recently described that restoring TET function via inducible shRNA model of TET-induced AML, or through vitamin C administration, the latter being a cofactor for α-KG dependent dioxygenases reverses leukemogenicity induced by the mutant TET protein [189]. These results imply that metabolic control of TET activity could be harnessed for therapeutic benefit in patients with TET mutations. Notably, cytosine methylation signatures of TET2-mutated AML show significant overlaps with those found in IDH1/IDH2 mutated patients. It is to note that IDH1/IDH2 and TET2 mutations are mutually exclusive in AML [169] but signal a common mechanism of leukemogenesis based on aberrant DNA methylation. Besides vitamin C, proteolytic processes and micro-RNA miR22 regulate TET2 function [190,191].
TET2 and hematopoiesis. TET2 has pleiotropic roles in hematopoiesis, including stem-cell self-renewal, lineage commitment, and terminal differentiation of specific lineages. The TET2 gene is highly expressed in HSCs and progenitor cells and is downregulated with differentiation. Several studies on mouse models proved that TET2 acts as a tumor suppressor as well [187,188,192,193].
The role of TET mutations in MDS and AML. TET2 alteration was one of the most prevalent genetic abnormalities (25–35%) identified in MDS [194,195,196,197]. A more extensive series failed to identify a strong association of TET2 alteration with clinical phenotype, risk scores or overall survival [194]. Nevertheless, in higher-risk MDS and AML with low blast count, the TET2 status can be associated with a better response to the demethylating agent azacitidine [196]. The prevalence of TET2 mutations is higher in secondary than in de novo AMLs. In addition, TET2 genetic alterations could be associated with adverse outcomes in cytogenetically defined subgroups of AML patients [198,199,200,201].
TET2 mutations in other hematologic and non-hematologic malignancies. Somatic alterations in TET2, including deletions and missense, nonsense and frameshift mutations, have been identified in 10–26% of MPN patients [202,203]. Acquired somatic alterations in TET2 were identified in 2–20% of classical MPNs, including polycythemia vera, essential thrombocytosis and primary myelofibrosis [197,204,205]. These mutations, which can be an early genetic event in the course of MPN, have no clear prognostic impact; they do not increase the risk of leukemic transformation. In some cases, however, TET2 mutations are only seen when the disease progresses to acute leukemia [206,207]. In chronic myeloid leukemia, TET2 mutations were associated with acute blastic transformation [208]. Additionally, TET2 mutations were identified in 20% of mastocytosis, mostly in aggressive forms of the disease [209,210]. More recently, TET2 mutations have been also found in 30% of patients with blastic plasmacytoid dendritic cell neoplasms [181,211]. TET2 mutations have been also identified in mature B-cell (2%) and T-cell (11.9%) lymphomas [212,213], in 33% of angioimmunoblastic T-cell lymphomas [212], in mantle cell lymphomas [214], and diffuse large B-cell lymphomas. The latter two are also associated with an altered DNA gene methylation pattern on genes involved in hematopoietic development [215]. Interestingly, in angioimmunoblastic T-cell lymphomas, TET2 mutations are frequently associated with DNA methyltransferase 3A mutations [216], and IDH2 mutations [217]. TET2 mutations have also been detected in a small number of solid tumors, such as in prostate cancers [218].
TET2 in AML/MDS therapy. TET2 gene alterations were anticipated to predict an increased response to hypomethylating agents and this predictive effect was explored in several cohorts of high-risk MDS, AML with low blast counts, and severe CMML patients [219,220,221,222,223]. In some of these studies, the TET2 status appeared as an independent genetic predictor of azacitidine response, although not conferring a survival advantage [196,221].

3.3.5. DNMT3 Enzyme

The DNMT enzyme family. The DNMT family, including DNMT1, DNMT3A, and DNMT3B encode methyltransferases that catalyze DNA methylation that involves adding a methyl group to the carbon-5 position of cytosine in CpG dinucleotides, leading to the formation of 5-methylcytosine (5-mC). DNMT3A and DNMT3B are involved in de novo DNA methylation, whereas DNMT1 plays a role in the maintenance of DNA methylation [224]. DNMT3A is highly expressed in T-lymphocytes and neutrophils, while DNMT3B is downregulated in hematopoietic differentiation. Aberrant CpG island promoter methylation in tumor suppressor genes is an important pathogenetic mechanism in malignant tumors, suggesting that DNMTs play important roles in oncogenesis [225].
DNMT isoforms in the mitochondria. DNMT1 has been observed to translocate into the mitochondria and interact with the mtDNA in the matrix of some tissues, such as mouse embryonic fibroblasts, human colon cancer cells [141], and human neurons [226]. The presence of DNMT3A has also been described in mitochondria. Interestingly, a particular isoform, the DMNT1-3A, was the one that has been identified to be able to both transfer to mitochondria and methylate the mtDNA [227,228,229].
DNMT3A and leukemogenesis. The mechanism of leukemogenesis by DNMT3A is not entirely clear; however, studies have shown that heterozygous DNMT3A ablation in mice leads to an expansion of the hematopoietic stem cell pool [230]. The formation of myeloid malignancies, however, may require additional genetic alterations. This reinforces the notion that DNMT3A mutation, just as mutations in other epigenetic regulators, do not necessarily lead to frank leukemic transformation on their own but rather create a premalignant state that lays the ground for malignancy. It has recently been reported that mutant DNMT3A (R882H) interacts with the Polycomb repressive complex 1 (PRC1) in order to silence genes, suggesting that PRC1 activity could be an attractive target in DNMT3A-mutant tumors [231].
The role of DNMT3A mutations in MDS and AML. DNMT3A mutations occur in 30–35% of AMLs with normal karyotype, 10% of MDSs, and 20% of T-lineage acute lymphoblastic leukemia [232,233,234]. As mentioned above, DNMT3A mutations result in loss of function, and can be present in pre-leukemic hematopoietic stem cells, which can remain in the tumor cells after transformation to MDS or AML [235,236]. DNMT3A mutations are associated with poor prognosis and decreased overall survival [237]. MDS patients with DNMT3A mutations have a shorter OS and higher risks of leukemic transformation [238,239]. Additionally, DNMT3A mutations have been observed in non-leukemic T-cells from AML patients as well and in normal elderly individuals with no signs of leukemia [240].
DNMT3a in MDS/AML therapy. DNMT3A mutations are associated with a positive response to DNA methyltransferase inhibitors (a.k.a. DNA demethylating agents), namely azacitidine, decitabine and guadecitabine [233,241]. The effect of histone deacethylase (HDAC) inhibitors, such as pracinostat, vorinostat and valproic acid, have been modest in AML. In combination with DNA demethylating agents, however, HDAC inhibitors show increased activity [152].

3.3.6. Additional Sex Like Comb Protein 1 (ASXL1)

ASXL superfamily. The ASXL superfamily of genes and the encoded proteins consist of three members, namely ASXL1, ASXL2 and ASXL3. They are human homologs of the Drosophila Asx gene, encoding epigenetic scaffolding proteins that are involved in the regulation or recruitment of the polycomb-group repressor complex (PRC) and trithorax-group (trxG) activator complex, participating in epigenetic regulation and histone modification [242]. ASXL1 directly interacts with various protein-coding genes, such as BAP1, KDM1A (LSD1), NCOA1, and nuclear hormone receptors, such as retinoic acid receptors, estrogen- and androgen receptors. The loss of the ASXL1 gene is associated with leukemoid transformation and increased self-renewal in hematopoietic cells [242].
The role of ASXL1 in mitochondrial and endoplasmic reticulum function. A recent publication reported a significant mitochondria–endoplasmic reticulum deficiency in a leukemia cell line carrying ASXL1 mutation. This deficiency was related to decreased number of matrix granules, mitochondria-associated endoplasmic reticulum membrane (MAM), and mitochondrial-derived vesicle (MDV) precursors, which are implicated in the regulation of cell death pathways via mitophagy and intracellular Ca2+ concentration changes. These cells are also thought to have increased mitochondrial respiration and defective mitophagy [243].
ASXL1 mutations in hematologic malignancies. Nonsense point mutations or frame-shift mutations of ASXL1 occur in hematological malignancies, such as MDS, MPNs, MDS/MPNs, AMLs and CLL [244,245,246]. Whole-exome sequencing analyses revealed a 2.9% ASXL1 is mutation rate in CLL [246], with generally higher percentages in myeloid neoplasms (45.3% in CMML, 34.5% in MPN, 30% in secondary AML, 16.2% in MDS, and 6.5% in de novo AMLs) [247]. In MDS, ASXL1 mutations are independent adverse prognostic factors both in overall and leukemia-free survival. In addition, ASXL1 mutations are associated with shorter overall survival in CMML patients [194,220,247,248,249]. In AML, ASXL1 mutations were more frequently found in male [250,251,252,253,254], and older age patients [255], and in patients with lower platelet count and hemoglobin level. Of note, ASXL1 mutations are frequently seen with other gene alterations, such as with EZH2 IDH1/2, RUNX1, and TET2 [194,256], most of which are adverse prognostic factors themselves in myeloid neoplasms, potentially explaining why ASXL1 mutations are associated with poor prognosis in many cases [257].
ASXL1 in MDS and AML therapy. Leukemic cells with ASXL1 mutation have been shown to overexpress anti-apoptotic Bcl-2 and have increased global cytosine methylation levels. It is not surprising that these cells were sensitive to both the Bcl-2 inhibitor venetoclax and the DNMT-inhibitor azacytidine [258]. In addition, clinical trials have shown the efficacy of venetoclax combined with hypomethylating agents or low-dose cytarabine in a subgroup of AML patients [259].

3.3.7. Enhancer of Zeste Homolog 2 (EZH2)

EZH2 and carcinogenesis. Enhancer of zeste homolog 2 (EZH2) is an epigenetic modulator, part of the polycomb repressive complex 2 (PRC2) that can suppress gene expression via histone 3 (H3K27me3) di- or trimethylation. In addition, EZH2 methylates non-histone proteins, such as the transcription factor GATA4, and can activate downstream genes in a PRC2-independent manner, contributing to its complex effect on cells. Mutation or altered expression of EZH2 has been linked to various tumors, including hematologic malignancies and solid tumors, such as breast cancer, esophageal cancer, gastric cancer, and anaplastic thyroid carcinoma. Interestingly, both increased and decreased EZH2/PRC2 functions are associated with carcinogenesis. With increased function, EZH2 silences genes that promote differentiation and restrain proliferation. As a tumor suppressor, loss of EZH2 accelerates Ras-driven neoplastic processes and can amplify Akt and ERK activation. In JAK2-V617F transgene mice with concurrent EZH2 knockout, a synergistic effect with very high platelet and neutrophil count, accelerated myelofibrosis, and reduced survival was described. Additionally, PRC2 is suppressed in T-cell acute lymphoblastic leukemia (T-ALL) by NOTCH1 signaling, resulting in loss of H3K27me3 repression as part of leukemogenesis. In CML, EZH2 overexpression is induced by BCR-ABL1 via signal transducer and activator of transcription (STAT) 5 phosphorylation [260,261,262,263,264,265].
EZH2 in malignancies. The type of EZH2 alteration is different in different tumor types. Solid tumors often overexpress EZH2, whereas hematologic neoplasms show a more diverse picture of how EZH2 is altered. Gain-of-function mutations, promoting H3K27me3 and subsequent gene suppression, have been described in germinal center-type diffuse large B-cell lymphomas (30%) and follicular lymphomas (20%). Gain of copy-number, as part of gains of chromosome 7 (including the 7q36.1 region), has also been described in follicular lymphomas, resulting in the same overall effect as the gain-of-function mutations in these patients. In patients with various B-cell lymphomas, high-risk MDS, and AML, overexpression of EZH2 is the most common. Interestingly, loss-of-function mutations have also been described in MDS (6%), MDS/MPN (10–12%), and myelofibrosis (13%). [117,263,264,266]. Loss of chromosome 7, or deletion of 7q (7q-), which are frequently seen in AML and MDS patients, also involves EZH2 [264]. In general, EZH2 mutations are associated with inferior overall survival in myeloid malignancies, with no increased rate of AML transformation in MDS patients [246,267]. In addition, decreased EZH2 expression is associated with worse overall survival in AML [268]. Also, increased EZH2 expression was found in AML patients with complete remission when compared to newly diagnosed patients, who had no EZH2 mutations in either of the mutated hotspots [269]. Loss of EZH2 function can either attenuate and promote leukemic transformation in MDS and MPN, depending on the disease context and cooperating mutations [268].
EZH2 and mitochondria. EZH2 alters apoptotic pathways, with no data on its effect on mitochondrial dynamics. In glioma cells, its downregulation results in apoptosis and cell cycle arrest in the G0/G1 phase via cytochrome c release [270]. Similarly, in multiple myeloma cells, inhibition of EZH2 showed caspase-3-dependent apoptosis [271]. On the other hand, in EZH2 overexpressing melanoma cells, the EZH2 inhibitor GSK126 caused caspase-independent apoptosis, mediated by apoptosis-inducing factor, mitochondrion associated 1 (AIFM1) protein [272]. In contrast, EZH2 inhibition enhanced cell proliferation and reduced apoptosis in AML cells, in line with myeloid cell leukemogenesis and chemotherapy resistance with reduced EZH2 expression or function [269]. Nevertheless, EZH2 inhibition reduced glycolysis, promoted OXPHOS in pancreatic cells in vitro [273], and inhibited glycolysis in prostate cancer cells [274]. Moreover, EZH2 regulates lipid metabolism through the EZH2-TERT-lipid metabolism network, with EZH2 knockdown glioblastoma cells showing reduced fatty acid synthase expression and resultant diminished fatty acid levels [263,275].
EZH2 as targeted therapy. Patients with a gain of function mutations, or increased copy-number have been proposed to be ideal candidates for EZH2 inhibitor treatment. EZH2 inhibitors have been tested in clinical trials in various malignancies, including myeloid and lymphoid neoplasms, listed in more detail in the review of Duan and co-workers [263,266], with the existing pre-clinical data supportive of treatment in these diseases [276,277].

3.4. RNA Splicing Gene Mutations

3.4.1. RNA Splicing

RNA splicing or alternative splicing is where mature RNA is formed from pre-messenger RNA (pre-mRNA) through intron removal and exon splicing. Alternative splicing increases the transcriptomic and proteomic complexity by generating distinct RNA isoforms with different functions than the other products from the same gene. Alternative splicing can contribute to carcinogenesis when dysregulated [278,279]. Mis-splicing of genes has been found in multiple malignancies, including some carcinomas, neuroblastoma, CLL and AML. Also, mutations in genes involved in RNA splicing, such as SF3B1, SRSF2, ZRSR2 or U2AF1/2, occur in approximately 50% of MDSs and are currently intensively examined for their therapeutic relevance [280,281,282,283].

3.4.2. Alternative Splicing

Alternative splicing is also regulated by epigenetic changes [284,285]. Histone modifications and DNA methylation can affect exon usage by controlling the elongation speed of RNA polymerase II and, consequently, the choice of splice sites. Moreover, chromatin modifications have been found to regulate the activity of alternative or cryptic transcriptional start sites (TSSs) in the genome [286]. Indeed, deregulation of alternative promoter usage has been recently identified as a common phenomenon in cancer [287]. It is, however, largely unknown how genetic changes can result in alternative promoter choices. In addition, their role in mitochondrial metabolism and epigenetic regulation is still unknown. Recently many RNA splicing factor mutations were detected as a possible contributor in MDS and AML pathogenesis. Until now, its role as a therapeutic target remained unclear in clinical settings, but preliminary data from experiments on cell lines showed promising results [225].

3.4.3. SF3B1

SF3B1 encodes a subunit of the splicing factor 3b complex. SF3B1 disrupts the usage of thousands of splice junctions, leading to altered expression of hundreds of genes [225]. SF3B1 mutations are seen in 57–75% of patients with refractory anemia with ring sideroblasts (RARS) and 6–18% of patients with MDSs without ring sideroblasts [288,289].
SF3B1 in mitochondrial metabolism. The mutated SF3B1 downregulates genes essential to various mitochondrial pathways, including ACACA (acetylcoenzyme A carboxylase α) and RGL1 (ral guanine nucleotide dissociation stimulator like-1) genes. SF3B1-mutated RARS have abnormal splicing of the ABCB7 gene in the mitochondria, which leads to deficiency of the ABCB7 protein, resulting in mitochondrial iron overload, reduced heme synthesis, and ineffective erythropoiesis [283,290]. SF3B1-mutated MDS is associated with thrombocytosis, increased ring sideroblasts, fewer cytopenias, lower blasts percentage, and are associated with a favorable prognosis [289].
Metabolic effect of SF3B1 mutation. A recent in vitro study showed that SF3B1 mutated cells expressed less mitochondrial complex III gene, resulting in decreased cellular respiration and reduced citric acid cycle metabolites. In addition, further misspliced and downregulated mitochondrial metabolic enzymes, such as dihydrolipoamide S-succinyltransferase (DLST) and methylmalonyl-CoA mutase (MUT), were noted. This metabolic reprogramming bears a particular relevance in MDSs with ring sideroblasts (MDS-RS) with SF3B1 mutation [291].

3.4.4. SRSF2

SRSF2, encoding the serine/arginine-rich splicing factor 2, is critically involved in splice site selection, spliceosome assembly, and constitutive and alternative splicing [291]. SRSF2 mutations are stable during disease evolution in MDS, suggesting that they may play a role in disease initiation. SRSF2 mutations are seen in 11–15% of patients with MDS, frequently co-existing with RUNX1, IDH1, IDH2, and ASXL1 mutations [283], and confer an inferior overall survival [292,293].
SRSF2 and mitochondria. In mammalian cells, nutrients and growth factors signal through an array of upstream proteins to regulate the mTORC1 growth control pathway. SRSF2 gene function impacts the mTORC1 pathway, and mitochondrial stress influences mTORC1 signaling as well [294].

3.4.5. U2AF1

The U2AF1 gene encodes the U2 auxiliary factor, facilitating the binding of U2 snRNP to the pre-mRNA branch site. Recurrent mutations of the U2AF1 gene occur in 9% of patients with MDS. The prognostic impact of U2AF1 mutations and the exact mechanism of U2AF1S34F mutation confers a clonal growth advantage in MDS remain unclear [283,295]. FOXO3a activation is likely involved by increasing oxidative stress, which subsequently alters multiple processes, including apoptosis, autophagy, and immune-inflammatory responses [296].

3.5. Transcription Factor Mutations

The main effects of the transcription factors discussed are summarized in Table 2.

3.5.1. Runt-Related Transcription Factor 1 (RUNX1)

All RUNX proteins contain the runt-homology domain (RHD), which is responsible for DNA-binding and interaction with a common heterodimeric partner, CBFb [297]. The fusion oncogene RUNX1/RUNX1T1 is expressed due to the chromosomal translocation t(8; 21), which involves the RUNX1 gene on chromosome 21 and the RUNX1T1 gene on chromosome 8 [298]. When expressed in hematopoietic cells, the fusion protein occupies more than 4000 genomic sites and forms transcription regulatory complexes by recruiting cofactors [299,300,301,302,303,304]. These complexes trigger local chromatin remodeling of a wide range of genes and thereby affect their expression [299,305,306]. The change of target gene expression leads to a block of cell differentiation, increased self-renewal, inhibition of apoptosis, which eventually results in malignant transformation [301,305,307,308]. The details of all these events, however, are not entirely understood.
RUNX1 in mitochondria. In a mouse model with both RUNX1 mutation and CBL exon deletion increased Drp1-dependent mitochondrial fission and ROS production in the tumor cells, leading to impaired granulopoiesis dysplasia and an overall MDS phenotype. These changes were reversible with Drp1 inhibition, making it a promising therapeutic candidate in MDS patients harboring RUNX1 mutation/RUNX1/RUNX1T1 fusion gene [120].
RUNX1 in AML.RUNX1 and CBFB are frequent targets of chromosome abnormalities in human AML and ALL. Core inding factor (CBF) leukemia subtypes, which includes leukemias harboring fusion genes CBFB/MYH11 or RUNX1/RUNX1T1, are associated with younger age and range from 20% in pediatric to less than 5% in older AML patients [309]. Moreover, CBF leukemias are generally associated with a relatively good prognosis. Somatic mutations in RUNX1 are detected in approximately 3% of pediatric and 15% of adult de novo AML patients. In cytogenetically normal cases, the presence of RUNX1 mutations is associated with poor prognosis [310,311,312].
RUNX1 in MDS. Several studies reported somatic mutations in RUNX1 in patients with primary MDS, therapy-related MDS (t-MDS), and AML from MDS progression [313,314,315,316,317,318,319,320]. RUNX1 mutations also occur in 20% of Fanconi anemia and 64% of congenital neutropenia (CN) patients, who later develop MDS [310,321]. RUNX1 mutations in MDS are distributed throughout the gene, affecting both major functional domains. RUNX1 is one of the most frequently mutated genes in MDS, accounting for roughly 10% of the cases [320,322]. In primary MDS cases, there is a positive correlation between RUNX1 mutations and shorter survival [323].

3.5.2. Tumor Protein 53 (TP53)

TP53 and apoptosis inhibition, mutant p53. TP53 is a tumor suppressor gene encoding the tumor suppressor protein p53. When p53 forms a homotetrameric transcription factor, it directly regulates about 500 target genes, resulting in various changes, including cell cycle arrest and cell senescence, DNA repair, metabolic adaptation, and apoptosis. Of note, there is significant cross-talk between TP53 and MYC. Additionally, p53 induces apoptosis and autophagy in a transcription-independent way, interacting with other proteins in the cytoplasm [324]. The induction of apoptosis by p53 involves the transcriptional activation of the pro-apoptotic proteins PUMA, and-to a lesser extent-NOXA. PUMA and NOXA inhibit the anti-apoptotic bcl-2 and bcl-Xl proteins, leading to the release of bax/bak inhibition and cytochrome-c release from the mitochondria [325]. In the absence of functional p53, the cells enter to S-phase [326]. In addition, the loss of p53 stabilizes topoisomerase IIα on the DNA and causes fork retardation, leading to torsional stress between the two machines [326,327]. TP53 is an unusual tumor suppressor. In contrast to other tumor suppressors, such as RB and PTEN, where protein expression is significantly deceased or lost, mutant p53 expression can be high compared to the wild type (WT) p53 [325]. The mutant p53 can participate in tumorigenesis in three different ways: 1. Loss of WT p53 activity 2. Dominant negative effect of the mutant form over the WT p53, via formation of unfunctional mixed tetramers 3. Gain of function mutations, where the mutant p53 affects other transcription factors and tumors suppressors (such as p63, and p73) [325].
TP53 and metabolism. Although inhibiting apoptosis is an important part of TP53-related carcinogenesis, it also greatly impacts cellular metabolism and mitochondrial dynamics. This is supported by the in vivo mouse model, where the animals lacking the critical effectors of p53-apoptosis do not develop tumors spontaneously, as the TP53-deficient mice [325,328]. The transcription of glucose transporters GLUT1, GLUT3, and GLUT4 are downregulated by p53. In addition, p53 also inhibits GLUT1 translocations, and downregulates the expression of several glycolytic enzymes, such as hexokinase 2, and phosphoglycerate mutase 1. Furthermore, p53 increases parkin expression, which initiates the degradation of HIF-1α, the latter of which promotes glycolysis under normal conditions. Additionally, the p53-inducible gene TP53-induced glycolysis and apoptosis regulator (TIGAR) lowers the levels of reactive oxygen species (ROS) and fructose-2,6-biphosphate, thereby further inhibiting glycolysis. P53 further represses glycolysis by the transcriptional inhibition of PFKFB3 and PFKFB4 genes, which reduce fructose-2,6-bisphosphate expression. In addition, p53 inhibits the pentose phosphate pathway (PPP) by direct binding to glucose-6-phosphate dehydrogenase (G6PD), and it also inhibits the expression membrane-bound lactate transporter. OXPHOS is induced by p53, which represses the transcription of pyruvate dehydrogenase kinase 2 (PDK2), a negative regulator of pyruvate dehydrogenase (PDH) that converts pyruvate to acetyl-CoA, a primary substrate in the TCA cycle. OXPHOS is enhanced by p53 via the activation of mitochondrial glutaminase 2 (GLS2), catalyzing the hydrolysis of glutamine to glutamate. Additionally, p53 suppresses fatty acid synthesis and enhances lipid synthesis, unfavorable for tumor cells. Interestingly, TP53 with a gain-of-function mutation promotes glycolysis [329,330,331]. The widespread effect of p53 on cellular metabolism explains why tumors harboring an altered TP53 are more aggressive. These effects also represent possible therapeutic targets.
TP53 and mitochondrial dynamics. Supporting cellular senescence, p53 promotes the formation of highly interconnected mitochondria by blocking fission protein Drp1 translocation to the mitochondria. This translocation inhibition can occur via inhibitory phosphorylation of Drp1 at the Ser637 site. This inhibitory action is part of its tumor suppressor effect. P53-dependent Protein kinase A (PKA) activation has also been linked to mitochondrial elongation [332,333]. On the contrary, a recent study showed that mitochondrial fusion protein OPA1 is cleaved by p53 via another inner mitochondrial membrane protein, OMA1, resulting in bax/bak-dependent apoptosis induction. This pathway is suppressed in normal cells, where different stressors can activate it. OMA1/OPA1 expression is, however, variable in different solid tumors, and their exact role in cancer needs to be further investigated [334]. In patient-derived T-ALL xenografts, increased OPA1-cleavage and resultant mitochondrial fission and cell death could be triggered by ROS-generating therapy [335].
TP53 in tumors.TP53 mutations have long been described in the literature along with Li-Fraumeni syndrome, where the patient carries a heterozygous mutation of the TP53 genes, resulting in the development of multiple malignancies [336,337]. Most of the mutations are point mutations in the DNA-binding domain, resulting in the loss of function of the gene. Gain of function mutations, and dominant negative effect, however, have also been reported. Regarding the dominant negative effect, numerous mutant p53 would form mixed tetramers with the WT p53 in a patient with retained TP53, resulting in impaired function. These mutations depend on the tumor cell type, with many malignancies harboring various possible TP53 mutations [325].
TP53 in hematologic malignancies. In MDS patients, TP53 mutations, especially mutli-hit TP53 mutations, are associated with high-risk disease, rapid transformation to AML, and therapy resistance. On the other hand, patients with monoallelic TP53 mutations had a higher incidence of other genetic mutations, including genes, such as TET2, SF3B1, ASXL1, RUNX1, SRSF2, JAK2, BCOR and CBL [338]. In AML, TP53 mutations occur in about 5–10% of patients, with a higher incidence in the t-AMLs, and are associated with therapy resistance and short overall survival [339,340]. In addition, AML patients with TP53-aneuploidy showed inferior outcomes [341]. Furthermore, TP53 mutations are frequently seen in lymphoid malignancies, such as in DLBCL (20–25%) and mantle cell lymphoma (25%) [342,343]. Additionally, patients with secondary DLBCL harboring TP53 mutations and treated with R-CHOP had inferior outcomes [342,343]. In B-cell lymphomas, altered TP53 and MYC expressions can co-exist, with a significant cross-talk between the two pathways, accelerating each other’s effect. Given that both TP53 and MYC alterations are associated with worse outcomes, their co-existence results in particularly poor prognosis [324]. In multiple myeloma, TP53 alterations include monoallelic deletion as part of deletion of chromosome 17p (del17p) (~8%), monoallelic mutations (~6%), and biallelic inactivation (~4%), associated with high-risk disease [344].

3.6. Signaling Molecules

RAS

RAS proto-oncogenes, such as HRAS, NRAS, and KRAS, encode small GTPases and are frequently seen in various tumors. They exist in GTP-bound active, and GDP-bound inactive forms. When activated, they modulate a wide range of gene transcription. Oncogenic RAS mutations commonly result in amino acid substitutions, locking Ras proteins in a GTP-bound and constitutively active state [345]. Ras overexpression can cause a so-called “replication stress”, which is linked to carcinogenesis similarly to myc. Ras-induced accelerated transcription is driven by transcription factors, such as TBP, which are required for all promoters. On the other hand, c-Myc amplifies transcription indirectly by up- or downregulating target genes that enable RNA production [326]. Interestingly, Ras and Myc have a functionally connected network, and Ras pathways can regulate Myc [345,346]. The major downstream Ras effector pathways include the phosphatidylinositiol-3-kinase (PI3K), mitogen-activated protein kinase (MAPK), and Ras-like (Ral) small GPTase signaling pathways [346].
RAS-induced metabolic changes. Ras and Myc similarly suppress OXPHOS and enhance glycolysis, described in various tumors [13]. Ras, however, reprograms cellular metabolism more robustly than Myc [345]. The signal transducer and activator of transcription 3 (STAT3) is one of the key transcription factors regulating glycolysis and OXPHOS in a Ras-dependent manner [347]. In addition, both Ras and Myc are associated with increased ROS production [326].
RAS-induced changes in mitochondrial dynamics. Increased Ras expression and Erk activation leads to enhanced Drp-1-dependent mitochondrial fission, which was necessary for Ras-induced carcinogenesis in preclinical models [3,80]. Interestingly, T-ALL cells show Erk/Drp1-dependent mitochondrial fission in vitro, triggered by surrounding mesenchymal stem cells [348].
RAS in hematologic malignancies. Both AML and MDS frequently harbor RAS mutations, with NRAS predominating [346,349]. Although NRAS mutations are frequently associated with AML progression from MDS [350], several large cohort studies indicated that NRAS mutations in primary AML did not influence the prognosis of patients. However, a co-existent USAF1 mutation in AML with RAS mutation, however, was related to chemotherapy resistance [351]. Targeting downstream pathways of accelerated Ras expression have been used in various clinical trials with variable outcome. Targeting Ras-like (Ral) GTPases, which are critical mediators of Ras-driven transformation in AML, is a further therapeutic option [346], along with targeting cellular metabolism and mitochondrial fission.

3.7. Myeloproliferative Neoplasms (MPNs)

As the name suggests, myeloproliferative neoplasms (MPNs) are characterized by overproduction of a specific cell line, such as white blood cells, red blood cells, and platelets. The quintessential disorder in this category of hematologic disorders is chronic myeloid leukemia (CML). However, it also includes polycythemia vera (PV), essential thrombocythemia (ET), primary myelofibrosis (PMF), chronic neutrophilic leukemia (CNL), chronic eosinophilic leukemia (not otherwise specified) (CEL), and other unclassifiable myeloproliferative neoplasms. The chromosomal abnormality associated with CML is t(9;22), resulting in the fusion of BCR-ABL1 genes, also termed the Philadelphia chromosome. Although chromosomal abnormalities do not characterize most myeloproliferative disorders, other potential abnormalities include del(9p), del(13q), del(20q), gains of chromosome 8 or 9; and those more suggestive of primary myelofibrosis include der(6)t(1;6)(q21-23;p21.3). In addition, the presence of various genetic mutations are common, and include mutations in Janus kinase 2 (JAK2), and less frequently calreticulin (CALR) or thrombopoietin receptor (myeloproliferative leukemia virus gene, MPL). The most prevalent alteration in JAK2 is a phenylalanine replacement of valine at position 617, commonly referred to as JAK2V617F. Other mutations in exon 12-15 of JAK2 also occur to a lesser extent [6].

3.7.1. BCR/ABL1, JAK2, CALR and MPL

The most common genetic alterations described above, including BCR/ABL, JAK2, MPL, and CALR, exert their effects through alterations in tyrosine kinase signaling. The translocation of the Abelson 1 (ABL1) oncogene on chromosome 9 to the Breakpoint Cluster Region (BCR) on chromosome 22 results in the fused BCR-ABL1 gene. The following protein product induces neoplastic proliferation through deregulated tyrosine kinase activity. The combined BCR-ABL1 product loses its regulatory activity and is thus constitutively active. The unaltered ABL1 protein plays a central role in numerous processes, including cellular survival, proliferation, migration, and stress response. Therefore, loss of its regulation leads to enhanced and persistent signaling through the PI3K/AKT/mTOR, RAS/RAF/MEK/ERK, and JAK2/STAT pathways [352,353]. Likewise, alterations in the JAK2, CALR, and MPL genes lead to transformed tyrosine kinase activity and enhancement of similar signaling pathways. The most common mutation, JAK2V617F, results in loss of the JAK2 inhibitory domain, allowing for persistent downstream activation. The MPL product is the thrombopoietin receptor, which normally activates the JAK2 pathway when bound by thrombopoietin. Mutations in MPL also result in loss of auto-inhibition and thus autoactivation of the JAK2/STAT signaling pathways [354]. On the other hand, CALR functions as a chaperone protein to control calcium equilibrium and protein folding. Mutations in CALR were more recently discovered and are thought to exert oncogenic effects through its altered C terminus. The altered CALR can subsequently bind with MPL intracellularly before being trafficked to the cell surface where it can constitutively activate the thrombopoietin receptor without ligand binding [355].

3.7.2. Metabolic Changes in MPNs with Altered JAK2

Numerous biological derangements are associated with the previously described mutations commonly detected in MPNs. The subsequent metabolic alterations are closely linked to mitochondrial biogenesis and include modifications in apoptosis, glycolysis, fatty acid metabolism, glutaminolysis, and subsequently in OXPHOS [356,357,358]. Ligand independent activation of the JAK/STAT pathway is a common oncogenic modality amongst MPNs. Thus, the underlying metabolic changes induced by this activation are essential for disease progression, neoplastic cell survival and proliferation. Glucose usage is dramatically increased amongst JAK2V617F expressing cells. This upregulation of metabolic processes appears to result from enhanced glucose transportation into the cell via the GLUT1 transporter and increased activity of the glycolytic rate-limiting enzyme, 6-phosphofructo-1-kinase. The enhanced activity of 6-phosphofructo-1-kinase depends on 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3 (PFKFB3), which is controlled by the JAK2V617F kinase. These metabolic changes enhance energy production through glycolysis and the subsequent shunting into OXPHOS within the mitochondria [358]. Not only are the metabolic processes of the neoplastic cells altered, but they also induce changes in neighboring cells by inducing stromal cells to release amino acids, lipids, and other molecules that can be used for energy production. This process contributes to disease progression and can ultimately result in cachexia through starvation of healthy cells. MPN cells also demonstrate increased lipid metabolism and pentose phosphate pathway due to hypoglycemia and inflammatory cytokines. Furthermore, glutamine metabolism is ramped up through increased glutamine uptake from surrounding cells to meet high energy demands [356].

3.7.3. MPN Treatment

The treatment of many MPNs is often based on cytoreductive therapy (hydroxyurea) and the prevention of secondary complications (thrombosis) with low-dose aspirin or phlebotomy for PV. Bone marrow transplantation is also available to some patients with MPN, such as high-risk PMF. However, Janus kinase inhibitors, such as ruxolitinib and fedratinib, are also available to patients who fail or cannot tolerate initial therapy. Although rather efficacious in treatment, Janus kinase inhibitors often do not entirely eliminate abnormal myeloid progenitors, contributing to disease relapse. Therefore, these medications are marred by potential full relapse when discontinued. Current investigations have shown that deubiquitinase (DUB) inhibitors (WP1130 and G9) are potentially helpful treatment modalities to induce apoptosis in JAK2V617F mutants. DUB inhibitors can exert their effect by preventing the deubiquitination of JAK2 so that altered JAK2 molecules are passed through the ubiquitin/proteasome pathway. Additionally, DUBs are important in maintaining mitochondrial quality control, and thus their inhibition leads to amplified ROS production Moreover, DUB inhibitor WP1130 induces the activity of Bak and, to a lesser degree, Bax. Consequently, the induction of ROS and Bak/Bax leads to activation of the mitochondrial-dependent apoptotic pathway [359]. Patients with CALR mutations are managed similarly to most other MPNs. However, the trafficking of mutant CALR to the cell surface provides a viable potential treatment modality with either mutant CALR-binding antibody or CALR targeting by T-cell therapy [355]. Metabolism targeting drugs have also shown promise in MPN treatment, such as blocking0 of the glycolytic enzyme PFKB3 by PFK15 [356].

3.7.4. Metabolic Changes Related to BCR/ABL1

Fusion of BCR/ABL1 results in similar features to those described with JAK2, and it is thought to be due to rampant cellular signaling through multiple pathways due to the kinase activity. BCR-ABL is a vital anti-apoptotic factor by activation of multiple pathways, including Ras, PI3K/AKT, and JAK/STAT pathway, in addition to Myc expression. The PI3K/AKT pathway leads to the inactivation of pro-apoptotic factors by phosphorylation, including Bcl-2 associated death promoter (Bad), forkhead box O transcription factors (FOXO), procaspase 9, and Yes-associated protein (YAP). Similarly, Ras pathway activation induces the PI3K/AKT pathway. Activation of the JAK/STAT pathway promotes cell survival by regulating Bcl-XL and Bcl-2, critical anti-apoptotic proteins [360]. Interestingly, evidence has also shown that BCR/ABL translocation can simulate hypoxic-like signaling even in the presence of oxygen. This results in the activation of hypoxia-inducible factors 1 and 2 (HIF1, HIF2), STAT5, and the glucose transporter genes SLC2A1 and SLC2A3, independent of the environmental oxygen levels. STAT5 and HIF2 are important regulators of glucose uptake and utilization via the SLC2A1/3 transporters. Along the same lines, hexokinase expression is elevated, corresponding with high pyruvate levels for the TCA cycle. Furthermore, HIF1 enhances pyruvate dehydrogenase kinase (PDK2/4) so that the elevation in pyruvate can be quickly used in the TCA cycle. In addition, multiple amino acid levels were elevated compared to normal to provide additional substrates to the TCA cycle. This was partially completed through the use of the SLC1A5 glutamine transporter. Metabolism of glutamine can produce glutamate, which can further be processed into α-ketoglutarate and directly shunted into the TCA cycle within mitochondria. Glutamine appears to be an important factor for intensified OXPHOS, as treatment of BCR/ABL positive CML cell lines with a glutaminase inhibitor drastically hampered OXPHOS [361].

3.7.5. CML Treatment

Due to the underlying pathogenesis of BCR/ABL fusion, treatment is often via tyrosine kinase inhibitors (TKIs), such as imatinib. Subsequent second and third-generation TKIs have also been produced in an attempt to overcome TKI resistance. However, despite their effectiveness, some CML patients develop resistance to them, possibly by altered apoptotic pathways, such as by upregulated anti-apoptotic, and pro-survival molecules, such as SRC kinases, FOXO1, XPO1, and STAT3, or by mitochondrial protein changes [352,357]. The TKI resistance is furthermore thought to occur through accumulated mtDNA damage from increased ROS production. mtDNA damage may accumulate in genes required for OXPHOS, apoptosis, and more, leading to reduced effectiveness of these pathways or even heightened production of ROS, which further damages the mtDNA. In addition, as described previously, mitochondria of CML cells have increased uptake and utilization of glucose through glycolysis and oxidative phosphorylation, which also lends a hand to the overproduction of ROS. Considering the high level of ROS detected in CML cells, oxidative damage may also play a role in a TKI resistance. Therefore, targeting mitochondrial metabolism may prove helpful in the future for TKI resistant CML patients [357].

3.8. Lymphomas, Acute and Chronic Lymphoid Leukemias

3.8.1. Lymphomas

Lymphomas are a diverse group of lymphoid malignancies with heterogeneous molecular, chromosomal, and epigenetic changes. Lymphomas are typically divided into Hodgkin (HL) (10%) and non-Hodgkin lymphomas (90%), with the latter being either B-cell (90%), T-cell or natural killer cell (NK cell) (10%) types [362].

3.8.2. Hodgkin Lymphomas (HLs)

In classical HL (cHL) (95%), there is only a small percentage of neoplastic Reed-Sternberg (RS) and Hodgkin cells intermixed with inflammatory cells. On the other hand, in the nodular lymphocyte-predominant Hodgkin lymphoma (NLP-HL) (5%), the neoplastic cells are different and are designated as lymphocyte-predominant cells, a.k.a. “popcorn” cells. The mainstream treatment of HL is typically with ABVD (doxorubicin, bleomycin, vinblastine, and dacarbazine) and irradiation, where the chemotherapy has been described to primarily increase oxidative stress [1,363,364]. In relapsing/refractory (r/r) HL, about 10% of HL cases, autologous stem cell transplantation (ASCT) has about a 50% success rate. Several HL patients, however, are not eligible for ASCT or can have a relapse post-ASCT [365]. Newer drugs, such as anti-CD30 antibody-drug conjugates and anti-PD1 checkpoint inhibitors, can be used in these patients, with some non-responsive to these treatments [365]. However, additional therapeutic options for these patients are still in need, which include drugs targeting mitochondria. A previous study showed that the cHL cells have largely increased OXPHOS with increased mitochondrial mass and biogenesis and low levels of lactate production, which is unusual for other types of cancer cells. In addition, increased mitophagy, mitochondrial turnover, and mitochondrial anti-oxidant capacity have been reported in HL, contributing to drug resistance [1,366].

3.8.3. HL and Genetic Abnormalities

Genetic analysis in cHL is challenging due to the presence of less than 5% tumor cells in the tissue. A recent study, however, analyzed the tumor cells after microdissection, and found that a large proportion (87%) of the cells harbored mutations in the JAK-STAT signaling pathway, including genes STAT3, STAT5B, JAK1, JAK2, and PTN1 [367]. Furthermore, about 10% of cHL patients had TP53 point mutations/deletions, and about 60% had gains of MDM2, latter being a negative regulator of p53 [368]. In addition, chromosomal translocations involving the BCL6 gene was detected in 48% of NLP-HLs [369,370].

3.8.4. Non-Hodgkin Lymphomas (nHLs)

The by itself also heterogeneous group of B-cell nHLs includes several subtypes, where the tumor cells have a gene-expression profile reflecting their equivalent healthy cells of origin. In addition, they show recurrent genetic, epigenetic, and other molecular changes [371]. Typical translocations include t(14;18) in follicular lymphoma (FL) (involving Bcl2), t(11;14) in mantle cell lymphoma (MCL) (involving cyclin D1), t(8;14) in Burkitt lymphoma (overexpressing the MYC gene). T-cell and NK/T-cell lymphomas are rare, making up about 10–15% of nHLs. Molecular and chromosomal studies are rarely used in addition to immunophenotyping and T-cell receptor rearrangement studies, with the only exception being the frequently recurring chromosomal translocation t(2;5)(p23;q35), characteristic of ALK-positive anaplastic large-cell lymphomas, where the ALK-negative form has an inferior outcome. In peripheral T-cell lymphomas, alterations in p53, bcl-2, bcl-xl, CD26, EBV, CCND2, CCR4, PRDM1, and TCR gamma are associated with unfavorable prognosis [372]. Cytotoxic T-cell and monocytic/dendritic signatures are also associated with poor prognosis in T-cell lymphomas. In addition, high expression of GATA3 and TBX21 are predictors of inferior outcome in the “Peripheral T-cell lymphoma, not otherwise specified (PTCL-NOS)” subgroup. MYC and IDH2 alterations were detected in a smaller portion of the patients, which holds a great additional therapeutic potential [373].
B-cell nHLs and genetic abnormalities. In DLBCLs, the activated B-cell subtype (ABC) displays worse outcome compared to the germinal center B-cell subtype (GCB). Patients in the ABC type frequently carry mutations in the B-cell receptor (BCR) and the NF-kB pathway genes, such as MYD88, CD79A/B, CARD11, and TNFAIP3, and display active BCR signaling. In addition, MYD88 mutations lead to constitutive activation of the JAK-STAT signaling pathway in these patients. On the other hand, BCL2, MYC, EZH2, and PTEN alterations are more commonly seen in patients with GCB type. Treatment sensitivity, however, are heterogenous in both groups, where MYC, BCL2, and TP53 alterations have been related to therapy resistance. In addition, TP53 mutations carry poor prognosis, especially when co-existing with altered MYD88 [370,374,375]. In FL, besides altered BCL2, genes involved in the JAK-STAT and NF-κB signaling pathways in addition to MYC dysregulations are common. Alterations in BCL2, TP53, EZH2, CREBBP, KMT2D, and MEF2B are more often seen in patients with FL transformation to DLBLC or Burkitt lymphoma [370,376,377,378,379,380]. In Burkitt lymphoma itself, c-MYC deregulations, such as translocations and mutations, is highly characteristic. Besides others, TP53, MDM4, CCND3, and TCF3 mutations have also been reported in Burkitt lymphoma [370]. MCLs characteristically have increased CCND1 expression. Additionally, dysregulation in TP53, MEF2B, NOTCH2, WHSC1, and BIRC3. The mTOR pathway is often activated in MCL, which is related to therapy resistance. In addition, TP53 mutations are related to poor prognosis and therapy resistant in MCL [370,380,381]. “Double-hit” (with MYC, BCL2, or BCL6 alterations) and “triple-hit” (with MYC, BCL2, and BCL6 alterations) lymphomas are also aggressive in nature, and are part of the high-grade B-cell lymphomas. Double-hit FLs, however, seem to have an indolent clinical behavior [382,383].
nHL treatment. nHL treatment is diverse and includes therapies from R-CHOP (rituxiban-cyclophosphamide, hydroxydaunorubicin, Oncovin, prednisone), through bone marrow transplantation, to gene transfer therapy, using chimeric antigen receptor-modified T-cell (CAR-T) therapies, and more. We will not discuss these in details, but excellent reviews, such as those from Nebrinsky and co-workers, Chung et al., Klener and Klanova, or Othman and co-workers are suggested in this topic for those interested [370,381,384,385]. In addition, therapeutic challenges for NK/T-cell lymphomas are discussed in articles, such as the ones from Xue and Zhang, or Wang and co-workers [386,387].
nHLs and mitochondria. Several genes involved in the development of nHL have been or will be discussed separately with their effect on mitochondrial metabolism and dynamics. Generally: In vitro studies showed that inhibiting glycolysis in lymphoma cells can induce apoptosis and may overcome multidrug resistance [388]. In DLBCL, different metabolic subgroups exist, with one being mitochondria-predominant, and the other one predominantly using fatty-acid oxidation and glutathione synthesis, latter insensitive to inhibitors of BCR signaling [389]. In addition, tumors with activated mTOR pathways may become resistant to mTOR inhibitors via upregulating glutaminase to enhance glutamine metabolism to replenish the TCA cycle, even in hypoxic conditions, which could potentially be targeted to fight therapy resistance. Tumors with MYC mutations for example are sensitive to glutaminase inhibition [380]. The heterogeneity of the nHL cells in vivo, however, underlines the need of individualized treatments. Lastly, in Burkitt lymphoma cells, increased citrate synthase, IDH, and decreased succinate dehydrogenase expression was described [390]. Interestingly, somatic mtDNA mutations are not significant in the pathogenesis of DLBCL [391], higher increased copies of mitochondrial DNA (mtDNA) associated with persistent disease status in nHL patients [390].

3.8.5. Chronic Lymphocytic Leukemia (CLL)

CLL is the most frequent type of leukemia in adults, characterized by the clonal expansion of predominantly mature B-lymphocytes. It has a generally good prognosis. A frequent genetic lesion, the 13q14 deletion is seen in 50–60% of patients and has a favorable prognosis. Interestingly, unlike in other types of mature B-cell malignancies, translocations involving the immunoglubulin heavy chain is rare. Deletion of 11q22-q23 (del11q), involving the tumor suppressor gene ataxia telangiectasia (ATM) is associated with poor prognosis. The frequent trisomy 12 carries an immediate risk by itself, but has a poor prognosis when co-existent with NOTCH1 mutations. In addition, Richter transformation is noted to be more common in patients with trisomy 12. Similarly, loss of TP53 with deletion of the short arm of chromosome 17, and TP53 mutations exhibit a poorer prognosis, and resistance to chemotherapy in CLL. Further mutations, such as SF3B1, KRAS, NRAS, BRAF, MYD88, EGR2, MAP2K1, ATM, NOTCH1, POT1, CHD2, XPO1, BIRC3, MED12, FBXW7, ASXL1, NFKBIE, TRAF3, RPS15, and DDX3X are also seen in CLL patients. SF3B1 carries a good prognosis [392,393,394,395].
CLL and mitochondrial changes. Increased mitochondrial mass with accelerated OXPHOS, and increased ROS generation are characteristic of CLL cells. Similarly, TP53 deletion or deficiency enhances glycolysis and increases mitochondrial mass, latter likely by enhanced PGC-1α expression in CLL cells. Both TP53 deficiency and 17p deletion subsequently leads to altered metabolism, impaired autophagy and subsequent Ibratinib-resistance [1,396,397,398]. Nonetheless, according to one study, there is a positive association of increased mtDNA copy number and future CLL risk [399].

3.8.6. Acute Lymphoblastic Leukemia (ALL)

The vast majority of ALLs are B lymphoblastic leukemias (B-ALLs), mostly occurring in children and adolescents. Inferior clinical outcome has been described in relation to the presence of BCR-ABL1 fusion gene, KMT2A (MLL) rearrangements, or hypodiploidity (<44 chromosomes). The genetic mutations are diverse in B-ALL and include: (1) Transcriptional factors promoting early lymphoid cell development, such as PAX5, IKZF1, EBF1, ETV6, ERG, GATA3, and LMO2. (2) Tumor suppressor genes and cell cycle regulators, such as TP53, RB1, and CDKN2A/CDKN2B, BAK1. (3) Cytokine receptors, such as CRLF2, RPOR (4) Kinases, such as ABL1, ABL2, CSF1R, JAK2, PDGFRB. (5) Ras signaling pathway genes, such as KRAS, NF1, NRAS, PTPN11. (6) Lymphoid signaling genes, such as BTLA, and CD200, and (7) Epigenetic modifiers, such as EZH2, CREBBP, SETD2, MLL2, and NSD2. Amongst others, alterations in IKZF1, TP53-, CDKN2A, CREBBP, NR3C2, MSH6, PRPS1, PRPS2, NT5C2 are ETV6 are associated with unfavorable outcome and/or therapy resistance in B-ALL [400,401,402,403]. Regarding mitochondrial biogenesis in B-ALL, a recent pre-clinical data showed high oxygen consumption, which can be a future therapeutic target in therapy-resistant B-ALL [404].

3.8.7. MYC

The MYC family includes three proto-oncogenes: MYC, MYCN, and MYCL. The MYC gene codes the c-Myc protein, a transcription factor involved in cell growth, differentiation, metabolism, and cell death regulation [405,406]. c-Myc has been reported to indirectly increase transcription of multiple genes via activating discrete gene sets, resulting in increased global mRNA expression and turnover [326,407]. Similarly to Ras, c- Myc can activate p53, induce ROS generation, replication stress, and can cause changes in cellular metabolism [326,345,408]. In addition, it increases glycolytic activity via increasing the expression of key metabolic enzymes in the glycolytic pathway, such as glucose transporter GLUT1 and GLUT3, and hexokinases that convert pyruvate to lactate. MYC also increases lactate and glutamine transporters. OXPHOS was found to be suppressed or increased with glycolysis typically predominating [7,329,345,408,409,410,411,412]. Besides increased ATP generation, these changes can effectively increase nucleotide, amino acid and fatty acid synthesis [41,412]. In addition, MYC increases β-oxidation [413] and glutaminolysis [329]. Myc also promotes mitochondrial fission via increased Drp1 and Mff [81], and increases mitochondrial biogenesis and replication via increased PLOG, PLOG2, and NRF1 expression [414]. The expression of mitochondrial trafficking proteins is also induced by c-Myc via upregulation of RHOT1, RHOT2, TRAK2, and Kif5B, resulting in enhanced organelle movements and mitochondrial recruitment to the cortical cytoskeleton [50].
Aberrant expression of c-Myc has been described in various hematologic malignancies, including Burkitt lymphoma, “double-hit/ “triple-hit” DLBCLs, high-grade B-cell lymphomas, multiple myeloma (MM), some mantle cell lymphomas, and AMLs. Increased c-Myc expressions alone or with additional Bcl-2 or Bcl-6 alterations correlate with poor clinical outcome, or therapy resistance [6,324,415,416,417]. Blocking c-Myc expression have been proven to interferes with tumor cell survival, motility and metastasis in various tumor models [50,410,418], with blocking mitochondrial pathways being an alternative or additional treatment possibility for tumors harboring MYC mutations or alterations resulting in overexpression.

3.9. Plasma Cell Myeloma, a.k.a Multiple Myeloma (MM)

3.9.1. MM and Genetic Changes

MM is a neoplastic disorder of plasma cells, which are terminally differentiated B-cells. Numerous genetic alterations, including mutations and epigenetic changes, have been described in the dysregulation of plasma cells that lead to dyscrasias. Some significant genetic changes include MYC, CCND1/CCND2/CCND3 (Cyclin D), FGFR3 (fibroblast growth factor receptor 3), MMSET (histone methyltransferase multiple myeloma SET domain), KRAS, NRAS, BRAF, TP53, and MAF/MAFA/MAFB (musculoaponeurotic fibrosarcoma) [6,329,419]. According to the 2017 WHO Classification of Tumors of Hematopoietic and Lymphoid Tissue, MAF translocations (t14;16 and t14;20) and deletion of 17p (TP53), although relatively less common, portrays a poor prognosis [6]. Nonetheless, despite the broad knowledge of genetic abnormalities in multiple myeloma, dysregulation of mitochondrial functions has been less explored.

3.9.2. MM and Necroptosis

As previously mentioned, since the initial discovery of necroptosis in the early 2000s, it has been increasingly investigated as a potential source of treatment for many conditions, including variety of neoplastic processes [420,421]. Prior studies on induction of necroptosis and mitochondrial fission in different malignancies, such as in multiple myeloma, had shown this to be a potential therapeutic target as seen by in vitro studies showing improved survival in multiple myeloma cells when regulators of fission were blocked [420].

3.9.3. MM and Cellular Metabolism

Changes in cellular metabolism in MM are numerous, including a radical increase in glycolysis and OXPHOS. Enhanced glycolysis leads to higher pyruvate levels within the cell, which can be turned into lactate or acetyl-CoA as a substrate for the TCA cycle to go through OXPHOS. Acetyl-CoA can also be used for fatty acid synthesis. These processes are essential for neoplastic plasma cells in MM to proliferate and produce immunoglobulins in a stringent environment such as the bone marrow. The boosting of metabolic substrates and their subsequent processes within MM cells occur through numerous genetic changes that connect to the previously discovered changes occurring in MM discussed previously.

3.9.4. MM, Cellular Metabolism and Effect of Genetic Changes

Genetic alterations in MYC, MAF, MMSET, FGFR3, and CCND1 lead to distinctly altered cellular metabolism through influence on enzymes, metabolic substrates, and various cellular transporters. The changes induced by these alterations enable enhanced cell survival and proliferation in MM. One of the main effectors in altered cellular metabolism is the protein kinase B cascade (AKT). This pathway plays a central role in multiple myeloma through metabolic orchestration leading to upregulation of glycolysis, glutaminolysis, pentose phosphate shunting, lipid synthesis, and nucleotide production while also decreasing fatty acid oxidation [329]. Furthermore, AKT is an essential inhibitor of FOXO, which is necessary for p53-dependent cell death. AKT is activated by cyclin D1, MYC, and FGFR3, while AKT further upregulates MMSET, which plays a pivotal role in opening chromatin for enhanced activity by demethylation. MMSET is also crucial for glycolytic processing and subsequent shunting through the pentose phosphate pathway (PPP). Interestingly, MMSET appears to be of particular importance in multiple myeloma cells containing the t(4;14) alteration. Cyclin D1 and MYC are stabilized by the activity of KRAS and ERK, while the activity of ubiquitin-specific peptidase 5 (USP5) helps prevent MAF degradation in t(4;14), t(11;14), t(14;16), and t(14;20) cells. Cancer lines with high levels of MAF are known to be resistant to proteasome inhibitors, such as bortezomib, through increased proteasomal activity. In addition, MYC appears to enhance both glycolysis and glutaminolysis in MM by acting on glutamine, lactate, and glucose transporters for augmented uptake from the environment [329].

3.9.5. MM and Mitochondrial Dynamics and Mitochondrial Biogenesis

Interestingly, MM cell lines have demonstrated increased fusion, which is thought to be regulated by Myc [329]. In addition, high expression of the mitochondrial fusion gene MFN1 is associated with inferior survival in MM [422]. On the other hand, mitochondrial fission, controlled by fission proteins Fis1 and Drp1, are necessary to handle a sudden rise in oxidative stress. To this same effect, multiple myeloma cells have been shown to induce the trafficking of mitochondria from neighboring stromal cells, which has been implicated in drug resistance development [1]. Furthermore, increased mitophagy has been described in MM, which can be related to drug resistance [1]. Increased expression of mtDNA translation, polymerase and helicase genes TFAM, PGC1A, POLG2, and TWNK have also been linked to poor prognosis in MM [422].

3.9.6. Mitochondrial Transfer to MM Cells

Interestingly, MM cells are able to internalize mitochondria, transferred from the neighboring nonmalignant bone marrow stromal cells, which seems to be at least in part CD38-driven [423].

3.9.7. MAF BZIP Transcription Factor (MAF)

Musculoaponeurotic fibrosarcoma (MAF) proteins, encoded by MAF, are basic region leucine zipper (bZIP)-type transcription factors. Large MAFs (lMAFs), such as MAFA, MAFB, c-MAF and NRL are pro-oncogenes, and bind to DNA sequences called MAF recognition elements (MAREs), where they modulate expression of several genes, such as CCDN2, FOS, JUN, CREB, MTORC2 and ATF [329,424,425]. On the other hand, homodimers of small MAFs (sMAF), such as MAFF, MAFG and MAFK bind to the same region via competitive binding, inhibiting transcription. The ratio of lMAFs and sMAFs are therefore important determinants of transcription activation. In addition, posttranslational modifications also alter the activity of the MAF proteins [329].
MAF has multiple roles in cellular metabolism in MM, including elevated insulin secretion for enhanced MM cell glucose uptake, activating ARK5, which has been linked to increased OXPHOS, and glutamine metabolism. Glycolysis and glutaminolysis are of particular importance in MM due to the need for antibody production. The ARK5/CDK4 inhibitor ON123300 has been shown to induce apoptosis in MM cell lines. MAF translocations, however, were not predictive of ARK5/CDK4 treatment in this study [329,426].
In about 60% of angioimmunoblastic T-cell lymphomas, c-MAF was found to be overexpressed, whereas in MM, aberrant c-MAF, MAFA, and MAFB gene expressions have been described. As mentioned previously, MAF mutation is associated with a poor prognosis in MM [329,425,427].

4. Therapies Targeting Mitochondria

Here we aim to summarize potential therapies targeting mitochondria, some of which we already have addressed in previous chapters. In addition, specific therapies for malignancies with certain genetic mutations can be found in the chapters where those genes are discussed in detail.

4.1. Drugs Affecting Cellular Metabolism

Cellular metabolism can be inhibited via various enzymes in different metabolic pathways, which are promising therapeutic targets in cancer therapy. These include drugs targeting ETC, TCA, glycolysis, and fatty acid oxidation and synthesis enzymes. Some of these enzymes are localized in the mitochondria, whereas others are found in the cytosol (Figure 1 and Table 3) [7,428,429,430,431,432,433,434,435,436,437,438]. In addition, increasing mitochondrial ROS production is also an important mitochondria-related cancer therapy strategy [14,20]. For further details on these topics in relation to mitochondria see Section 2.1.

4.2. Drugs Affecting mtDNA Pathways and Cell Death Regulation

Inhibition of mtDNA translation and transcription [7,29,30,31], along with apoptosis [41,43,44,45,439] and necroptosis [440,441] induction are promising anti-cancer therapies (Figure 2 and Table 3), discussed in more details in Section 2.2 and Section 2.3.

4.3. Drugs Affecting Mitochondrial Dynamics

Interfering with mitochondrial dynamics, including mitochondrial fission/fusion [442,443], mitophagy [444], and mitochondrial trafficking opens up a large number of new therapeutic targets in cancer therapy (Table 3), discussed in more details in Section 2.4 and Section 2.5.

5. Conclusions

There is substantial evidence that genetic alterations in hematologic malignancies are in close relationship with mitochondrial metabolism, dynamics, cell death pathways, and mtDNA transcription and translation. Targeting these pathways are promising future therapies, especially in therapy-resistant cases.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/life11121351/s1, Table S1: Abbreviations.

Author Contributions

Conceptualization, E.A.W.-G. and I.C.; writing—original draft preparation, E.A.W.-G., I.C., A.L.G., M.W.; writing—review and editing, E.A.W.-G., I.C., A.L.G., Y.L., J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barbato, A.; Scandura, G.; Puglisi, F.; Cambria, D.; La Spina, E.; Palumbo, G.A.; Lazzarino, G.; Tibullo, D.; Di Raimondo, F.; Giallongo, C.; et al. Mitochondrial Bioenergetics at the Onset of Drug Resistance in Hematological Malignancies: An Overview. Front. Oncol. 2020, 10, 604143. [Google Scholar] [CrossRef]
  2. Popov, L. Mitochondrial biogenesis: An update. J. Cell. Mol. Med. 2020, 24, 4892–4899. [Google Scholar] [CrossRef] [Green Version]
  3. Serasinghe, M.N.; Chipuk, J.E. Mitochondrial Fission in Human Diseases. Handb. Exp. Pharmacol. 2017, 240, 159–188. [Google Scholar] [CrossRef] [PubMed]
  4. Paul, B.T.; Manz, D.H.; Torti, F.M.; Torti, S.V. Mitochondria and Iron: Current questions. Expert Rev. Hematol. 2017, 10, 65–79. [Google Scholar] [CrossRef] [Green Version]
  5. Miyazono, Y.; Hirashima, S.; Ishihara, N.; Kusukawa, J.; Nakamura, K.-I.; Ohta, K. Uncoupled mitochondria quickly shorten along their long axis to form indented spheroids, instead of rings, in a fission-independent manner. Sci. Rep. 2018, 8, 350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Swerdlow, S.H.; Campo, E.; Harris, N.L.; Jaffe, E.S.; Pileri, S.A.; Stein, H.; Thiele, J.; Arber, D.; Hasserjian, R.; Le Beau, M. WHO Classification of Tumours of Haematopoietic and Lymphoid Tissues; International Agency for Research on Cancer: Lyon, France, 2017. [Google Scholar]
  7. Moindjie, H.; Rodrigues-Ferreira, S.; Nahmias, C. Mitochondrial Metabolism in Carcinogenesis and Cancer Therapy. Cancers 2021, 13, 3311. [Google Scholar] [CrossRef]
  8. Hipólito, A.; Martins, F.; Mendes, C.; Lopes-Coelho, F.; Serpa, J. Molecular and Metabolic Reprogramming: Pulling the Strings Toward Tumor Metastasis. Front. Oncol. 2021, 11, 656851. [Google Scholar] [CrossRef] [PubMed]
  9. Rusz, M.; Del Favero, G.; El Abiead, Y.; Gerner, C.; Keppler, B.K.; Jakupec, M.A.; Koellensperger, G. Morpho-metabotyping the oxidative stress response. Sci. Rep. 2021, 11, 15471. [Google Scholar] [CrossRef] [PubMed]
  10. Warburg, O. The Metabolism of Carcinoma Cells. J. Cancer Res. 1925, 9, 148–163. [Google Scholar] [CrossRef] [Green Version]
  11. Liberti, M.V.; Locasale, J.W. The Warburg Effect: How Does it Benefit Cancer Cells? Trends Biochem. Sci. 2016, 41, 211–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Smolková, K.; Plecitá-Hlavatá, L.; Bellance, N.; Benard, G.; Rossignol, R.; Ježek, P. Waves of gene regulation suppress and then restore oxidative phosphorylation in cancer cells. Int. J. Biochem. Cell Biol. 2011, 43, 950–968. [Google Scholar] [CrossRef]
  13. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  14. Altman, B.J.; Stine, Z.E.; Dang, C.V. Erratum: From Krebs to clinic: Glutamine metabolism to cancer therapy. Nat. Rev. Cancer 2016, 16, 773. [Google Scholar] [CrossRef] [Green Version]
  15. Jeyaraju, D.; Hurren, R.; Wang, X.; MacLean, N.; Gronda, M.; Shamas-Din, A.; Minden, M.D.; Giaever, G.; Schimmer, A.D. A novel isoflavone, ME-344, targets the cytoskeleton in acute myeloid leukemia. Oncotarget 2016, 7, 49777–49785. [Google Scholar] [CrossRef] [Green Version]
  16. Golub, D.; Iyengar, N.; Dogra, S.; Wong, T.; Bready, D.; Tang, K.; Modrek, A.S.; Placantonakis, D.G. Mutant Isocitrate Dehydrogenase Inhibitors as Targeted Cancer Therapeutics. Front. Oncol. 2019, 9, 417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Aggarwal, V.; Tuli, H.S.; Varol, A.; Thakral, F.; Yerer, M.B.; Sak, K.; Varol, M.; Jain, A.; Khan, A.; Sethi, G. Role of Reactive Oxygen Species in Cancer Progression: Molecular Mechanisms and Recent Advancements. Biomolecules 2019, 9, 735. [Google Scholar] [CrossRef] [Green Version]
  18. Pelicano, H.; Feng, L.; Zhou, Y.; Carew, J.S.; Hileman, E.O.; Plunkett, W.; Keating, M.J.; Huang, P. Inhibition of Mitochondrial Respiration: A novel strategy to enhance drug-induced apoptosis in human leukemia cells by a reactive oxygen species-mediated mechanism. J. Biol. Chem. 2003, 278, 37832–37839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Klein, K.; He, K.; Younes, A.I.; Barsoumian, H.B.; Chen, D.; Ozgen, T.; Mosaffa, S.; Patel, R.R.; Gu, M.; Novaes, J.; et al. Role of Mitochondria in Cancer Immune Evasion and Potential Therapeutic Approaches. Front. Immunol. 2020, 11, 573326. [Google Scholar] [CrossRef] [PubMed]
  20. Ansell, S.M. Hodgkin lymphoma: MOPP chemotherapy to PD-1 blockade and beyond. Am. J. Hematol. 2016, 91, 109–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Li, L.; Qi, R.; Zhang, L.; Yu, Y.; Hou, J.; Gu, Y.; Song, D.; Wang, X. Potential biomarkers and targets of mitochondrial dynamics. Clin. Transl. Med. 2021, 11, e529. [Google Scholar] [CrossRef] [PubMed]
  22. Joaquim, M.; Escobar-Henriques, M. Role of Mitofusins and Mitophagy in Life or Death Decisions. Front. Cell Dev. Biol. 2020, 8, 572182. [Google Scholar] [CrossRef]
  23. Rong, Z.; Tu, P.; Xu, P.; Sun, Y.; Yu, F.; Tu, N.; Guo, L.; Yang, Y. The Mitochondrial Response to DNA Damage. Front. Cell Dev. Biol. 2021, 9, 669379. [Google Scholar] [CrossRef]
  24. Falkenberg, M. Mitochondrial DNA replication in mammalian cells: Overview of the pathway. Essays Biochem. 2018, 62, 287–296. [Google Scholar] [CrossRef]
  25. Moreira, O.C.; Estébanez, B.; Martínez-Florez, S.; De Paz, J.A.; Cuevas, M.J.; González-Gallego, J. Mitochondrial Function and Mitophagy in the Elderly: Effects of Exercise. Oxidative Med. Cell. Longev. 2017, 2017, 2012798. [Google Scholar] [CrossRef]
  26. Macias, E.; Rao, D.; Carbajal, S.; Kiguchi, K.; DiGiovanni, J. Stat3 Binds to mtDNA and Regulates Mitochondrial Gene Expression in Keratinocytes. J. Investig. Dermatol. 2014, 134, 1971–1980. [Google Scholar] [CrossRef] [Green Version]
  27. Rusecka, J.; Kaliszewska, M.; Bartnik, E.; Tońska, K. Nuclear genes involved in mitochondrial diseases caused by instability of mitochondrial DNA. J. Appl. Genet. 2018, 59, 43–57. [Google Scholar] [CrossRef] [Green Version]
  28. Fontana, G.A.; Gahlon, H.L. Mechanisms of replication and repair in mitochondrial DNA deletion formation. Nucleic Acids Res. 2020, 48, 11244–11258. [Google Scholar] [CrossRef]
  29. Bonekamp, N.A.; Peter, B.; Hillen, H.S.; Felser, A.; Bergbrede, T.; Choidas, A.; Horn, M.; Unger, A.; Di Lucrezia, R.; Atanassov, I.; et al. Small-molecule inhibitors of human mitochondrial DNA transcription. Nature 2020, 588, 712–716. [Google Scholar] [CrossRef] [PubMed]
  30. Škrtić, M.; Sriskanthadevan, S.; Jhas, B.; Gebbia, M.; Wang, X.; Wang, Z.; Hurren, R.; Jitkova, Y.; Gronda, M.; Maclean, N.; et al. Inhibition of Mitochondrial Translation as a Therapeutic Strategy for Human Acute Myeloid Leukemia. Cancer Cell 2011, 20, 674–688. [Google Scholar] [CrossRef] [Green Version]
  31. Kuntz, E.M.; Baquero, P.; Michie, A.M.; Dunn, K.; Tardito, S.; Holyoake, T.L.; Helgason, G.V.; Gottlieb, E. Targeting mitochondrial oxidative phosphorylation eradicates therapy-resistant chronic myeloid leukemia stem cells. Nat. Med. 2017, 23, 1234–1240. [Google Scholar] [CrossRef] [Green Version]
  32. Hu, L.; Yao, X.; Shen, Y. Altered mitochondrial DNA copy number contributes to human cancer risk: Evidence from an updated meta-analysis. Sci. Rep. 2016, 6, 35859. [Google Scholar] [CrossRef] [Green Version]
  33. Chatterjee, A.; Mambo, E.; Sidransky, D. Mitochondrial DNA mutations in human cancer. Oncogene 2006, 25, 4663–4674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hertweck, K.; Dasgupta, S. The Landscape of mtDNA Modifications in Cancer: A Tale of Two Cities. Front. Oncol. 2017, 7, 262. [Google Scholar] [CrossRef]
  35. Chaudhary, S.; Ganguly, S.; Palanichamy, J.K.; Singh, A.; Bakhshi, R.; Jain, A.; Chopra, A.; Bakhshi, S. PGC1A driven enhanced mitochondrial DNA copy number predicts outcome in pediatric acute myeloid leukemia. Mitochondrion 2021, 58, 246–254. [Google Scholar] [CrossRef]
  36. Kim, C.; Bassig, B.A.; Seow, W.J.; Hu, W.; Purdue, M.; Huang, W.-Y.; Liu, C.-S.; Cheng, W.-L.; Männistö, S.; Vermeulen, R.; et al. Mitochondrial DNA Copy Number and Chronic Lymphocytic Leukemia/Small Lymphocytic Lymphoma Risk in Two Prospective Studies. Cancer Epidemiology Biomarkers Prev. 2014, 24, 148–153. [Google Scholar] [CrossRef] [Green Version]
  37. Jain, A.; Bakhshi, S.; Thakkar, H.; Gerards, M.; Singh, A. Elevated mitochondrial DNA copy numbers in pediatric acute lymphoblastic leukemia: A potential biomarker for predicting inferior survival. Pediatr. Blood Cancer 2018, 65, e26874. [Google Scholar] [CrossRef]
  38. Jones, S.W.; Ball, A.L.; Chadwick, A.E.; Alfirevic, A. The Role of Mitochondrial DNA Variation in Drug Response: A Systematic Review. Front. Genet. 2021, 12, 698825. [Google Scholar] [CrossRef]
  39. D’Arcy, M.S. Cell death: A review of the major forms of apoptosis, necrosis and autophagy. Cell Biol. Int. 2019, 43, 582–592. [Google Scholar] [CrossRef]
  40. Dadsena, S.; Zollo, C.; García-Sáez, A.J. Mechanisms of mitochondrial cell death. Biochem. Soc. Trans. 2021, 49, 663–674. [Google Scholar] [CrossRef]
  41. Panuzzo, C.; Jovanovski, A.; Pergolizzi, B.; Pironi, L.; Stanga, S.; Fava, C.; Cilloni, D. Mitochondria: A Galaxy in the Hematopoietic and Leukemic Stem Cell Universe. Int. J. Mol. Sci. 2020, 21, 3928. [Google Scholar] [CrossRef]
  42. Degterev, A.; Huang, Z.; Boyce, M.; Li, Y.; Jagtap, P.; Mizushima, N.; Cuny, G.D.; Mitchison, T.J.; Moskowitz, M.A.; Yuan, J. Chemical inhibitor of nonapoptotic cell death with therapeutic potential for ischemic brain injury. Nat. Chem. Biol. 2005, 1, 112–119. [Google Scholar] [CrossRef]
  43. Reyna, D.E.; Garner, T.P.; Lopez, A.; Kopp, F.; Choudhary, G.S.; Sridharan, A.; Narayanagari, S.-R.; Mitchell, K.; Dong, B.; Bartholdy, B.A.; et al. Direct Activation of BAX by BTSA1 Overcomes Apoptosis Resistance in Acute Myeloid Leukemia. Cancer Cell 2017, 32, 490–505. [Google Scholar] [CrossRef] [Green Version]
  44. Reichenbach, F.; Wiedenmann, C.; Schalk, E.; Becker, D.; Funk, K.; Scholz-Kreisel, P.; Todt, F.; Wolleschak, D.; Döhner, K.; Marquardt, J.U.; et al. Mitochondrial BAX Determines the Predisposition to Apoptosis in Human AML. Clin. Cancer Res. 2017, 23, 4805–4816. [Google Scholar] [CrossRef] [Green Version]
  45. Niu, X.; Zhao, J.; Ma, J.; Xie, C.; Edwards, H.; Wang, G.; Caldwell, J.T.; Xiang, S.; Zhang, X.; Chu, R.; et al. Binding of Released Bim to Mcl-1 is a Mechanism of Intrinsic Resistance to ABT-199 which can be Overcome by Combination with Daunorubicin or Cytarabine in AML Cells. Clin. Cancer Res. 2016, 22, 4440–4451. [Google Scholar] [CrossRef] [Green Version]
  46. Gong, Y.; Fan, Z.; Luo, G.; Yang, C.; Huang, Q.; Fan, K.; Cheng, H.; Jin, K.; Ni, Q.; Yu, X.; et al. The role of necroptosis in cancer biology and therapy. Mol. Cancer 2019, 18, 100. [Google Scholar] [CrossRef] [Green Version]
  47. Xie, L.; Shi, F.; Tan, Z.; Li, Y.; Bode, A.M.; Cao, Y. Mitochondrial network structure homeostasis and cell death. Cancer Sci. 2018, 109, 3686–3694. [Google Scholar] [CrossRef]
  48. Wappler, E.A.; Institoris, A.; Dutta, S.; Katakam, P.; Busija, D.W. Mitochondrial Dynamics Associated with Oxygen-Glucose Deprivation in Rat Primary Neuronal Cultures. PLoS ONE 2013, 8, e63206. [Google Scholar] [CrossRef]
  49. Yu, R.; Lendahl, U.; Nistér, M.; Zhao, J. Regulation of Mammalian Mitochondrial Dynamics: Opportunities and Challenges. Front. Endocrinol. 2020, 11, 374. [Google Scholar] [CrossRef]
  50. Agarwal, E.; Altman, B.J.; Seo, J.H.; Bertolini, I.; Ghosh, J.C.; Kaur, A.; Kossenkov, A.V.; Languino, L.R.; Gabrilovich, D.I.; Speicher, D.W.; et al. Myc Regulation of a Mitochondrial Trafficking Network Mediates Tumor Cell Invasion and Metastasis. Mol. Cell. Biol. 2019, 39, e00109-19. [Google Scholar] [CrossRef] [Green Version]
  51. Suen, D.-F.; Norris, K.L.; Youle, R.J. Mitochondrial dynamics and apoptosis. Genes Dev. 2008, 22, 1577–1590. [Google Scholar] [CrossRef] [Green Version]
  52. Horbay, R.; Bilyy, R. Mitochondrial dynamics during cell cycling. Apoptosis 2016, 21, 1327–1335. [Google Scholar] [CrossRef]
  53. Ong, S.-B.; Kalkhoran, S.B.; Cabrera-Fuentes, H.A.; Hausenloy, D.J. Mitochondrial fusion and fission proteins as novel therapeutic targets for treating cardiovascular disease. Eur. J. Pharmacol. 2015, 763, 104–114. [Google Scholar] [CrossRef] [Green Version]
  54. Harris, J.; Deen, N.; Zamani, S.; Hasnat, M.A. Mitophagy and the release of inflammatory cytokines. Mitochondrion 2018, 41, 2–8. [Google Scholar] [CrossRef]
  55. Li, S.; Zhang, J.; Liu, C.; Wang, Q.; Yan, J.; Hui, L.; Jia, Q.; Shan, H.; Tao, L.; Zhang, M. The Role of Mitophagy in Regulating Cell Death. Oxidative Med. Cell. Longev. 2021, 2021, 6617256. [Google Scholar] [CrossRef]
  56. Palikaras, K.; Lionaki, E.; Tavernarakis, N. Mechanisms of mitophagy in cellular homeostasis, physiology and pathology. Nat. Cell Biol. 2018, 20, 1013–1022. [Google Scholar] [CrossRef]
  57. Strappazzon, F.; Nazio, F.; Corrado, M.; Cianfanelli, V.; Romagnoli, A.; Fimia, G.M.; Campello, S.; Nardacci, R.; Piacentini, M.; Campanella, M.; et al. AMBRA1 is able to induce mitophagy via LC3 binding, regardless of PARKIN and p62/SQSTM1. Cell Death Differ. 2015, 22, 419–432. [Google Scholar] [CrossRef] [Green Version]
  58. Pendin, D.; Filadi, R.; Pizzo, P. The Concerted Action of Mitochondrial Dynamics and Positioning: New Characters in Cancer Onset and Progression. Front. Oncol. 2017, 7, 102. [Google Scholar] [CrossRef] [Green Version]
  59. Margineantu, D.H.; Cox, W.G.; Sundell, L.; Sherwood, S.W.; Beechem, J.M.; Capaldi, R.A. Cell cycle dependent morphology changes and associated mitochondrial DNA redistribution in mitochondria of human cell lines. Mitochondrion 2002, 1, 425–435. [Google Scholar] [CrossRef]
  60. Rehman, J.; Zhang, H.J.; Toth, P.T.; Zhang, Y.; Marsboom, G.; Hong, Z.; Salgia, R.; Husain, A.N.; Wietholt, C.; Archer, S.L. Inhibition of mitochondrial fission prevents cell cycle progression in lung cancer. FASEB J. 2012, 26, 2175–2186. [Google Scholar] [CrossRef] [Green Version]
  61. Inoue-Yamauchi, A.; Oda, H. Depletion of mitochondrial fission factor DRP1 causes increased apoptosis in human colon cancer cells. Biochem. Biophys. Res. Commun. 2012, 421, 81–85. [Google Scholar] [CrossRef]
  62. Gomes, L.C.; Di Benedetto, G.; Scorrano, L. During autophagy mitochondria elongate, are spared from degradation and sustain cell viability. Nat. Cell Biol. 2011, 13, 589–598. [Google Scholar] [CrossRef] [Green Version]
  63. Galloway, C.; Lee, H.; Yoon, Y. Mitochondrial morphology—emerging role in bioenergetics. Free. Radic. Biol. Med. 2012, 53, 2218–2228. [Google Scholar] [CrossRef] [Green Version]
  64. Benard, G.; Bellance, N.; James, D.; Parrone, P.; Fernandez, H.; Letellier, T.; Rossignol, R. Mitochondrial bioenergetics and structural network organization. J. Cell Sci. 2007, 120, 838–848. [Google Scholar] [CrossRef] [Green Version]
  65. Filadi, R.; Pendin, D.; Pizzo, P. Mitofusin 2: From functions to disease. Cell Death Dis. 2018, 9, 330. [Google Scholar] [CrossRef]
  66. Neuspiel, M.; Zunino, R.; Gangaraju, S.; Rippstein, P.; McBride, H. Activated Mitofusin 2 Signals Mitochondrial Fusion, Interferes with Bax Activation, and Reduces Susceptibility to Radical Induced Depolarization. J. Biol. Chem. 2005, 280, 25060–25070. [Google Scholar] [CrossRef] [Green Version]
  67. Karbowski, M.; Lee, Y.-J.; Gaume, B.; Jeong, S.-Y.; Frank, S.; Nechushtan, A.; Santel, A.; Fuller, M.T.; Smith, C.L.; Youle, R.J. Spatial and temporal association of Bax with mitochondrial fission sites, Drp1, and Mfn2 during apoptosis. J. Cell Biol. 2002, 159, 931–938. [Google Scholar] [CrossRef] [Green Version]
  68. Milani, M.; Beckett, A.J.; Al-Zebeeby, A.; Luo, X.; Prior, I.A.; Cohen, G.M.; Varadarajan, S. DRP-1 functions independently of mitochondrial structural perturbations to facilitate BH3 mimetic-mediated apoptosis. Cell Death Discov. 2019, 5, 117. [Google Scholar] [CrossRef] [Green Version]
  69. Tanaka, A.; Youle, R.J. A chemical inhibitor of DRP1 uncouples mitochondrial fission and apoptosis. Mol. Cell 2008, 29, 409–410. [Google Scholar] [CrossRef]
  70. Hasnat, M.; Yuan, Z.; Ullah, A.; Naveed, M.; Raza, F.; Baig, M.M.F.A.; Khan, A.; Xu, D.; Su, Y.; Sun, L.; et al. Mitochondria-dependent apoptosis in triptolide-induced hepatotoxicity is associated with the Drp1 activation. Toxicol. Mech. Methods 2020, 30, 124–133. [Google Scholar] [CrossRef]
  71. Parone, P.A.; James, D.I.; Da Cruz, S.; Mattenberger, Y.; Donzé, O.; Barja, F.; Martinou, J.-C. Inhibiting the Mitochondrial Fission Machinery Does Not Prevent Bax/Bak-Dependent Apoptosis. Mol. Cell. Biol. 2006, 26, 7397–7408. [Google Scholar] [CrossRef] [Green Version]
  72. Estaquier, J.; Arnoult, D. Inhibiting Drp1-mediated mitochondrial fission selectively prevents the release of cytochrome c during apoptosis. Cell Death Differ. 2007, 14, 1086–1094. [Google Scholar] [CrossRef] [Green Version]
  73. Frezza, C.; Cipolat, S.; De Brito, O.M.; Micaroni, M.; Beznoussenko, G.V.; Rudka, T.; Bartoli, D.; Polishuck, R.S.; Danial, N.N.; De Strooper, B.; et al. OPA1 Controls Apoptotic Cristae Remodeling Independently from Mitochondrial Fusion. Cell 2006, 126, 177–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Dong, T.; Zhang, X.; Liu, Y.; Xu, S.; Chang, H.; Chen, F.; Pan, L.; Hu, S.; Wang, M.; Lu, M. Opa1 Prevents Apoptosis and Cisplatin-Induced Ototoxicity in Murine Cochleae. Front. Cell Dev. Biol. 2021, 9, 744838. [Google Scholar] [CrossRef]
  75. Brooks, C.; Wei, Q.; Feng, L.; Dong, G.; Tao, Y.; Mei, L.; Xie, Z.-J.; Dong, Z. Bak regulates mitochondrial morphology and pathology during apoptosis by interacting with mitofusins. Proc. Natl. Acad. Sci. USA 2007, 104, 11649–11654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Yu, T.; Sheu, S.-S.; Robotham, J.L.; Yoon, Y. Mitochondrial fission mediates high glucose-induced cell death through elevated production of reactive oxygen species. Cardiovasc. Res. 2008, 79, 341–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Tian, Y.; Huang, Z.; Wang, Z.; Yin, C.; Zhou, L.; Zhang, L.; Huang, K.; Zhou, H.; Jiang, X.; Li, J.; et al. Identification of Novel Molecular Markers for Prognosis Estimation of Acute Myeloid Leukemia: Over-Expression of PDCD7, FIS1 and Ang2 May Indicate Poor Prognosis in Pretreatment Patients with Acute Myeloid Leukemia. PLoS ONE 2014, 9, e84150. [Google Scholar] [CrossRef]
  78. Kitamura, S.; Yanagi, T.; Imafuku, K.; Hata, H.; Abe, R.; Shimizu, H. Drp1 regulates mitochondrial morphology and cell proliferation in cutaneous squamous cell carcinoma. J. Dermatol. Sci. 2017, 88, 298–307. [Google Scholar] [CrossRef] [Green Version]
  79. Zhan, L.; Cao, H.; Wang, G.; Lyu, Y.; Sun, X.; An, J.; Wu, Z.; Huang, Q.; Liu, B.; Xing, J. Drp1-mediated mitochondrial fission promotes cell proliferation through crosstalk of p53 and NF-κB pathways in hepatocellular carcinoma. Oncotarget 2016, 7, 65001–65011. [Google Scholar] [CrossRef] [Green Version]
  80. Kashatus, J.A.; Nascimento, A.; Myers, L.J.; Sher, A.; Byrne, F.; Hoehn, K.; Counter, C.M.; Kashatus, D. Erk2 Phosphorylation of Drp1 Promotes Mitochondrial Fission and MAPK-Driven Tumor Growth. Mol. Cell 2015, 57, 537–551. [Google Scholar] [CrossRef] [Green Version]
  81. Seo, J.H.; Agarwal, E.; Chae, Y.C.; Lee, Y.G.; Garlick, D.S.; Storaci, A.M.; Ferrero, S.; Gaudioso, G.; Gianelli, U.; Vaira, V.; et al. Mitochondrial fission factor is a novel Myc-dependent regulator of mitochondrial permeability in cancer. EBioMedicine 2019, 48, 353–363. [Google Scholar] [CrossRef]
  82. Xu, K.; Chen, G.; Li, X.; Wu, X.; Chang, Z.; Xu, J.; Zhu, Y.; Yin, P.; Liang, X.; Dong, L. MFN2 suppresses cancer progression through inhibition of mTORC2/Akt signaling. Sci. Rep. 2017, 7, 41718. [Google Scholar] [CrossRef]
  83. Xin, Y.; Li, J.; Wu, W.; Liu, X. Mitofusin-2: A New Mediator of Pathological Cell Proliferation. Front. Cell Dev. Biol. 2021, 9, 647631. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, W.; Lu, J.; Zhu, F.; Wei, J.; Jia, C.; Zhang, Y.; Zhou, L.; Xie, H.; Zheng, S. Pro-apoptotic and anti-proliferative effects of mitofusin-2 via Bax signaling in hepatocellular carcinoma cells. Med Oncol. 2012, 29, 70–76. [Google Scholar] [CrossRef]
  85. Szabadkai, G.; Simoni, A.M.; Chami, M.; Wieckowski, M.; Youle, R.J.; Rizzuto, R. Drp-1-Dependent Division of the Mitochondrial Network Blocks Intraorganellar Ca2+ Waves and Protects against Ca2+-Mediated Apoptosis. Mol. Cell 2004, 16, 59–68. [Google Scholar] [CrossRef] [PubMed]
  86. Ham, S.J.; Lee, D.; Yoo, H.; Jun, K.; Shin, H.; Chung, J. Decision between mitophagy and apoptosis by Parkin via VDAC1 ubiquitination. Proc. Natl. Acad. Sci. USA 2020, 117, 4281–4291. [Google Scholar] [CrossRef] [PubMed]
  87. Ney, P.A. Mitochondrial autophagy: Origins, significance, and role of BNIP3 and NIX. Biochim. Biophys. Acta 2015, 1853, 2775–2783. [Google Scholar] [CrossRef] [Green Version]
  88. Stergiou, I.; Kapsogeorgou, E. Autophagy and Metabolism in Normal and Malignant Hematopoiesis. Int. J. Mol. Sci. 2021, 22, 8540. [Google Scholar] [CrossRef]
  89. Lazarini, M.; Machado-Neto, J.A.; Duarte, A.D.S.S.; Pericole, F.V.; Vieira, K.P.; Niemann, F.S.; Alvarez, M.; Traina, F.; Saad, S.T.O. BNIP3L in myelodysplastic syndromes and acute myeloid leukemia: Impact on disease outcome and cellular response to decitabine. Haematologica 2016, 101, e445–e448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Jiang, H.; Yang, L.; Guo, L.; Cui, N.; Zhang, G.; Liu, C.; Xing, L.; Shao, Z.; Wang, H. Impaired Mitophagy of Nucleated Erythroid Cells Leads to Anemia in Patients with Myelodysplastic Syndromes. Oxidative Med. Cell. Longev. 2018, 2018, 6328051. [Google Scholar] [CrossRef] [PubMed]
  91. van de Loosdrecht, A.A.; Brada, S.J.; Blom, N.R.; Hendriks, D.W.; Smit, J.W.; Berg, E.V.D.; de Wolf, J.T.; Vellenga, E. Mitochondrial disruption and limited apoptosis of erythroblasts are associated with high risk myelodysplasia. An ultrastructural analysis. Leuk. Res. 2001, 25, 385–393. [Google Scholar] [CrossRef]
  92. Houwerzijl, E.J.; Pol, H.-W.D.; Blom, N.R.; van der Want, J.; De Wolf, J.T.; Vellenga, E. Erythroid precursors from patients with low-risk myelodysplasia demonstrate ultrastructural features of enhanced autophagy of mitochondria. Leukemia 2009, 23, 886–891. [Google Scholar] [CrossRef] [PubMed]
  93. Stergiou, I.; Kambas, K.; Poulaki, A.; Giannouli, S.; Katsila, T.; Dimitrakopoulou, A.; Vidali, V.; Mouchtouris, V.; Kloukina, I.; Xingi, E.; et al. Exploiting the Role of Hypoxia-Inducible Factor 1 and Pseudohypoxia in the Myelodysplastic Syndrome Pathophysiology. Int. J. Mol. Sci. 2021, 22, 4099. [Google Scholar] [CrossRef]
  94. Furnish, M.; Caino, M.C. Altered mitochondrial trafficking as a novel mechanism of cancer metastasis. Cancer Rep. 2020, 3, e1157. [Google Scholar] [CrossRef]
  95. Shanmughapriya, S.; Langford, D.; Natarajaseenivasan, K. Inter and Intracellular mitochondrial trafficking in health and disease. Ageing Res. Rev. 2020, 62, 101128. [Google Scholar] [CrossRef]
  96. Sneeggen, M.; Guadagno, N.A.; Progida, C. Intracellular Transport in Cancer Metabolic Reprogramming. Front. Cell Dev. Biol. 2020, 8, 597608. [Google Scholar] [CrossRef]
  97. Schuler, M.-H.; Lewandowska, A.; Di Caprio, G.; Skillern, W.; Upadhyayula, S.; Kirchhausen, T.; Shaw, J.M.; Cunniff, B. Miro1-mediated mitochondrial positioning shapes intracellular energy gradients required for cell migration. Mol. Biol. Cell 2017, 28, 2159–2169. [Google Scholar] [CrossRef] [PubMed]
  98. Spoor, J.; Fard, H.F.; Rezaei, N. Congenital neutropenia and primary immunodeficiency diseases. Crit. Rev. Oncol. 2019, 133, 149–162. [Google Scholar] [CrossRef]
  99. Farruggia, P.; Di Marco, F.; Dufour, C. Pearson syndrome. Expert Rev. Hematol. 2018, 11, 239–246. [Google Scholar] [CrossRef] [PubMed]
  100. Finsterer, J.; Frank, M. Haematological abnormalities in mitochondrial disorders. Singap. Med. J. 2015, 56, 412–419. [Google Scholar] [CrossRef] [Green Version]
  101. Ghosh, P.; Vidal, C.; Dey, S.; Zhang, L. Mitochondria Targeting as an Effective Strategy for Cancer Therapy. Int. J. Mol. Sci. 2020, 21, 3363. [Google Scholar] [CrossRef]
  102. Fenaux, P.; Haase, D.; Santini, V.; Sanz, G.; Platzbecker, U.; Mey, U. Myelodysplastic syndromes: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2021, 32, 142–156. [Google Scholar] [CrossRef]
  103. Estey, E.; Döhner, H. Acute myeloid leukaemia. Lancet 2006, 368, 1894–1907. [Google Scholar] [CrossRef]
  104. Lagunas-Rangel, F.A.; Chávez-Valencia, V.; Gómez-Guijosa, M.Á.; Cortes-Penagos, C. Acute Myeloid Leukemia-Genetic Altera-tions and Their Clinical Prognosis. Int. J. Hematol. Oncol. Stem. Cell Res. 2017, 11, 328–339. [Google Scholar]
  105. Bullinger, L.; Döhner, K.; Döhner, H. Genomics of Acute Myeloid Leukemia Diagnosis and Pathways. J. Clin. Oncol. 2017, 35, 934–946. [Google Scholar] [CrossRef]
  106. Østgård, L.S.G.; Medeiros, B.C.; Sengeløv, H.; Nørgaard, M.; Andersen, M.K.; Dufva, I.H.; Friis, L.S.; Kjeldsen, E.; Marcher, C.W.; Preiss, B.; et al. Epidemiology and Clinical Significance of Secondary and Therapy-Related Acute Myeloid Leukemia: A National Population-Based Cohort Study. J. Clin. Oncol. 2015, 33, 3641–3649. [Google Scholar] [CrossRef]
  107. Hulegårdh, E.; Nilsson, C.; Lazarevic, V.; Garelius, H.; Antunovic, P.; Derolf, R.; Möllgård, L.; Uggla, B.; Wennström, L.; Wahlin, A.; et al. Characterization and prognostic features of secondary acute myeloid leukemia in a population-based setting: A report from the Swedish Acute Leukemia Registry. Am. J. Hematol. 2015, 90, 208–214. [Google Scholar] [CrossRef]
  108. Kayser, S.; Döhner, K.; Krauter, J.; Köhne, C.-H.; Horst, H.A.; Held, G.; von Lilienfeld-Toal, M.; Wilhelm, S.; Kündgen, A.; Götze, K.; et al. The impact of therapy-related acute myeloid leukemia (AML) on outcome in 2853 adult patients with newly diagnosed AML. Blood 2011, 117, 2137–2145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Heuser, M.; Ofran, Y.; Boissel, N.; Mauri, S.B.; Craddock, C.; Janssen, J.; Wierzbowska, A.; Buske, C.; ESMO Guidelines Committee. Acute myeloid leukaemia in adult patients: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2020, 31, 697–712. [Google Scholar] [CrossRef] [PubMed]
  110. Sun, Y.; Chen, B.-R.; Deshpande, A.; Sun, Y.; Chen, B.-R.; Deshpande, A. Epigenetic Regulators in the Development, Maintenance, and Therapeutic Targeting of Acute Myeloid Leukemia. Front. Oncol. 2018, 8, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Rubnitz, J.E.; Gibson, B.; Smith, F.O. Acute Myeloid Leukemia. Hematol. Clin. N. Am. 2010, 24, 35–63. [Google Scholar] [CrossRef]
  112. Nichol, J.N.; Assouline, S.; Miller, W.H.; Goldman, J.M.; Dutcher, J.P.; Kyle, R.A.; Wiernik, P.H. The etiology of acute leukemia. In Neoplastic Diseases of the Blood, 5th ed.; Springer: New York, NY, USA, 2013. [Google Scholar]
  113. Kelly, L.M.; Gilliland, D.G. GENETICS OF MYELOID LEUKEMIAS. Annu. Rev. Genom. Hum. Genet. 2002, 3, 179–198. [Google Scholar] [CrossRef]
  114. Takahashi, S. Current findings for recurring mutations in acute myeloid leukemia. J. Hematol. Oncol. 2011, 4, 36. [Google Scholar] [CrossRef] [Green Version]
  115. Chen, S.-J.; Shen, Y.; Chen, Z. A panoramic view of acute myeloid leukemia. Nat. Genet. 2013, 45, 586–587. [Google Scholar] [CrossRef]
  116. Pourrajab, F.; Zare-Khormizi, M.R.; Hashemi, A.S.; Hekmatimoghaddam, S. Genetic Characterization and Risk Stratification of Acute Myeloid Leukemia. Cancer Manag. Res. 2020, 12, 2231–2253. [Google Scholar] [CrossRef] [Green Version]
  117. Panuzzo, C.; Signorino, E.; Calabrese, C.; Ali, M.S.; Petiti, J.; Bracco, E.; Cilloni, D. Landscape of Tumor Suppressor Mutations in Acute Myeloid Leukemia. J. Clin. Med. 2020, 9, 802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Pei, S.; Minhajuddin, M.; Adane, B.; Khan, N.; Stevens, B.M.; Mack, S.C.; Lai, S.; Rich, J.N.; Inguva, A.; Shannon, K.M.; et al. AMPK/FIS1-Mediated Mitophagy Is Required for Self-Renewal of Human AML Stem Cells. Cell Stem Cell 2018, 23, 86–100.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Dany, M.; Gencer, S.; Nganga, R.; Thomas, R.J.; Oleinik, N.; Baron, K.D.; Szulc, Z.M.; Ruvolo, P.; Kornblau, S.; Andreeff, M.; et al. Targeting FLT3-ITD signaling mediates ceramide-dependent mitophagy and attenuates drug resistance in AML. Blood 2016, 128, 1944–1958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Aoyagi, Y.; Hayashi, Y.; Harada, Y.; Choi, K.; Matsunuma, N.; Sadato, D.; Maemoto, Y.; Ito, A.; Yanagi, S.; Starczynowski, D.T.; et al. Mitochondrial Fragmentation Triggers Ineffective Hematopoiesis in Myelodysplastic Syndromes. Cancer Discov. 2021, candisc.0032.2021. [Google Scholar] [CrossRef]
  121. Zheng, Q.; Zhao, Y.; Guo, J.; Zhao, S.; Fei, C.; Xiao, C.; Wu, D.; Wu, L.; Li, X.; Chang, C. Iron overload promotes mitochondrial fragmentation in mesenchymal stromal cells from myelodysplastic syndrome patients through activation of the AMPK/MFF/Drp1 pathway. Cell Death Dis. 2018, 9, 515. [Google Scholar] [CrossRef] [Green Version]
  122. Marlein, C.R.; Zaitseva, L.; Piddock, R.E.; Robinson, S.D.; Edwards, D.R.; Shafat, M.S.; Zhou, Z.; Lawes, M.; Bowles, K.M.; Rushworth, S. NADPH oxidase-2 derived superoxide drives mitochondrial transfer from bone marrow stromal cells to leukemic blasts. Blood 2017, 130, 1649–1660. [Google Scholar] [CrossRef] [PubMed]
  123. Moschoi, R.; Imbert, V.; Nebout, M.; Chiche, J.; Mary, D.; Prebet, T.; Saland, E.; Castellano, R.; Pouyet, L.; Collette, Y.; et al. Protective mitochondrial transfer from bone marrow stromal cells to acute myeloid leukemic cells during chemotherapy. Blood 2016, 128, 253–264. [Google Scholar] [CrossRef] [Green Version]
  124. Horvath, S.; Raj, K. DNA methylation-based biomarkers and the epigenetic clock theory of ageing. Nat. Rev. Genet. 2018, 19, 371–384. [Google Scholar] [CrossRef] [PubMed]
  125. Solary, E.; Bernard, O.; Tefferi, A.A.; Fuks, F.; Vainchenker, W. The Ten-Eleven Translocation-2 (TET2) gene in hematopoiesis and hematopoietic diseases. Leukemia 2014, 28, 485–496. [Google Scholar] [CrossRef] [PubMed]
  126. Shayevitch, R.; Askayo, D.; Keydar, I.; Ast, G. The importance of DNA methylation of exons on alternative splicing. RNA 2018, 24, 1351–1362. [Google Scholar] [CrossRef] [Green Version]
  127. Jones, P.A. Functions of DNA methylation: Islands, start sites, gene bodies and beyond. Nat. Rev. Genet. 2012, 13, 484–492. [Google Scholar] [CrossRef]
  128. Horvath, S. DNA methylation age of human tissues and cell types. Genome Biol. 2013, 14, R115. [Google Scholar] [CrossRef] [Green Version]
  129. Dou, X.; Boyd-Kirkup, J.D.; McDermott, J.; Zhang, X.; Li, F.; Rong, B.; Zhang, R.; Miao, B.; Chen, P.; Cheng, H.; et al. The strand-biased mitochondrial DNA methylome and its regulation by DNMT3A. Genome Res. 2019, 29, 1622–1634. [Google Scholar] [CrossRef]
  130. Prasad, G.R.; Jho, E.-H. A concise review of human brain methylome during aging and neurodegenerative diseases. BMB Rep. 2019, 52, 577–588. [Google Scholar] [CrossRef] [PubMed]
  131. Zheng, Y.; Joyce, B.T.; Liu, L.; Zhang, Z.; Kibbe, W.A.; Zhang, W.; Hou, L. Prediction of genome-wide DNA methylation in repetitive elements. Nucleic Acids Res. 2017, 45, 8697–8711. [Google Scholar] [CrossRef] [Green Version]
  132. Atilano, S.R.; Malik, D.; Chwa, M.; Cáceres-Del-Carpio, J.; Nesburn, A.B.; Boyer, D.S.; Kuppermann, B.D.; Jazwinski, S.M.; Miceli, M.V.; Wallace, D.C.; et al. Mitochondrial DNA variants can mediate methylation status of inflammation, angiogenesis and signaling genes. Hum. Mol. Genet. 2015, 24, 4491–4503. [Google Scholar] [CrossRef] [Green Version]
  133. De Paepe, B. How mitochondrial DNA-driven changes to chromosomal DNA methylation add a layer of complexity to mitochondrial disease. Epigenomics 2019, 11, 1749–1751. [Google Scholar] [CrossRef]
  134. Cortés-Pereira, E.; Fernández-Tajes, J.; Fernandez-Moreno, M.; Mosquera, M.E.V.; Relaño, S.; Ramos-Louro, P.; Durán-Sotuela, A.; Dalmao-Fernández, A.; Oreiro, N.; Blanco, F.J.; et al. Differential Association of Mitochondrial DNA Haplogroups J and H with the Methylation Status of Articular Cartilage: Potential Role in Apoptosis and Metabolic and Developmental Processes. Arthritis Rheumatol. 2019, 71, 1191–1200. [Google Scholar] [CrossRef]
  135. Matilainen, O.; Quiros, P.M.; Auwerx, J. Mitochondria and Epigenetics—Crosstalk in Homeostasis and Stress. Trends Cell Biol. 2017, 27, 453–463. [Google Scholar] [CrossRef]
  136. Bellizzi, D.; D’Aquila, P.; Giordano, M.; Montesanto, A.; Passarino, G. Global DNA methylation levels are modulated by mitochondrial DNA variants. Epigenomics 2012, 4, 17–27. [Google Scholar] [CrossRef] [Green Version]
  137. Iyer, L.M.; Tahiliani, M.; Rao, A.; Aravind, L. Prediction of novel families of enzymes involved in oxidative and other complex modifications of bases in nucleic acids. Cell Cycle 2009, 8, 1698–1710. [Google Scholar] [CrossRef] [Green Version]
  138. Laukka, T.; Mariani, C.J.; Ihantola, T.; Cao, J.Z.; Hokkanen, J.; Kaelin, W.G., Jr.; Godley, L.A.; Koivunen, P. Fumarate and Succinate Regulate Expression of Hypoxia-inducible Genes via TET Enzymes. J. Biol. Chem. 2016, 291, 4256–4265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Maddocks, O.; Labuschagne, C.F.; Adams, P.D.; Vousden, K.H. Serine Metabolism Supports the Methionine Cycle and DNA/RNA Methylation through De Novo ATP Synthesis in Cancer Cells. Mol. Cell 2016, 61, 210–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Zhang, N. Role of methionine on epigenetic modification of DNA methylation and gene expression in animals. Anim. Nutr. 2018, 4, 11–16. [Google Scholar] [CrossRef] [PubMed]
  141. Shock, L.S.; Thakkar, P.V.; Peterson, E.J.; Moran, R.G.; Taylor, S.M. DNA methyltransferase 1, cytosine methylation, and cytosine hydroxymethylation in mammalian mitochondria. Proc. Natl. Acad. Sci. USA 2011, 108, 3630–3635. [Google Scholar] [CrossRef] [Green Version]
  142. Lopes, A.F.C. Mitochondrial metabolism and DNA methylation: A review of the interaction between two genomes. Clin. Epigenetics 2020, 12, 182. [Google Scholar] [CrossRef]
  143. Dawid, I.B. 5-Methylcytidylic Acid: Absence from Mitochondrial DNA of Frogs and HeLa Cells. Science 1974, 184, 80–81. [Google Scholar] [CrossRef]
  144. Nass, M.M. Differential methylation of mitochondrial and nuclear DNA in cultured mouse, hamster and virus-transformed hamster cells In vivo and in vitro methylation. J. Mol. Biol. 1973, 80, 155–175. [Google Scholar] [CrossRef]
  145. Vanyushin, B.; Kirnos, M. Structure of animal mitochondrial DNA (base composition, pyrimidine clusters, character of methylation). Biochim. Biophys. Acta 1977, 475, 323–336. [Google Scholar] [CrossRef]
  146. Mechta, M.; Ingerslev, L.R.; Fabre, O.; Picard, M.; Barrès, R. Evidence Suggesting Absence of Mitochondrial DNA Methylation. Front. Genet. 2017, 8, 166. [Google Scholar] [CrossRef] [Green Version]
  147. Kopinski, P.K.; Janssen, K.; Schaefer, P.M.; Trefely, S.; Perry, C.E.; Potluri, P.; Tintos-Hernandez, J.A.; Singh, L.N.; Karch, K.; Campbell, S.L.; et al. Regulation of nuclear epigenome by mitochondrial DNA heteroplasmy. Proc. Natl. Acad. Sci. USA 2019, 116, 16028–16035. [Google Scholar] [CrossRef] [Green Version]
  148. Fetterman, J.L.; Ballinger, S.W. Mitochondrial genetics regulate nuclear gene expression through metabolites. Proc. Natl. Acad. Sci. USA 2019, 116, 15763–15765. [Google Scholar] [CrossRef] [Green Version]
  149. Scola, G.; Kim, H.K.; Young, L.T.; Salvador, M.; Andreazza, A.C. Lithium reduces the effects of rotenone-induced complex I dysfunction on DNA methylation and hydroxymethylation in rat cortical primary neurons. Psychopharmacology 2014, 231, 4189–4198. [Google Scholar] [CrossRef] [PubMed]
  150. Lozoya, O.A.; Xu, F.; Grenet, D.; Wang, T.; Grimm, S.A.; Godfrey, V.; Waidyanatha, S.; Woychik, R.P.; Santos, J.H. Single Nucleotide Resolution Analysis Reveals Pervasive, Long-Lasting DNA Methylation Changes by Developmental Exposure to a Mitochondrial Toxicant. Cell Rep. 2020, 32, 108131. [Google Scholar] [CrossRef] [PubMed]
  151. Feeley, K.P.; Bray, A.W.; Westbrook, D.G.; Johnson, L.W.; Kesterson, R.A.; Ballinger, S.W.; Welch, D.R. Mitochondrial Genetics Regulate Breast Cancer Tumorigenicity and Metastatic Potential. Cancer Res. 2015, 75, 4429–4436. [Google Scholar] [CrossRef] [Green Version]
  152. San José-Enériz, E.; Gimenez-Camino, N.; Agirre, X.; Prosper, F. HDAC Inhibitors in Acute Myeloid Leukemia. Cancers 2019, 11, 1794. [Google Scholar] [CrossRef] [Green Version]
  153. Yu, X.; Li, H.; Hu, P.; Qing, Y.; Wang, X.; Zhu, M.; Wang, H.; Wang, Z.; Xu, J.; Guo, Q.; et al. Natural HDAC-1/8 inhibitor baicalein exerts therapeutic effect in CBF-AML. Clin. Transl. Med. 2020, 10, e154. [Google Scholar] [CrossRef] [PubMed]
  154. Yang, J.; He, J.; Ismail, M.; Tweeten, S.; Zeng, F.; Gao, L.; Ballinger, S.; Young, M.; Prabhu, S.D.; Rowe, G.; et al. HDAC inhibition induces autophagy and mitochondrial biogenesis to maintain mitochondrial homeostasis during cardiac ischemia/reperfusion injury. J. Mol. Cell. Cardiol. 2019, 130, 36–48. [Google Scholar] [CrossRef] [PubMed]
  155. Testa, U.; Castelli, G.; Pelosi, E. Isocitrate Dehydrogenase Mutations in Myelodysplastic Syndromes and in Acute Myeloid Leukemias. Cancers 2020, 12, 2427. [Google Scholar] [CrossRef] [PubMed]
  156. Haferlach, T.; Nagata, Y.; Grossmann, V.; Okuno, Y.; Bacher, U.; Nagae, G.; Schnittger, S.; Sanada, M.; Kon, A.; Alpermann, T.; et al. Landscape of genetic lesions in 944 patients with myelodysplastic syndromes. Leukemia 2014, 28, 241–247. [Google Scholar] [CrossRef] [Green Version]
  157. Molenaar, R.J.; Thota, S.; Nagata, Y.; Patel, B.B.; Clemente, M.J.; Przychodzen, B.; Hirsh, C.; Viny, A.; Hosano, N.; Bleeker, F.E.; et al. Clinical and biological implications of ancestral and non-ancestral IDH1 and IDH2 mutations in myeloid neoplasms. Leukemia 2015, 29, 2134–2142. [Google Scholar] [CrossRef] [PubMed]
  158. Dinardo, C.D.; Jabbour, E.; Ravandi, F.; Takahashi, K.; Daver, N.; Routbort, M.J.; Patel, K.P.; Brandt, M.L.; Pierce, S.; Kantarjian, H.M.; et al. IDH1 and IDH2 mutations in myelodysplastic syndromes and role in disease progression. Leukemia 2016, 30, 980–984. [Google Scholar] [CrossRef] [Green Version]
  159. Medeiros, B.C.; Fathi, A.T.; Dinardo, C.D.; Pollyea, D.A.; Chan, S.M.; Swords, R. Isocitrate dehydrogenase mutations in myeloid malignancies. Leukemia 2017, 31, 272–281. [Google Scholar] [CrossRef]
  160. Patnaik, M.M.; Hanson, C.A.; Hodnefield, J.M.; Lasho, T.L.; Finke, C.M.; Knudson, R.A.; Ketterling, R.; Pardanani, A.; Tefferi, A. Differential prognostic effect of IDH1 versus IDH2 mutations in myelodysplastic syndromes: A Mayo Clinic Study of 277 patients. Leukemia 2012, 26, 101–105. [Google Scholar] [CrossRef] [PubMed]
  161. Wang, N.; Wang, F.; Shan, N.; Sui, X.; Xu, H. IDH1 Mutation Is an Independent Inferior Prognostic Indicator for Patients with Myelodysplastic Syndromes. Acta Haematol. 2017, 138, 143–151. [Google Scholar] [CrossRef]
  162. Pellagatti, A.; Roy, S.; Di Genua, C.; Burns, A.; McGraw, K.; Valletta, S.; Larrayoz, M.J.; Fernandez-Mercado, M.; Mason, J.; Killick, S.; et al. Targeted resequencing analysis of 31 genes commonly mutated in myeloid disorders in serial samples from myelodysplastic syndrome patients showing disease progression. Leukemia 2016, 30, 248–250. [Google Scholar] [CrossRef]
  163. Marcucci, G.; Maharry, K.; Wu, Y.-Z.; Radmacher, M.D.; Mrózek, K.; Margeson, D.; Holland, K.B.; Whitman, S.P.; Becker, H.; Schwind, S.; et al. IDH1andIDH2Gene Mutations Identify Novel Molecular Subsets within De Novo Cytogenetically Normal Acute Myeloid Leukemia: A Cancer and Leukemia Group B Study. J. Clin. Oncol. 2010, 28, 2348–2355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Voso, M.T.; Fabiani, E.; Fianchi, L.; Falconi, G.; Criscuolo, M.; Santangelo, R.; Chiusolo, P.; Betti, S.; Dalo, F.; Hohaus, S.; et al. Mutations of epigenetic regulators and of the spliceosome machinery in therapy-related myeloid neoplasms and in acute leukemias evolved from chronic myeloproliferative diseases. Leukemia 2013, 27, 982–985. [Google Scholar] [CrossRef]
  165. Singhal, D.; Wee, L.Y.A.; Kutyna, M.M.; Chhetri, R.; Geoghegan, J.; Schreiber, A.; Feng, J.; Wang, P.P.-S.; Babic, M.; Parker, W.T.; et al. The mutational burden of therapy-related myeloid neoplasms is similar to primary myelodysplastic syndrome but has a distinctive distribution. Leukemia 2019, 33, 2842–2853. [Google Scholar] [CrossRef]
  166. Ok, C.Y.; Patel, K.P.; Garcia-Manero, G.; Routbort, M.J.; Fu, B.; Tang, G.; Goswami, M.; Singh, R.; Kanagal-Shamanna, R.; Pierce, S.A.; et al. Mutational profiling of therapy-related myelodysplastic syndromes and acute myeloid leukemia by next generation sequencing, a comparison with de novo diseases. Leuk. Res. 2015, 39, 348–354. [Google Scholar] [CrossRef] [Green Version]
  167. Hartmann, L.; Nadarajah, N.; Meggendorfer, M.; Höllein, A.; Vetro, C.; Kern, W.; Haferlach, T.; Haferlach, C.; Stengel, A. Molecular characterization of a second myeloid neoplasm developing after treatment for acute myeloid leukemia. Leukemia 2020, 34, 811–820. [Google Scholar] [CrossRef] [PubMed]
  168. Kuzmanovic, T.; Patel, B.J.; Sanikommu, S.R.; Nagata, Y.; Awada, H.; Kerr, C.M.; Przychodzen, B.P.; Jha, B.K.; Hiwase, D.; Singhal, D.; et al. Genomics of therapy-related myeloid neoplasms. Haematologica 2020, 105, e98–e101. [Google Scholar] [CrossRef]
  169. Figueroa, M.E.; Abdel-Wahab, O.; Lu, C.; Ward, P.S.; Patel, J.; Shih, A.; Li, Y.; Bhagwat, N.; VasanthaKumar, A.; Fernandez, H.F.; et al. Leukemic IDH1 and IDH2 Mutations Result in a Hypermethylation Phenotype, Disrupt TET2 Function, and Impair Hematopoietic Differentiation. Cancer Cell 2010, 18, 553–567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Lu, C.; Ward, P.S.; Kapoor, G.S.; Rohle, D.; Turcan, S.; Abdel-Wahab, O.; Edwards, C.R.; Khanin, R.; Figueroa, M.E.; Melnick, A.; et al. IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 2012, 483, 474–478. [Google Scholar] [CrossRef] [Green Version]
  171. Chan, S.M.; Thomas, D.; Corces-Zimmerman, M.R.; Xavy, S.; Rastogi, S.; Hong, W.-J.; Zhao, F.; Medeiros, B.C.; Tyvoll, D.A.; Majeti, R. Isocitrate dehydrogenase 1 and 2 mutations induce BCL-2 dependence in acute myeloid leukemia. Nat. Med. 2015, 21, 178–184. [Google Scholar] [CrossRef] [Green Version]
  172. Yen, K.; Travins, J.; Wang, F.; David, M.; Artin, E.; Straley, K.; Padyana, A.; Gross, S.; DelaBarre, B.; Tobin, E.; et al. AG-221, a First-in-Class Therapy Targeting Acute Myeloid Leukemia Harboring Oncogenic IDH2 Mutations. Cancer Discov. 2017, 7, 478–493. [Google Scholar] [CrossRef] [Green Version]
  173. Stein, E.M.; Dinardo, C.D.; Pollyea, D.A.; Fathi, A.T.; Roboz, G.J.; Altman, J.K.; Stone, R.M.; DeAngelo, D.J.; Levine, R.L.; Flinn, I.W.; et al. Enasidenib in mutant IDH2 relapsed or refractory acute myeloid leukemia. Blood 2017, 130, 722–731. [Google Scholar] [CrossRef]
  174. Stein, E.M.; Dinardo, C.D.; Fathi, A.T.; Pollyea, D.A.; Stone, R.M.; Altman, J.K.; Roboz, G.J.; Patel, M.R.; Collins, R.; Flinn, I.W.; et al. Molecular remission and response patterns in patients with mutant-IDH2 acute myeloid leukemia treated with enasidenib. Blood 2019, 133, 676–687. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Pollyea, D.A.; Tallman, M.S.; De Botton, S.; Kantarjian, H.M.; Collins, R.; Stein, A.S.; Frattini, M.G.; Xu, Q.; Tosolini, A.; See, W.L.; et al. Enasidenib, an inhibitor of mutant IDH2 proteins, induces durable remissions in older patients with newly diagnosed acute myeloid leukemia. Leukemia 2019, 33, 2575–2584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Stein, E.M.; Fathi, A.T.; DiNardo, C.D.; Pollyea, D.A.; Roboz, G.J.; Collins, R.; Sekeres, M.A.; Stone, R.M.; Attar, E.C.; Frattini, M.G.; et al. Enasidenib in patients with mutant IDH2 myelodysplastic syndromes: A phase 1 subgroup analysis of the multicentre, AG221-C-001 trial. Lancet Haematol. 2020, 7, e309–e319. [Google Scholar] [CrossRef]
  177. Richard-Carpentier, G.; DeZern, A.E.; Takahashi, K.; Konopleva, M.Y.; Loghavi, S.; Masarova, L.; Alvarado, Y.; Ravandi, F.; Bravo, G.M.; Naqvi, K.; et al. Preliminary Results from the Phase II Study of the IDH2-Inhibitor Enasidenib in Patients with High-Risk IDH2-Mutated Myelodysplastic Syndromes (MDS). Blood 2019, 134 (Suppl. 1), 678. [Google Scholar] [CrossRef]
  178. Wang, Y.; Agarwal, E.; Bertolini, I.; Ghosh, J.C.; Seo, J.H.; Altieri, D.C. IDH2 reprograms mitochondrial dynamics in cancer through a HIF-1α–regulated pseudohypoxic state. FASEB J. 2019, 33, 13398–13411. [Google Scholar] [CrossRef] [Green Version]
  179. Tahiliani, M.; Koh, K.P.; Shen, Y.; Pastor, W.A.; Bandukwala, H.; Brudno, Y.; Agarwal, S.; Iyer, L.M.; Liu, D.R.; Aravind, L.; et al. Conversion of 5-Methylcytosine to 5-Hydroxymethylcytosine in Mammalian DNA by MLL Partner TET1. Science 2009, 324, 930–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Kaelin, W.G.; McKnight, S.L. Influence of Metabolism on Epigenetics and Disease. Cell 2013, 153, 56–69. [Google Scholar] [CrossRef] [Green Version]
  181. Yin, R.; Mao, S.-Q.; Zhao, B.; Chong, Z.; Yang, Y.; Zhao, C.; Zhang, D.; Huang, H.; Gao, J.; Li, Z.; et al. Ascorbic Acid Enhances Tet-Mediated 5-Methylcytosine Oxidation and Promotes DNA Demethylation in Mammals. J. Am. Chem. Soc. 2013, 135, 10396–10403. [Google Scholar] [CrossRef]
  182. Blaschke, K.; Ebata, K.; Karimi, M.M.; Zepeda-Martínez, J.A.; Goyal, P.; Mahapatra, S.; Tam, A.; Laird, D.J.; Hirst, M.; Rao, A.; et al. Vitamin C induces Tet-dependent DNA demethylation and a blastocyst-like state in ES cells. Nature 2013, 500, 222–226. [Google Scholar] [CrossRef]
  183. Coulter, J.B.; O’Driscoll, C.M.; Bressler, J.P. Hydroquinone Increases 5-Hydroxymethylcytosine Formation through Ten Eleven Translocation 1 (TET1) 5-Methylcytosine Dioxygenase. J. Biol. Chem. 2013, 288, 28792–28800. [Google Scholar] [CrossRef] [Green Version]
  184. Dzitoyeva, S.; Chen, H.; Manev, H. Effect of aging on 5-hydroxymethylcytosine in brain mitochondria. Neurobiol. Aging 2012, 33, 2881–2891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Ji, F.; Zhao, C.; Wang, B.; Tang, Y.; Miao, Z.; Wang, Y. The role of 5-hydroxymethylcytosine in mitochondria after ischemic stroke. J. Neurosci. Res. 2018, 96, 1717–1726. [Google Scholar] [CrossRef]
  186. Guillamot, M.; Cimmino, L.; Aifantis, I. The Impact of DNA Methylation in Hematopoietic Malignancies. Trends Cancer 2016, 2, 70–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Moran-Crusio, K.; Reavie, L.; Shih, A.; Abdel-Wahab, O.; Ndiaye-Lobry, D.; Lobry, C.; Figueroa, M.E.; Vasanthakumar, A.; Patel, J.; Zhao, X.; et al. Tet2 Loss Leads to Increased Hematopoietic Stem Cell Self-Renewal and Myeloid Transformation. Cancer Cell 2011, 20, 11–24. [Google Scholar] [CrossRef] [Green Version]
  188. Quivoron, C.; Couronné, L.; Della Valle, V.; Lopez, C.K.; Plo, I.; Wagner-Ballon, O.; Cruzeiro, M.D.; Delhommeau, F.; Arnulf, B.; Stern, M.-H.; et al. TET2 Inactivation Results in Pleiotropic Hematopoietic Abnormalities in Mouse and Is a Recurrent Event during Human Lymphomagenesis. Cancer Cell 2011, 20, 25–38. [Google Scholar] [CrossRef] [Green Version]
  189. Cimmino, L.; Dolgalev, I.; Wang, Y.; Yoshimi, A.; Martin, G.H.; Wang, J.; Ng, V.; Xia, B.; Witkowski, M.T.; Mitchell-Flack, M.; et al. Restoration of TET2 Function Blocks Aberrant Self-Renewal and Leukemia Progression. Cell 2017, 170, 1079–1095.e20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Song, S.J.; Ito, K.; Ala, U.; Kats, L.; Webster, K.; Sun, S.M.; Jongen-Lavrencic, M.; Manova-Todorova, K.; Teruya-Feldstein, J.; Avigan, D.E.; et al. The Oncogenic MicroRNA miR-22 Targets the TET2 Tumor Suppressor to Promote Hematopoietic Stem Cell Self-Renewal and Transformation. Cell Stem Cell 2013, 13, 87–101. [Google Scholar] [CrossRef] [Green Version]
  191. Wang, Y.; Zhang, Y. Regulation of TET Protein Stability by Calpains. Cell Rep. 2014, 6, 278–284. [Google Scholar] [CrossRef] [Green Version]
  192. Ko, M.; Bandukwala, H.S.; An, J.; Lamperti, E.D.; Thompson, E.C.; Hastie, R.; Tsangaratou, A.; Rajewsky, K.; Koralov, S.B.; Rao, A. Ten-Eleven-Translocation 2 (TET2) negatively regulates homeostasis and differentiation of hematopoietic stem cells in mice. Proc. Natl. Acad. Sci. USA 2011, 108, 14566–14571. [Google Scholar] [CrossRef] [Green Version]
  193. Li, Z.; Cai, X.; Cai, C.-L.; Wang, J.; Zhang, W.; Petersen, B.E.; Yang, F.-C.; Xu, M. Deletion of Tet2 in mice leads to dysregulated hematopoietic stem cells and subsequent development of myeloid malignancies. Blood 2011, 118, 4509–4518. [Google Scholar] [CrossRef] [Green Version]
  194. Bejar, R.; Stevenson, K.; Abdel-Wahab, O.; Galili, N.; Nilsson, B.; Garcia-Manero, G.; Kantarjian, H.; Raza, A.; Levine, R.L.; Neuberg, D.; et al. Clinical Effect of Point Mutations in Myelodysplastic Syndromes. N. Engl. J. Med. 2011, 364, 2496–2506. [Google Scholar] [CrossRef] [Green Version]
  195. Kosmider, O.; Gelsi-Boyer, V.; Cheok, M.; Grabar, S.; Della-Valle, V.; Picard, F.; Viguié, F.; Quesnel, B.; Beyne-Rauzy, O.; Solary, E.; et al. TET2 mutation is an independent favorable prognostic factor in myelodysplastic syndromes (MDSs). Blood 2009, 114, 3285–3291. [Google Scholar] [CrossRef] [Green Version]
  196. Itzykson, R.; Kosmider, O.; Cluzeau, T.; Mansat-De Mas, V.; Dreyfus, F.; Beyne-Rauzy, O.; Quesnel, B.; Vey, N.; Gelsi-Boyer, V.; Raynaud, S.; et al. Impact of TET2 mutations on response rate to azacitidine in myelodysplastic syndromes and low blast count acute myeloid leukemias. Leukemia 2011, 25, 1147–1152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Tefferi, A.; Pardanani, A.; Lim, K.-H.; Abdel-Wahab, O.; Lasho, T.L.; Patel, J.; Gangat, N.; Finke, C.M.; Schwager, S.; Mullally, A.; et al. TET2 mutations and their clinical correlates in polycythemia vera, essential thrombocythemia and myelofibrosis. Leukemia 2009, 23, 905–911. [Google Scholar] [CrossRef] [Green Version]
  198. Nibourel, O.; Kosmider, O.; Cheok, M.; Boissel, N.; Renneville, A.; Philippe, N.; Dombret, H.; Dreyfus, F.; Quesnel, B.; Geffroy, S.; et al. Incidence and prognostic value of TET2 alterations in de novo acute myeloid leukemia achieving complete remission. Blood 2010, 116, 1132–1135. [Google Scholar] [CrossRef]
  199. Weissmann, S.; Alpermann, T.; Grossmann, V.; Kowarsch, A.; Nadarajah, N.; Eder, C.; Dicker, F.; Fasan, A.; Haferlach, C.; Kern, W.; et al. Landscape of TET2 mutations in acute myeloid leukemia. Leukemia 2012, 26, 934–942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Konstandin, N.; Bultmann, S.; Szwagierczak, A.; Dufour, A.; Ksienzyk, B.; Schneider, F.; Herold, T.; Mulaw, M.A.; Kakadia, P.M.; Spiekermann, K.; et al. Genomic 5-hydroxymethylcytosine levels correlate with TET2 mutations and a distinct global gene expression pattern in secondary acute myeloid leukemia. Leukemia 2011, 25, 1649–1652. [Google Scholar] [CrossRef] [Green Version]
  201. Chou, W.-C.; Chou, S.-C.; Liu, C.-Y.; Chen, C.-Y.; Hou, H.-A.; Kuo, Y.-Y.; Lee, M.-C.; Ko, B.-S.; Tang, J.-L.; Yao, M.; et al. TET2 mutation is an unfavorable prognostic factor in acute myeloid leukemia patients with intermediate-risk cytogenetics. Blood 2011, 118, 3803–3810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Delhommeau, F.; Dupont, S.; Della Valle, V.; James, C.; Trannoy, S.; Massé, A.; Kosmider, O.; Le Couedic, J.-P.; Robert, F.; Alberdi, A.; et al. Mutation inTET2in Myeloid Cancers. N. Engl. J. Med. 2009, 360, 2289–2301. [Google Scholar] [CrossRef]
  203. Langemeijer, S.M.C.; Kuiper, R.P.; Berends, M.; Knops, R.; Aslanyan, M.G.; Massop, M.; Stevens-Linders, E.; Van Hoogen, P.; Van Kessel, A.G.; Raymakers, R.A.P.; et al. Acquired mutations in TET2 are common in myelodysplastic syndromes. Nat. Genet. 2009, 41, 838–842. [Google Scholar] [CrossRef]
  204. Tefferi, A.; Lim, K.-H.; Abdel-Wahab, O.; Lasho, T.L.; Patel, J.; Patnaik, M.M.; Hanson, C.A.; Pardanani, A.; Gilliland, D.G.; Levine, R.L. Detection of mutant TET2 in myeloid malignancies other than myeloproliferative neoplasms: CMML, MDS, MDS/MPN and AML. Leukemia 2009, 23, 1343–1345. [Google Scholar] [CrossRef] [Green Version]
  205. Vannucchi, A.M.; Lasho, T.L.; Guglielmelli, P.; Biamonte, F.; Pardanani, A.; Pereira, A.; Finke, C.; Score, J.; Gangat, N.; Mannarelli, C.; et al. Mutations and prognosis in primary myelofibrosis. Leukemia 2013, 27, 1861–1869. [Google Scholar] [CrossRef] [PubMed]
  206. Beer, P.A.; Delhommeau, F.; LeCouédic, J.-P.; Dawson, M.A.; Chen, E.; Bareford, D.; Kušec, R.; McMullin, M.F.; Harrison, C.N.; Vannucchi, A.M.; et al. Two routes to leukemic transformation after a JAK2 mutation–positive myeloproliferative neoplasm. Blood 2010, 115, 2891–2900. [Google Scholar] [CrossRef] [Green Version]
  207. Schaub, F.X.; Looser, R.; Li, S.; Hao-Shen, H.; Lehmann, T.; Tichelli, A.; Skoda, R.C. Clonal analysis of TET2 and JAK2 mutations suggests that TET2 can be a late event in the progression of myeloproliferative neoplasms. Blood 2010, 115, 2003–2007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Roche-Lestienne, C.; Marceau, A.; Labis, E.; Nibourel, O.; Coiteux, V.; Guilhot, J.; Legros, L.; Nicolini, F.; Rousselot, P.; Gardembas, M.; et al. Mutation analysis of TET2, IDH1, IDH2 and ASXL1 in chronic myeloid leukemia. Leukemia 2011, 25, 1661–1664. [Google Scholar] [CrossRef] [Green Version]
  209. Tefferi, A.; Levine, R.L.; Lim, K.-H.; Abdel-Wahab, O.; Lasho, T.L.; Patel, J.; Finke, C.M.; Mullally, A.; Li, C.-Y.; Pardanani, A.; et al. Frequent TET2 mutations in systemic mastocytosis: Clinical, KITD816V and FIP1L1-PDGFRA correlates. Leukemia 2009, 23, 900–904. [Google Scholar] [CrossRef] [Green Version]
  210. Soucie, E.; Hanssens, K.; Mercher, T.; Georgin-Lavialle, S.; Damaj, G.; Livideanu, C.; Chandesris, M.O.; Acin, Y.; Létard, S.; De Sepulveda, P.; et al. In aggressive forms of mastocytosis, TET2 loss cooperates with c-KITD816V to transform mast cells. Blood 2012, 120, 4846–4849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Jardin, F.; Ruminy, P.; Parmentier, F.; Troussard, X.; Vaida, I.; Stamatoullas, A.; Leprêtre, S.; Penther, D.; Duval, A.B.; Picquenot, J.-M.; et al. TET2 and TP53 mutations are frequently observed in blastic plasmacytoid dendritic cell neoplasm. Br. J. Haematol. 2011, 153, 413–416. [Google Scholar] [CrossRef]
  212. Ward, P.S.; Patel, J.; Wise, D.R.; Abdel-Wahab, O.; Bennett, B.D.; Coller, H.A.; Cross, J.R.; Fantin, V.R.; Hedvat, C.V.; Perl, A.E.; et al. The Common Feature of Leukemia-Associated IDH1 and IDH2 Mutations Is a Neomorphic Enzyme Activity Converting α-Ketoglutarate to 2-Hydroxyglutarate. Cancer Cell 2010, 17, 225–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Lemonnier, F.; Couronné, L.; Parrens, M.; Jaïs, J.-P.; Travert, M.; Lamant, L.; Tournillac, O.; Rousset, T.; Fabiani, B.; Cairns, R.A.; et al. Recurrent TET2 mutations in peripheral T-cell lymphomas correlate with TFH-like features and adverse clinical parameters. Blood 2012, 120, 1466–1469. [Google Scholar] [CrossRef] [Green Version]
  214. Meissner, B.; Kridel, R.; Lim, R.S.; Rogic, S.; Tse, K.; Scott, D.W.; Moore, R.; Mungall, A.; Marra, M.A.; Connors, J.M.; et al. The E3 ubiquitin ligase UBR5 is recurrently mutated in mantle cell lymphoma. Blood 2013, 121, 3161–3164. [Google Scholar] [CrossRef] [Green Version]
  215. Asmar, F.; Punj, V.; Christensen, J.; Pedersen, M.T.; Pedersen, A.; Nielsen, A.B.; Hother, C.; Ralfkiaer, U.; Brown, P.; Ralfkiaer, E.; et al. Genome-wide profiling identifies a DNA methylation signature that associates with TET2 mutations in diffuse large B-cell lymphoma. Haematologica 2013, 98, 1912–1920. [Google Scholar] [CrossRef] [Green Version]
  216. Couronné, L.; Bastard, C.; Bernard, O.A. TET2andDNMT3AMutations in Human T-Cell Lymphoma. N. Engl. J. Med. 2012, 366, 95–96. [Google Scholar] [CrossRef] [PubMed]
  217. Cairns, R.A.; Iqbal, J.; Lemonnier, F.; Kucuk, C.; de Leval, L.; Jais, J.-P.; Parrens, M.; Martin, A.; Xerri, L.; Brousset, P.; et al. IDH2 mutations are frequent in angioimmunoblastic T-cell lymphoma. Blood 2012, 119, 1901–1903. [Google Scholar] [CrossRef] [Green Version]
  218. Nickerson, M.L.; Im, K.M.; Misner, K.J.; Tan, W.; Lou, H.; Gold, B.; Wells, D.W.; Bravo, H.C.; Fredrikson, K.; Harkins, T.T.; et al. Somatic alterations contributing to metastasis of a castration-resistant prostate cancer. Hum. Mutat. 2013, 34, 1231–1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Itzykson, R.; Kosmider, O.; Renneville, A.; Gelsi-Boyer, V.; Meggendorfer, M.; Morabito, M.; Berthon, C.; Adès, L.; Fenaux, P.; Beyne-Rauzy, O.; et al. Prognostic Score Including Gene Mutations in Chronic Myelomonocytic Leukemia. J. Clin. Oncol. 2013, 31, 2428–2436. [Google Scholar] [CrossRef] [PubMed]
  220. Itzykson, R.; Kosmider, O.; Renneville, A.; Morabito, M.; Preudhomme, C.; Berthon, C.; Adès, L.; Fenaux, P.; Platzbecker, U.; Gagey, O.; et al. Clonal architecture of chronic myelomonocytic leukemias. Blood 2013, 121, 2186–2198. [Google Scholar] [CrossRef]
  221. Pollyea, D.A.; Raval, A.; Kusler, B.; Gotlib, J.R.; Alizadeh, A.A.; Mitchell, B.S. Impact of TET2 mutations on mRNA expression and clinical outcomes in MDS patients treated with DNA methyltransferase inhibitors. Hematol. Oncol. 2011, 29, 157–160. [Google Scholar] [CrossRef]
  222. Voso, M.T.; Fabiani, E.; Piciocchi, A.; Matteucci, C.; Brandimarte, L.; Finelli, C.; Pogliani, E.; Angelucci, E.; Fioritoni, G.; Musto, P.; et al. Role of BCL2L10 methylation and TET2 mutations in higher risk myelodysplastic syndromes treated with 5-Azacytidine. Leukemia 2011, 25, 1910–1913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Braun, T.; Itzykson, R.; Renneville, A.; de Renzis, B.; Dreyfus, F.; Laribi, K.; Bouabdallah, K.; Vey, N.; Toma, A.; Recher, C.; et al. Molecular predictors of response to decitabine in advanced chronic myelomonocytic leukemia: A phase 2 trial. Blood 2011, 118, 3824–3831. [Google Scholar] [CrossRef] [Green Version]
  224. Smith, Z.D.; Meissner, A. DNA methylation: Roles in mammalian development. Nat. Rev. Genet. 2013, 14, 204–220. [Google Scholar] [CrossRef]
  225. Gill, H.; Leung, A.Y.H.; Kwong, Y.-L. Molecular and Cellular Mechanisms of Myelodysplastic Syndrome: Implications on Targeted Therapy. Int. J. Mol. Sci. 2016, 17, 440. [Google Scholar] [CrossRef] [Green Version]
  226. Chestnut, B.A.; Chang, Q.; Price, A.; Lesuisse, C.; Wong, M.; Martin, L.J. Epigenetic Regulation of Motor Neuron Cell Death through DNA Methylation. J. Neurosci. 2011, 31, 16619–16636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Wong, M.; Gertz, B.; Chestnut, B.A.; Martin, L.J. Mitochondrial DNMT3A and DNA methylation in skeletal muscle and CNS of transgenic mouse models of ALS. Front. Cell. Neurosci. 2013, 7, 279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Bellizzi, D.; D’Aquila, P.; Scafone, T.; Giordano, M.; Riso, V.; Riccio, A.; Passarino, G. The Control Region of Mitochondrial DNA Shows an Unusual CpG and Non-CpG Methylation Pattern. DNA Res. 2013, 20, 537–547. [Google Scholar] [CrossRef] [Green Version]
  229. Saini, S.K.; Mangalhara, K.C.; Prakasam, G.; Bamezai, R.N.K. DNA Methyltransferase1 (DNMT1) Isoform3 methylates mitochondrial genome and modulates its biology. Sci. Rep. 2017, 7, 1525. [Google Scholar] [CrossRef] [Green Version]
  230. Challen, G.A.; Sun, D.; Jeong, M.; Luo, M.; Jelinek, J.; Berg, J.; Bock, C.; Vasanthakumar, A.; Gu, H.; Xi, Y.; et al. Dnmt3a is essential for hematopoietic stem cell differentiation. Nat. Genet. 2011, 44, 23–31. [Google Scholar] [CrossRef] [Green Version]
  231. Koya, J.; Kataoka, K.; Sato, T.; Bando, M.; Kato, Y.; Tsuruta-Kishino, T.; Kobayashi, H.; Narukawa, K.; Miyoshi, H.; Shirahige, K.; et al. DNMT3A R882 mutants interact with polycomb proteins to block haematopoietic stem and leukaemic cell differentiation. Nat. Commun. 2016, 7, 10924. [Google Scholar] [CrossRef] [PubMed]
  232. Falini, B.; Sportoletti, P.; Brunetti, L.; Martelli, M.P. Perspectives for therapeutic targeting of gene mutations in acute myeloid leukaemia with normal cytogenetics. Br. J. Haematol. 2015, 170, 305–322. [Google Scholar] [CrossRef]
  233. Nebbioso, A.; Benedetti, R.; Conte, M.; Iside, C.; Altucci, L. Genetic mutations in epigenetic modifiers as therapeutic targets in acute myeloid leukemia. Expert Opin. Ther. Targets 2015, 19, 1187–1202. [Google Scholar] [CrossRef]
  234. Ley, T.J.; Ding, L.; Walter, M.J.; McLellan, M.D.; Lamprecht, T.L.; Larson, D.E.; Kandoth, C.; Payton, J.E.; Baty, J.; Welch, J.J.; et al. DNMT3AMutations in Acute Myeloid Leukemia. New Engl. J. Med. 2010, 363, 2424–2433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Shlush, L.I.; Zandi, S.; Mitchell, A.; Chen, W.C.; Brandwein, J.M.; Gupta, V.; Kennedy, J.A.; Schimmer, A.; Schuh, A.C.; Yee, K.W.; et al. Identification of pre-leukaemic haematopoietic stem cells in acute leukaemia. Nature 2014, 506, 328–333. [Google Scholar] [CrossRef]
  236. Xie, M.; Lu, C.; Wang, J.; McLellan, M.D.; Johnson, K.J.; Wendl, M.C.; McMichael, J.F.; Schmidt, H.K.; Yellapantula, V.; Miller, C.A.; et al. Age-related mutations associated with clonal hematopoietic expansion and malignancies. Nat. Med. 2014, 20, 1472–1478. [Google Scholar] [CrossRef] [PubMed]
  237. Ribeiro, A.F.T.; Pratcorona, M.; Erpelinck-Verschueren, C.; Rockova, V.; Sanders, M.; Abbas, S.; Figueroa, M.E.; Zeilemaker, A.; Melnick, A.; Löwenberg, B.; et al. Mutant DNMT3A: A marker of poor prognosis in acute myeloid leukemia. Blood 2012, 119, 5824–5831. [Google Scholar] [CrossRef]
  238. Walter, M.J.; Ding, L.; Shen, D.; Shao, J.; Grillot, M.; McLellan, M.; Fulton, R.; Schmidt, H.; Kalicki-Veizer, J.; O’Laughlin, M.; et al. Recurrent DNMT3A mutations in patients with myelodysplastic syndromes. Leukemia 2011, 25, 1153–1158. [Google Scholar] [CrossRef] [Green Version]
  239. Kolquist, K.A.; Schultz, R.A.; Furrow, A.; Brown, T.C.; Han, J.-Y.; Campbell, L.J.; Wall, M.; Slovak, M.L.; Shaffer, L.G.; Ballif, B.C. Microarray-based comparative genomic hybridization of cancer targets reveals novel, recurrent genetic aberrations in the myelodysplastic syndromes. Cancer Genet. 2011, 204, 603–628. [Google Scholar] [CrossRef]
  240. Jaiswal, S.; Fontanillas, P.; Flannick, J.; Manning, A.; Grauman, P.V.; Mar, B.G.; Lindsley, R.C.; Mermel, C.H.; Burtt, N.; Chavez, A.; et al. Age-Related Clonal Hematopoiesis Associated with Adverse Outcomes. N. Engl. J. Med. 2014, 371, 2488–2498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Traina, F.; Visconte, V.; Elson, P.; Tabarroki, A.; Jankowska, A.M.; Hasrouni, E.; Sugimoto, Y.; Szpurka, H.; Makishima, H.; O’Keefe, C.L.; et al. Impact of molecular mutations on treatment response to DNMT inhibitors in myelodysplasia and related neoplasms. Leukemia 2014, 28, 78–87. [Google Scholar] [CrossRef] [PubMed]
  242. Katoh, M. Functional and cancer genomics of ASXL family members. Br. J. Cancer 2013, 109, 299–306. [Google Scholar] [CrossRef] [Green Version]
  243. Mondet, J.; Presti, C.L.; Chevalier, S.; Bertrand, A.; Tondeur, S.; Blanchet, S.; McLeer, A.; Pernet-Gallay, K.; Mossuz, P. Mitochondria in human acute myeloid leukemia cell lines have ultrastructural alterations linked to deregulation of their respiratory profiles. Exp. Hematol. 2021, 98, 53–62.e3. [Google Scholar] [CrossRef] [PubMed]
  244. Carbuccia, N.; Murati, A.; Trouplin, V.; Brecqueville, M.; Adelaide, J.; Rey, J.M.G.; Vainchenker, W.; Bernard, O.A.; Chaffanet, M.; Vey, N.; et al. Mutations of ASXL1 gene in myeloproliferative neoplasms. Leukemia 2009, 23, 2183–2186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Gelsi-Boyer, V.; Trouplin, V.; Adélaïde, J.; Bonansea, J.; Cervera, N.; Carbuccia, N.; Lagarde, A.; Prebet, T.; Nezri, M.; Sainty, D.; et al. Mutations of polycomb-associated geneASXL1in myelodysplastic syndromes and chronic myelomonocytic leukaemia. Br. J. Haematol. 2009, 145, 788–800. [Google Scholar] [CrossRef]
  246. Quesada, V.; Conde, L.; Villamor, N.; Ordóñez, G.R.; Jares, P.; Bassaganyas, L.; Ramsay, A.J.; Beà, S.; Pinyol, M.; Martínez-Trillos, A.; et al. Exome sequencing identifies recurrent mutations of the splicing factor SF3B1 gene in chronic lymphocytic leukemia. Nat. Genet. 2011, 44, 47–52. [Google Scholar] [CrossRef]
  247. Gelsi-Boyer, V.; Trouplin, V.; Roquain, J.; Adélaïde, J.; Carbuccia, N.; Esterni, B.; Finetti, P.; Murati, A.; Arnoulet, C.; Zerazhi, H.; et al. ASXL1 mutation is associated with poor prognosis and acute transformation in chronic myelomonocytic leukaemia. Br. J. Haematol. 2010, 151, 365–375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Chen, T.-C.; Hou, H.-A.; Chou, W.-C.; Tang, J.-L.; Kuo, Y.-Y.; Chen, C.-Y.; Tseng, M.-H.; Huang, C.-F.; Lai, Y.-J.; Chiang, Y.-C.; et al. Dynamics of ASXL1 mutation and other associated genetic alterations during disease progression in patients with primary myelodysplastic syndrome. Blood Cancer J. 2014, 4, e177. [Google Scholar] [CrossRef] [PubMed]
  249. Thol, F.; Friesen, I.; Damm, F.; Yun, H.; Weissinger, E.M.; Krauter, J.; Wagner, K.; Chaturvedi, A.; Sharma, A.; Wichmann, M.; et al. Prognostic Significance of ASXL1 Mutations in Patients with Myelodysplastic Syndromes. J. Clin. Oncol. 2011, 29, 2499–2506. [Google Scholar] [CrossRef] [PubMed]
  250. Metzeler, K.; Becker, H.; Maharry, K.; Radmacher, M.D.; Kohlschmidt, J.; Mrózek, K.; Nicolet, D.; Whitman, S.P.; Wu, Y.-Z.; Schwind, S.; et al. ASXL1 mutations identify a high-risk subgroup of older patients with primary cytogenetically normal AML within the ELN Favorable genetic category. Blood 2011, 118, 6920–6929. [Google Scholar] [CrossRef]
  251. Shivarov, V.; Gueorguieva, R.; Ivanova, M.; Tiu, R.V. ASXL1 mutations define a subgroup of patients with acute myeloid leukemia with distinct gene expression profile and poor prognosis: A meta-analysis of 3311 adult patients with acute myeloid leukemia. Leuk. Lymphoma 2015, 56, 1–3. [Google Scholar] [CrossRef]
  252. Chou, W.-C.; Huang, H.-H.; Hou, H.-A.; Chen, C.-Y.; Tang, J.-L.; Yao, M.; Tsay, W.; Ko, B.-S.; Wu, S.-J.; Huang, S.-Y.; et al. Distinct clinical and biological features of de novo acute myeloid leukemia with additional sex comb-like 1 (ASXL1) mutations. Blood 2010, 116, 4086–4094. [Google Scholar] [CrossRef] [Green Version]
  253. El-Sharkawi, D.; Ali, A.; Evans, C.M.; Hills, R.K.; Burnett, A.K.; Linch, D.C.; Gale, R.E. ASXL1mutations are infrequent in young patients with primary acute myeloid leukemia and their detection has a limited role in therapeutic risk stratification. Leuk. Lymphoma 2014, 55, 1326–1331. [Google Scholar] [CrossRef]
  254. Schnittger, S.; Eder, C.; Jeromin, S.; Alpermann, T.; Fasan, A.; Grossmann, V.; Kohlmann, A.; Illig, T.; Klopp, N.; Wichmann, H.-E.; et al. ASXL1 exon 12 mutations are frequent in AML with intermediate risk karyotype and are independently associated with an adverse outcome. Leukemia 2013, 27, 82–91. [Google Scholar] [CrossRef]
  255. Pratcorona, M.; Abbas, S.; Sanders, M.A.; Koenders, J.E.; Kavelaars, F.G.; Erpelinck-Verschueren, C.A.J.; Zeilemakers, A.; Lowenberg, B.; Valk, P.J.M. Acquired mutations in ASXL1 in acute myeloid leukemia: Prevalence and prognostic value. Haematologica 2012, 97, 388–392. [Google Scholar] [CrossRef] [Green Version]
  256. Rocquain, J.; Carbuccia, N.; Trouplin, V.; Raynaud, S.; Murati, A.; Nezri, M.; Tadrist, Z.; Olschwang, S.; Vey, N.; Birnbaum, D.; et al. Combined mutations of ASXL1, CBL, FLT3, IDH1, IDH2, JAK2, KRAS, NPM1, NRAS, RUNX1, TET2 and WT1 genes in myelodysplastic syndromes and acute myeloid leukemias. BMC Cancer 2010, 10, 401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Lin, Y.; Zheng, Y.; Wang, Z.-C.; Wang, S.-Y. Prognostic significance of ASXL1 mutations in myelodysplastic syndromes and chronic myelomonocytic leukemia: A meta-analysis. Hematology 2016, 21, 454–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Rahmani, N.E.; Ramachandra, N.; Sahu, S.; Gitego, N.; Lopez, A.; Pradhan, K.; Bhagat, T.D.; Gordon-Mitchell, S.; Pena, B.R.; Kazemi, M.; et al. ASXL1 mutations are associated with distinct epigenomic alterations that lead to sensitivity to venetoclax and azacytidine. Blood Cancer J. 2021, 11, 157. [Google Scholar] [CrossRef]
  259. Guerra, V.A.; DiNardo, C.; Konopleva, M. Venetoclax-based therapies for acute myeloid leukemia. Best Pr. Res. Clin. Haematol. 2019, 32, 145–153. [Google Scholar] [CrossRef]
  260. Ntziachristos, P.; Tsirigos, A.; Van Vlierberghe, P.; Nedjic, J.; Trimarchi, T.; Flaherty, M.S.; Ferres-Marco, D.; Da Ros, V.G.; Tang, Z.; Siegle, J.; et al. Genetic inactivation of the polycomb repressive complex 2 in T cell acute lymphoblastic leukemia. Nat. Med. 2012, 18, 298–301. [Google Scholar] [CrossRef]
  261. Shimizu, T.; Kubovcakova, L.; Nienhold, R.; Zmajkovic, J.; Meyer, S.C.; Hao-Shen, H.; Geier, F.; Dirnhofer, S.; Guglielmelli, P.; Vannucchi, A.M.; et al. Loss of Ezh2 synergizes with JAK2-V617F in initiating myeloproliferative neoplasms and promoting myelofibrosis. J. Exp. Med. 2016, 213, 1479–1496. [Google Scholar] [CrossRef]
  262. Tan, J.-Z.; Yan, Y.; Wang, X.-X.; Jiang, Y.; Xu, H.E. EZH2: Biology, disease, and structure-based drug discovery. Acta Pharmacol. Sin. 2014, 35, 161–174. [Google Scholar] [CrossRef] [Green Version]
  263. Duan, R.; Du, W.; Guo, W. EZH2: A novel target for cancer treatment. J. Hematol. Oncol. 2020, 13, 104. [Google Scholar] [CrossRef] [PubMed]
  264. Rinke, J.; Chase, A.; Cross, N.C.P.; Hochhaus, A.; Ernst, T. EZH2 in Myeloid Malignancies. Cells 2020, 9, 1639. [Google Scholar] [CrossRef] [PubMed]
  265. Nishioka, C.; Ikezoe, T.; Yang, J.; Yokoyama, A. BCR/ABL increases EZH2 levels which regulates XIAP expression via miRNA-219 in chronic myeloid leukemia cells. Leuk. Res. 2016, 45, 24–32. [Google Scholar] [CrossRef] [PubMed]
  266. Huet, S.; Xerri, L.; Tesson, B.; Mareschal, S.; Taix, S.; Mescam-Mancini, L.; Sohier, E.; Carrère, C.; Lazarovici, J.; Casasnovas, R.-O.; et al. EZH2 alterations in follicular lymphoma: Biological and clinical correlations. Blood Cancer J. 2017, 7, e555. [Google Scholar] [CrossRef] [Green Version]
  267. Zhang, Q.; Han, Q.; Zi, J.; Ma, J.; Song, H.; Tian, Y.; McGrath, M.; Song, C.; Ge, Z. Mutations in EZH2 are associated with poor prognosis for patients with myeloid neoplasms. Genes Dis. 2019, 6, 276–281. [Google Scholar] [CrossRef]
  268. Stomper, J.; Meier, R.; Ma, T.; Pfeifer, D.; Ihorst, G.; Blagitko-Dorfs, N.; Greve, G.; Zimmer, D.; Platzbecker, U.; Hagemeijer, A.; et al. Integrative study of EZH2 mutational status, copy number, protein expression and H3K27 trimethylation in AML/MDS patients. Clin. Epigenetics 2021, 13, 1–14, Erratum in: Clin. Epigenetics 2021, 13, 94. [Google Scholar] [CrossRef]
  269. Wang, J.; He, N.; Wang, R.; Tian, T.; Han, F.; Zhong, C.; Zhang, C.; Hua, M.; Ji, C.; Ma, D. Analysis of TET2 and EZH2 gene functions in chromosome instability in acute myeloid leukemia. Sci. Rep. 2020, 10, 2706. [Google Scholar] [CrossRef] [Green Version]
  270. Zhang, R.; Wang, R.; Chang, H.; Wu, F.; Liu, C.; Deng, D.; Fan, W. Downregulation of Ezh2 expression by RNA interference induces cell cycle arrest in the G0/G1 phase and apoptosis in U87 human glioma cells. Oncol. Rep. 2012, 28, 2278–2284. [Google Scholar] [CrossRef] [Green Version]
  271. Zeng, D.; Liu, M.; Pan, J. Blocking EZH2 methylation transferase activity by GSK126 decreases stem cell-like myeloma cells. Oncotarget 2017, 8, 3396–3411. [Google Scholar] [CrossRef] [Green Version]
  272. Tiffen, J.C.; Gunatilake, D.; Gallagher, S.; Gowrishankar, K.; Heinemann, A.; Cullinane, C.; Dutton-Regester, K.; Pupo, G.M.; Strbenac, D.; Yang, J.; et al. Targeting activating mutations of EZH2 leads to potent cell growth inhibition in human melanoma by derepression of tumor suppressor genes. Oncotarget 2015, 6, 27023–27036. [Google Scholar] [CrossRef] [Green Version]
  273. Zhou, X.; Gao, W.; Hua, H.; Ji, Z. LncRNA-BLACAT1 Facilitates Proliferation, Migration and Aerobic Glycolysis of Pancreatic Cancer Cells by Repressing CDKN1C via EZH2-Induced H3K27me3. Front. Oncol. 2020, 10. [Google Scholar] [CrossRef]
  274. Tao, T.; Chen, M.; Jiang, R.; Guan, H.; Huang, Y.; Su, H.; Hu, Q.; Han, X.; Xiao, J. Involvement of EZH2 in aerobic glycolysis of prostate cancer through miR-181b/HK2 axis. Oncol. Rep. 2017, 37, 1430–1436. [Google Scholar] [CrossRef] [Green Version]
  275. Ahmad, F.; Patrick, S.; Sheikh, T.; Sharma, V.; Pathak, P.; Malgulwar, P.B.; Kumar, A.; Joshi, S.D.; Sarkar, C.; Sen, E. Telomerase reverse transcriptase (TERT)-enhancer of zeste homolog 2 (EZH2) network regulates lipid metabolism and DNA damage responses in glioblastoma. J. Neurochem. 2017, 143, 671–683. [Google Scholar] [CrossRef] [Green Version]
  276. Eich, M.-L.; Athar, M.; Ferguson, J.E.; Varambally, S. EZH2-Targeted Therapies in Cancer: Hype or a Reality. Cancer Res. 2020, 80, 5449–5458. [Google Scholar] [CrossRef] [PubMed]
  277. Wen, S.; Wang, J.; Liu, P.; Li, Y.; Lu, W.; Hu, Y.; Liu, J.; He, Z.; Huang, P. Novel combination of histone methylation modulators with therapeutic synergy against acute myeloid leukemia in vitro and in vivo. Cancer Lett. 2018, 413, 35–45. [Google Scholar] [CrossRef]
  278. Kozlovski, I.; Siegfried, Z.; Amar-Schwartz, A.; Karni, R. The role of RNA alternative splicing in regulating cancer metabolism. Hum. Genet. 2017, 136, 1113–1127. [Google Scholar] [CrossRef] [PubMed]
  279. Singh, B.; Eyras, E. The role of alternative splicing in cancer. Transcription 2017, 8, 91–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Qiu, J.; Zhou, B.; Thol, F.; Zhou, Y.; Chen, L.; Shao, C.; DeBoever, C.; Hou, J.; Li, H.; Chaturvedi, A.; et al. Distinct splicing signatures affect converged pathways in myelodysplastic syndrome patients carrying mutations in different splicing regulators. RNA 2016, 22, 1535–1549. [Google Scholar] [CrossRef] [Green Version]
  281. Lee, S.; Abdel-Wahab, O. Therapeutic targeting of splicing in cancer. Nat. Med. 2016, 22, 976–986. [Google Scholar] [CrossRef]
  282. Lee, S.C.-W.; Dvinge, H.; Kim, E.; Cho, H.; Micol, J.-B.; Chung, Y.R.; Durham, B.H.; Yoshimi, A.; Kim, Y.J.; Thomas, M.; et al. Modulation of splicing catalysis for therapeutic targeting of leukemia with mutations in genes encoding spliceosomal proteins. Nat. Med. 2016, 22, 672–678. [Google Scholar] [CrossRef] [Green Version]
  283. Zhang, L.; Padron, E.; Lancet, J. The molecular basis and clinical significance of genetic mutations identified in myelodysplastic syndromes. Leuk. Res. 2015, 39, 6–17. [Google Scholar] [CrossRef] [PubMed]
  284. Hnilicová, J.; Hozeifi, S.; Duskova, E.; Icha, J.; Tomankova, T.; Staněk, D. Histone Deacetylase Activity Modulates Alternative Splicing. PLoS ONE 2011, 6, e16727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Zhou, H.-L.; Luo, G.; Wise, J.A.; Lou, H. Regulation of alternative splicing by local histone modifications: Potential roles for RNA-guided mechanisms. Nucleic Acids Res. 2014, 42, 701–713. [Google Scholar] [CrossRef] [Green Version]
  286. Brocks, D.; Schmidt, C.R.; Daskalakis, M.; Jang, H.S.; Shah, N.M.; Li, D.; Li, J.; Zhang, B.; Hou, Y.; Laudato, S.; et al. DNMT and HDAC inhibitors induce cryptic transcription start sites encoded in long terminal repeats. Nat. Genet. 2017, 49, 1052–1060. [Google Scholar] [CrossRef]
  287. Demircioğlu, D.; Cukuroglu, E.; Kindermans, M.; Nandi, T.; Calabrese, C.; Fonseca, N.; Kahles, A.; Lehmann, K.-V.; Stegle, O.; Brazma, A.; et al. A Pan-cancer Transcriptome Analysis Reveals Pervasive Regulation through Alternative Promoters. Cell 2019, 178, 1465–1477.e17. [Google Scholar] [CrossRef]
  288. Papaemmanuil, E.; Cazzola, M.; Boultwood, J.; Malcovati, L.; Vyas, P.; Bowen, D.; Pellagatti, A.; Wainscoat, J.S.; Hellstrom-Lindberg, E.; Gambacorti-Passerini, C.; et al. Somatic SF3B1 Mutation in Myelodysplasia with Ring Sideroblasts. N. Engl. J. Med. 2011, 365, 1384–1395. [Google Scholar] [CrossRef] [Green Version]
  289. Patnaik, M.M.; Lasho, T.L.; Hodnefield, J.M.; Knudson, R.A.; Ketterling, R.P.; Garcia-Manero, G.; Steensma, D.P.; Pardanani, A.; Hanson, C.A.; Tefferi, A. SF3B1 mutations are prevalent in myelodysplastic syndromes with ring sideroblasts but do not hold independent prognostic value. Blood 2012, 119, 569–572. [Google Scholar] [CrossRef] [Green Version]
  290. Bravo, G.M.; Lee, E.; Merchan, B.; Kantarjian, H.M.; García-Manero, G. Integrating genetics and epigenetics in myelodysplastic syndromes: Advances in pathogenesis and disease evolution. Br. J. Haematol. 2014, 166, 646–659. [Google Scholar] [CrossRef] [PubMed]
  291. Dalton, W.B. The metabolic reprogramming and vulnerability of SF3B1 mutations. Mol. Cell. Oncol. 2020, 7, 1697619. [Google Scholar] [CrossRef]
  292. Wu, S.-J.; Kuo, Y.-Y.; Hou, H.-A.; Li, L.-Y.; Tseng, M.-H.; Huang, C.-F.; Lee, F.-Y.; Liu, M.-C.; Liu, C.-W.; Lin, C.-T.; et al. The clinical implication of SRSF2 mutation in patients with myelodysplastic syndrome and its stability during disease evolution. Blood 2012, 120, 3106–3111. [Google Scholar] [CrossRef]
  293. Thol, F.; Kade, S.; Schlarmann, C.; Löffeld, P.; Morgan, M.; Krauter, J.; Wlodarski, M.; Kölking, B.; Wichmann, M.; Görlich, K.; et al. Frequency and prognostic impact of mutations in SRSF2, U2AF1, and ZRSR2 in patients with myelodysplastic syndromes. Blood 2012, 119, 3578–3584. [Google Scholar] [CrossRef] [Green Version]
  294. Condon, K.J.; Orozco, J.M.; Adelmann, C.H.; Spinelli, J.B.; van der Helm, P.W.; Roberts, J.M.; Kunchok, T.; Sabatini, D.M. Genome-wide CRISPR screens reveal multitiered mechanisms through which mTORC1 senses mitochondrial dysfunction. Proc. Natl. Acad. Sci. USA 2021, 118, e2022120118. [Google Scholar] [CrossRef]
  295. Graubert, T.; Shen, D.; Ding, L.; Okeyo-Owuor, T.; Lunn, C.L.; Shao, J.; Krysiak, K.; Harris, C.C.; Koboldt, D.C.; Larson, D.; et al. Recurrent mutations in the U2AF1 splicing factor in myelodysplastic syndromes. Nat. Genet. 2011, 44, 53–57. [Google Scholar] [CrossRef] [Green Version]
  296. Zhu, Y.; Song, D.; Guo, J.; Jin, J.; Tao, Y.; Zhang, Z.; Xu, F.; He, Q.; Li, X.; Chang, C.; et al. U2AF1 mutation promotes tumorigenicity through facilitating autophagy flux mediated by FOXO3a activation in myelodysplastic syndromes. Cell Death Dis. 2021, 12, 655. [Google Scholar] [CrossRef]
  297. Liu, P.; Tarlé, S.A.; Hajra, A.; Claxton, D.F.; Marlton, P.; Freedman, M.; Siciliano, M.J.; Collins, F.S. Fusion Between Transcription Factor CBFβ/PEBP2β and a Myosin Heavy Chain in Acute Myeloid Leukemia. Science 1993, 261, 1041–1044. [Google Scholar] [CrossRef] [PubMed]
  298. Miyoshi, H.; Kozu, T.; Shimizu, K.; Enomoto, K.; Maseki, N.; Kaneko, Y.; Kamada, N.; Ohki, M. The t(8;21) translocation in acute my-eloid leukemia results in production of an AML1-MTG8 fusion transcript. EMBO J. 1993, 12, 2715–2721. [Google Scholar] [CrossRef] [PubMed]
  299. Gardini, A.; Cesaroni, M.; Luzi, L.; Okumura, A.J.; Biggs, J.R.; Minardi, S.P.; Venturini, E.; Zhang, D.-E.; Pelicci, P.G.; Alcalay, M. AML1/ETO Oncoprotein Is Directed to AML1 Binding Regions and Co-Localizes with AML1 and HEB on Its Targets. PLoS Genet. 2008, 4, e1000275. [Google Scholar] [CrossRef]
  300. Maiques-Diaz, A.; Chou, F.S.; Wunderlich, M.; López, G.G.; Jacinto, F.V.; Rodriguez-Perales, S.; Larrayoz, M.J.; Calasanz, M.J.; Mulloy, J.C.; Cigudosa, J.C.; et al. Chromatin modifications induced by the AML1-ETO fusion protein reversibly silence its genomic targets through AML1 and Sp1 binding motifs. Leukemia 2012, 26, 1329–1337. [Google Scholar] [CrossRef] [Green Version]
  301. Mandoli, A.; Singh, A.; Prange, K.; Tijchon, E.; Oerlemans, M.; Dirks, R.; Ter Huurne, M.; Wierenga, A.T.; Janssen-Megens, E.M.; Berentsen, K.; et al. The Hematopoietic Transcription Factors RUNX1 and ERG Prevent AML1-ETO Oncogene Overexpression and Onset of the Apoptosis Program in t(8;21) AMLs. Cell Rep. 2016, 17, 2087–2100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Trombly, D.J.; Whitfield, T.W.; Padmanabhan, S.; Gordon, J.A.R.; Lian, J.B.; Van Wijnen, A.J.; Zaidi, S.K.; Stein, J.L.; Stein, G.S. Genome-wide co-occupancy of AML1-ETO and N-CoR defines the t(8;21) AML signature in leukemic cells. BMC Genom. 2015, 16, 309. [Google Scholar] [CrossRef] [Green Version]
  303. Sun, X.-J.; Wang, Z.; Wang, L.; Jiang, Y.; Kost, N.; Soong, T.D.; Chen, W.-Y.; Tang, Z.; Nakadai, T.; Elemento, O.; et al. A stable transcription factor complex nucleated by oligomeric AML1–ETO controls leukaemogenesis. Nature 2013, 500, 93–97. [Google Scholar] [CrossRef] [Green Version]
  304. Li, Y.; Wang, H.; Wang, X.; Jin, W.; Tan, Y.; Fang, H.; Chen, S.; Chen, Z.; Wang, K. Genome-wide studies identify a novel interplay between AML1 and AML1/ETO in t(8;21) acute myeloid leukemia. Blood 2016, 127, 233–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Ptasinska, A.; Assi, S.A.; Mannari, D.; James, S.R.; Williamson, D.; Dunne, J.; Hoogenkamp, M.; Wu, M.; Care, M.; McNeill, H.; et al. Depletion of RUNX1/ETO in t(8;21) AML cells leads to genome-wide changes in chromatin structure and transcription factor binding. Leukemia 2012, 26, 1829–1841. [Google Scholar] [CrossRef]
  306. Wang, L.; Gural, A.; Sun, X.-J.; Zhao, X.; Perna, F.; Huang, G.; Hatlen, M.A.; Vu, L.; Liu, F.; Xu, H.; et al. The Leukemogenicity of AML1-ETO Is Dependent on Site-Specific Lysine Acetylation. Science 2011, 333, 765–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Spirin, P.V.; Lebedev, T.; Orlova, N.N.; Gornostaeva, A.S.; Prokofjeva, M.M.; Nikitenko, N.A.; Dmitriev, S.; Buzdin, A.A.; Borisov, N.M.; Aliper, A.M.; et al. Silencing AML1-ETO gene expression leads to simultaneous activation of both pro-apoptotic and proliferation signaling. Leukemia 2014, 28, 2222–2228. [Google Scholar] [CrossRef]
  308. Martinez-Soria, N.; McKenzie, L.; Draper, J.; Ptasinska, A.; Issa, H.; Potluri, S.; Blair, H.J.; Pickin, A.; Isa, A.; Chin, P.S.; et al. The Oncogenic Transcription Factor RUNX1/ETO Corrupts Cell Cycle Regulation to Drive Leukemic Transformation. Cancer Cell 2018, 34, 626–642.e8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  309. Grimwade, D.; Walker, H.; Harrison, G.; Oliver, F.; Chatters, S.; Harrison, C.; Wheatley, K.; Burnett, A.K.; Goldstone, A.H. The predictive value of hierarchical cytogenetic classification in older adults with acute myeloid leukemia (AML): Analysis of 1065 patients entered into the United Kingdom Medical Research Council AML11 trial. Blood 2001, 98, 1312–1320. [Google Scholar] [CrossRef] [Green Version]
  310. Skokowa, J.; Steinemann, D.; Katsman-Kuipers, J.E.; Zeidler, C.; Klimenkova, O.; Klimiankou, M.; Ünalan, M.; Kandabarau, S.; Makaryan, V.; Beekman, R.; et al. Cooperativity of RUNX1 and CSF3R mutations in severe congenital neutropenia: A unique pathway in myeloid leukemogenesis. Blood 2014, 123, 2229–2237. [Google Scholar] [CrossRef]
  311. Tang, J.-L.; Hou, H.-A.; Chen, C.-Y.; Liu, C.-Y.; Chou, W.-C.; Tseng, M.-H.; Huang, C.-F.; Lee, F.-Y.; Liu, M.-C.; Yao, M.; et al. AML1/RUNX1 mutations in 470 adult patients with de novo acute myeloid leukemia: Prognostic implication and interaction with other gene alterations. Blood 2009, 114, 5352–5361. [Google Scholar] [CrossRef]
  312. Greif, P.A.; Konstandin, N.P.; Metzeler, K.; Herold, T.; Pasalic, Z.; Ksienzyk, B.; Dufour, A.; Schneider, F.; Schneider, S.; Kakadia, P.M.; et al. RUNX1 mutations in cytogenetically normal acute myeloid leukemia are associated with a poor prognosis and up-regulation of lymphoid genes. Haematologica 2012, 97, 1909–1915. [Google Scholar] [CrossRef]
  313. Christiansen, D.H.; Andersen, M.K.; Pedersen-Bjergaard, J. Mutations of AML1 are common in therapy-related myelodysplasia following therapy with alkylating agents and are significantly associated with deletion or loss of chromosome arm 7q and with subsequent leukemic transformation. Blood 2004, 104, 1474–1481. [Google Scholar] [CrossRef] [PubMed]
  314. Harada, H.; Harada, Y.; Niimi, H.; Kyo, T.; Kimura, A.; Inaba, T. High incidence of somatic mutations in the AML1/RUNX1 gene in myelodysplastic syndrome and low blast percentage myeloid leukemia with myelodysplasia. Blood 2004, 103, 2316–2324. [Google Scholar] [CrossRef] [PubMed]
  315. Steensma, D.P.; Gibbons, R.; Mesa, R.A.; Tefferi, A.; Higgs, D.R. Somatic point mutations in RUNX1/CBFA2/AML1 are common in high-risk myelodysplastic syndrome, but not in myelofibrosis with myeloid metaplasia. Eur. J. Haematol. 2005, 74, 47–53. [Google Scholar] [CrossRef]
  316. Niimi, H.; Harada, H.; Harada, Y.; Ding, Y.; Imagawa, J.; Inaba, T.; Kyo, T.; Kimura, A. Hyperactivation of the RAS signaling pathway in myelodysplastic syndrome with AML1/RUNX1 point mutations. Leukemia 2006, 20, 635–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Chen, C.-Y.; Lin, L.-I.; Tang, J.-L.; Ko, B.-S.; Tsay, W.; Chou, W.-C.; Yao, M.; Wu, S.-J.; Tseng, M.-H.; Tien, H.-F. RUNX1 gene mutation in primary myelodysplastic syndrome-the mutation can be detected early at diagnosis or acquired during disease progression and is associated with poor outcome. Br. J. Haematol. 2007, 139, 405–414. [Google Scholar] [CrossRef]
  318. Migas, A.; Savva, N.; Mishkova, O.; Aleinikova, O.V. AML1/RUNX1 gene point mutations in childhood myeloid malignancies. Pediatr. Blood Cancer 2011, 57, 583–587. [Google Scholar] [CrossRef]
  319. Ismael, O.; Shimada, A.; Hama, A.; Elshazley, M.; Muramatsu, H.; Goto, A.; Sakaguchi, H.; Tanaka, M.; Takahashi, Y.; Yinyan, X.; et al. De novo childhood myelodysplastic/myeloproliferative disease with unique molecular characteristics. Br. J. Haematol. 2012, 158, 129–137. [Google Scholar] [CrossRef]
  320. Tsai, S.-C.; Shih, L.-Y.; Liang, S.-T.; Huang, Y.-J.; Kuo, M.-C.; Huang, C.-F.; Shih, Y.-S.; Lin, T.-H.; Chiu, M.-C.; Liang, D.-C. Biological Activities of RUNX1 Mutants Predict Secondary Acute Leukemia Transformation from Chronic Myelomonocytic Leukemia and Myelodysplastic Syndromes. Clin. Cancer Res. 2015, 21, 3541–3551. [Google Scholar] [CrossRef] [Green Version]
  321. Quentin, S.; Cuccuini, W.; Ceccaldi, R.; Nibourel, O.; Pondarre, C.; Pagès, M.-P.; Vasquez, N.; D’Enghien, C.D.; Larghero, J.; De Latour, R.P.; et al. Myelodysplasia and leukemia of Fanconi anemia are associated with a specific pattern of genomic abnormalities that includes cryptic RUNX1/AML1 lesions. Blood 2011, 117, e161–e170. [Google Scholar] [CrossRef] [Green Version]
  322. Haferlach, T.; Stengel, A.; Eckstein, S.; Perglerová, K.; Alpermann, T.; Kern, W.; Meggendorfer, M. The new provisional WHO entity ‘RUNX1 mutated AML’ shows specific genetics but no prognostic influence of dysplasia. Leukemia 2016, 30, 2109–2112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Sood, R.; Kamikubo, Y.; Liu, P. Role of RUNX1 in hematological malignancies. Blood 2017, 131, 373, Erratum in: Blood 2017, 129, 2070–2082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Yu, L.; Yu, T.-T.; Young, K.H. Cross-talk between Myc and p53 in B-cell lymphomas. Chronic Dis. Transl. Med. 2019, 5, 139–154. [Google Scholar] [CrossRef]
  325. Aubrey, B.; Kelly, G.L.; Janic, A.; Herold, M.; Strasser, A. How does p53 induce apoptosis and how does this relate to p53-mediated tumour suppression? Cell Death Differ. 2018, 25, 104–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Kotsantis, P.; Petermann, E.; Boulton, S.J. Mechanisms of Oncogene-Induced Replication Stress: Jigsaw Falling into Place. Cancer Discov. 2018, 8, 537–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Yeo, C.Q.X.; Alexander, I.; Lin, Z.; Lim, S.; Aning, O.A.; Kumar, R.; Sangthongpitag, K.; Pendharkar, V.; Ho, V.H.; Cheok, C.F. p53 Maintains Genomic Stability by Preventing Interference between Transcription and Replication. Cell Rep. 2016, 15, 132–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  328. Valente, L.J.; Gray, D.H.; Michalak, E.M.; Pinon-Hofbauer, J.; Egle, A.; Scott, C.L.; Janic, A.; Strasser, A. p53 Efficiently Suppresses Tumor Development in the Complete Absence of Its Cell-Cycle Inhibitory and Proapoptotic Effectors p21, Puma, and Noxa. Cell Rep. 2013, 3, 1339–1345. [Google Scholar] [CrossRef] [Green Version]
  329. Bloedjes, T.; de Wilde, G.; Guikema, J. Metabolic Effects of Recurrent Genetic Aberrations in Multiple Myeloma. Cancers 2021, 13, 396. [Google Scholar] [CrossRef]
  330. Liu, J.; Zhang, C.; Hu, W.; Feng, Z. Tumor suppressor p53 and metabolism. J. Mol. Cell Biol. 2019, 11, 284–292. [Google Scholar] [CrossRef] [Green Version]
  331. Gomes, A.S.; Ramos, H.; Soares, J.; Saraiva, L. p53 and glucose metabolism: An orchestra to be directed in cancer therapy. Pharmacol. Res. 2018, 131, 75–86. [Google Scholar] [CrossRef]
  332. Kim, Y.Y.; Um, J.; Yoon, J.; Lee, D.; Lee, Y.J.; Kim, D.H.; Park, J.; Yun, J. p53 regulates mitochondrial dynamics by inhibiting Drp1 translocation into mitochondria during cellular senescence. FASEB J. 2020, 34, 2451–2464. [Google Scholar] [CrossRef] [Green Version]
  333. Kim, Y.Y.; Um, J.; Shin, D.J.; Jeong, D.J.; Bin Hong, Y.; Yun, J. p53-mediated regulation of mitochondrial dynamics plays a pivotal role in the senescence of various normal cells as well as cancer cells. FASEB J. 2021, 35, e21319. [Google Scholar] [CrossRef]
  334. Alavi, M.V. Targeted OMA1 therapies for cancer. Int. J. Cancer 2019, 145, 2330–2341. [Google Scholar] [CrossRef]
  335. Silic-Benussi, M.; Scattolin, G.; Cavallari, I.; Minuzzo, S.A.; Del Bianco, P.; Francescato, S.; Basso, G.; Indraccolo, S.; D’Agostino, D.M.; Ciminale, V. Selective killing of human T-ALL cells: An integrated approach targeting redox homeostasis and the OMA1/OPA1 axis. Cell Death Dis. 2018, 9, 822. [Google Scholar] [CrossRef]
  336. Srivastava, S.; Zou, Z.; Pirollo, K.F.; Blattner, W.A.; Chang, E.H. Germ-line transmission of a mutated p53 gene in a cancer-prone family with Li–Fraumeni syndrome. Nature 1990, 348, 747–749. [Google Scholar] [CrossRef]
  337. Swaminathan, M.; Bannon, S.A.; Routbort, M.; Naqvi, K.; Kadia, T.M.; Takahashi, K.; Alvarado, Y.; Ravandi-Kashani, F.; Patel, K.P.; Champlin, R.; et al. Hematologic malignancies and Li–Fraumeni syndrome. Cold Spring Harb. Mol. Case Stud. 2019, 5, a003210. [Google Scholar] [CrossRef] [Green Version]
  338. Bernard, E.; Nannya, Y.; Hasserjian, R.P.; Devlin, S.M.; Tuechler, H.; Medina-Martinez, J.S.; Yoshizato, T.; Shiozawa, Y.; Saiki, R.; Malcovati, L.; et al. Implications of TP53 allelic state for genome stability, clinical presentation and outcomes in myelodysplastic syndromes. Nat. Med. 2020, 26, 1549–1556. [Google Scholar] [CrossRef] [PubMed]
  339. Hunter, A.M.; Sallman, D.A. Current status and new treatment approaches in TP53 mutated AML. Best Pr. Res. Clin. Haematol. 2019, 32, 134–144. [Google Scholar] [CrossRef] [PubMed]
  340. Welch, J.S. Patterns of mutations in TP53 mutated AML. Best Pr. Res. Clin. Haematol. 2018, 31, 379–383. [Google Scholar] [CrossRef]
  341. Papaemmanuil, E.; Gerstung, M.; Bullinger, L.; Gaidzik, V.I.; Paschka, P.; Roberts, N.D.; Potter, N.E.; Heuser, M.; Thol, F.; Bolli, N.; et al. Genomic Classification and Prognosis in Acute Myeloid Leukemia. N. Engl. J. Med. 2016, 374, 2209–2221. [Google Scholar] [CrossRef]
  342. Zlamalikova, L.; Moulis, M.; Ravcukova, B.; Liskova, K.; Malcikova, J.; Salek, D.; Jarkovsky, J.; Svitakova, M.; Hrabalkova, R.; Smarda, J.; et al. Complex analysis of the TP53 tumor suppressor in mantle cell and diffuse large B-cell lymphomas. Oncol. Rep. 2017, 38, 2535–2542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Voropaeva, E.N.; Pospelova, T.I.; Voevoda, M.I.; Maksimov, V.N. Frequency, spectrum, and functional significance of TP53 mutations in patients with diffuse large B-cell lymphoma. Mol. Biol. 2017, 51, 64–72. (In Russian) [Google Scholar] [CrossRef]
  344. Flynt, E.; Bisht, K.; Sridharan, V.; Ortiz, M.; Towfic, F.; Thakurta, A. Prognosis, Biology, and Targeting of TP53 Dysregulation in Multiple Myeloma. Cells 2020, 9, 287. [Google Scholar] [CrossRef] [Green Version]
  345. Maya-Mendoza, A.; Ostrakova, J.; Kosar, M.; Hall, A.; Duskova, P.; Mistrik, M.; Merchut-Maya, J.M.; Hodny, Z.; Bartkova, J.; Christensen, C.; et al. Myc and Ras oncogenes engage different energy metabolism programs and evoke distinct patterns of oxidative and DNA replication stress. Mol. Oncol. 2015, 9, 601–616. [Google Scholar] [CrossRef] [Green Version]
  346. Pomeroy, E.J.; Eckfeldt, C.E. Targeting Ras signaling in AML: RALB is a small GTPase with big potential. Small GTPases 2020, 11, 39–44. [Google Scholar] [CrossRef]
  347. Gough, D.J.; Corlett, A.; Schlessinger, K.; Wegrzyn, J.; Larner, A.C.; Levy, D.E. Mitochondrial STAT3 supports Ras-dependent oncogenic transformation. Science 2009, 324, 1713–1716. [Google Scholar] [CrossRef] [Green Version]
  348. Cai, J.; Wang, J.; Huang, Y.; Wu, H.-X.; Xia, T.; Xiao, J.; Chen, X.; Li, H.; Qiu, Y.; Wang, Y.; et al. ERK/Drp1-dependent mitochondrial fission is involved in the MSC-induced drug resistance of T-cell acute lymphoblastic leukemia cells. Cell Death Dis. 2016, 7, e2459. [Google Scholar] [CrossRef] [PubMed]
  349. Ganguly, B.B.; Kadam, N. Mutations of myelodysplastic syndromes (MDS): An update. Mutat. Res. Mutat. Res. 2016, 769, 47–62. [Google Scholar] [CrossRef]
  350. Shih, L.-Y.; Huang, C.F.; Wang, P.N.; Wu, J.H.; Lin, T.L.; Dunn, P.M.; Kuo, M.C. Acquisition of FLT3 or N-ras mutations is frequently associated with progression of myelodysplastic syndrome to acute myeloid leukemia. Leukemia 2004, 18, 466–475. [Google Scholar] [CrossRef] [PubMed]
  351. Wang, S.; Wu, Z.; Li, T.; Li, Y.; Wang, W.; Hao, Q.; Xie, X.; Wan, D.; Jiang, Z.; Wang, C.; et al. Mutational spectrum and prognosis in NRAS-mutated acute myeloid leukemia. Sci. Rep. 2020, 10, 12152. [Google Scholar] [CrossRef] [PubMed]
  352. Soverini, S.; Mancini, M.; Bavaro, L.; Cavo, M.; Martinelli, G. Chronic myeloid leukemia: The paradigm of targeting oncogenic tyrosine kinase signaling and counteracting resistance for successful cancer therapy. Mol. Cancer 2018, 17, 49. [Google Scholar] [CrossRef] [PubMed]
  353. Kleppe, M.; Kwak, M.; Koppikar, P.; Riester, M.; Keller, M.; Bastian, L.; Hricik, T.; Bhagwat, N.; McKenney, A.S.; Papalexi, E.; et al. JAK–STAT Pathway Activation in Malignant and Nonmalignant Cells Contributes to MPN Pathogenesis and Therapeutic Response. Cancer Discov. 2015, 5, 316–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Xia, D.; Hasserjian, R.P. Molecular testing forJAK2,MPL, andCALRin myeloproliferative neoplasms. Am. J. Hematol. 2016, 91, 1277–1280. [Google Scholar] [CrossRef]
  355. How, J.; Hobbs, G.S.; Mullally, A. Mutant calreticulin in myeloproliferative neoplasms. Blood 2019, 134, 2242–2248. [Google Scholar] [CrossRef] [Green Version]
  356. Rao, T.N.; Hansen, N.; Hilfiker, J.; Rai, S.; Majewska, J.-M.; Leković, D.; Gezer, D.; Andina, N.; Galli, S.; Cassel, T.; et al. JAK2-mutant hematopoietic cells display metabolic alterations that can be targeted to treat myeloproliferative neoplasms. Blood 2019, 134, 1832–1846. [Google Scholar] [CrossRef]
  357. Glowacki, S.; Synowiec, E.; Blasiak, J. The Role of Mitochondrial DNA Damage and Repair in the Resistance of BCR/ABL-Expressing Cells to Tyrosine Kinase Inhibitors. Int. J. Mol. Sci. 2013, 14, 16348–16364. [Google Scholar] [CrossRef] [Green Version]
  358. Sharma, V.; Wright, K.L.; Epling-Burnette, P.K.; Reuther, G.W. Metabolic Vulnerabilities and Epigenetic Dysregulation in Myeloproliferative Neoplasms. Front. Immunol. 2020, 11, 604142. [Google Scholar] [CrossRef] [PubMed]
  359. Akiyama, H.; Umezawa, Y.; Watanabe, D.; Okada, K.; Ishida, S.; Nogami, A.; Miura, O. Inhibition of USP9X Downregulates JAK2-V617F and Induces Apoptosis Synergistically with BH3 Mimetics Preferentially in Ruxolitinib-Persistent JAK2-V617F-Positive Leukemic Cells. Cancers 2020, 12, 406. [Google Scholar] [CrossRef] [PubMed]
  360. Danisz, K.; Blasiak, J. Role of anti-apoptotic pathways activated by BCR/ABL in the resistance of chronic myeloid leukemia cells to tyrosine kinase inhibitors. Acta Biochim. Pol. 2013, 60, 503–514. [Google Scholar] [CrossRef]
  361. Sontakke, P.; Koczula, K.M.; Jaques, J.; Wierenga, A.T.J.; Brouwers-Vos, A.Z.; Pruis, M.; Günther, U.L.; Vellenga, E.; Schuringa, J.J. Hypoxia-Like Signatures Induced by BCR-ABL Potentially Alter the Glutamine Uptake for Maintaining Oxidative Phosphorylation. PLoS ONE 2016, 11, e0153226. [Google Scholar] [CrossRef] [Green Version]
  362. Mugnaini, E.N.; Ghosh, N. Lymphoma. Prim. Care: Clin. Off. Pr. 2016, 43, 661–675. [Google Scholar] [CrossRef]
  363. Kaseb, H.; Babiker, H.M. Hodgkin Lymphoma. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2021. [Google Scholar]
  364. Kaya, E.; Keskin, L.; Aydogdu, I.; Kuku, I.; Bayraktar, N.; Erkut, M.A. Oxidant/antioxidant Parameters and their Relationship with Chemotherapy in Hodgkin’s Lymphoma. J. Int. Med Res. 2005, 33, 687–692. [Google Scholar] [CrossRef] [Green Version]
  365. Othman, T.; Herrera, A.; Mei, M. Emerging Therapies in Relapsed and Refractory Hodgkin Lymphoma: What Comes Next After Brentuximab Vedotin and PD-1 Inhibition? Curr. Hematol. Malign- Rep. 2021, 16, 1–7. [Google Scholar] [CrossRef]
  366. Birkenmeier, K.; Dröse, S.; Wittig, I.; Winkelmann, R.; Käfer, V.; Döring, C.; Hartmann, S.; Wenz, T.; Reichert, A.S.; Brandt, U.; et al. Hodgkin and Reed-Sternberg cells of classical Hodgkin lymphoma are highly dependent on oxidative phosphorylation. Int. J. Cancer 2016, 138, 2231–2246. [Google Scholar] [CrossRef] [Green Version]
  367. Tiacci, E.; Ladewig, E.; Schiavoni, G.; Penson, A.; Fortini, E.; Pettirossi, V.; Wang, Y.; Rosseto, A.; Venanzi, A.; Vlasevska, S.; et al. Pervasive mutations of JAK-STAT pathway genes in classical Hodgkin lymphoma. Blood 2018, 131, 2454–2465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  368. Küppers, R. The biology of Hodgkin’s lymphoma. Nat. Rev. Cancer 2009, 9, 15–27. [Google Scholar] [CrossRef] [PubMed]
  369. Wlodarska, I.; Nooyen, P.; Maes, B.; Martı́n-Subero, J.I.; Siebert, R.; Pauwels, P.; De Wolf-Peeters, C.; Hagemeijer, A. Frequent occurrence of BCL6 rearrangements in nodular lymphocyte predominance Hodgkin lymphoma but not in classical Hodgkin lymphoma. Blood 2003, 101, 706–710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  370. Esmeray, E.; Küçük, C. Genetic alterations in B cell lymphoma subtypes as potential biomarkers for non-invasive diagnosis, prognosis, therapy, and disease monitoring. Turk. J. Boil. 2020, 44, 1–14. [Google Scholar] [CrossRef]
  371. Armitage, J.O.; Gascoyne, R.D.; Lunning, M.A.; Cavalli, F. Non-Hodgkin lymphoma. Lancet 2017, 390, 298–310. [Google Scholar] [CrossRef]
  372. Armitage, J.O. The aggressive peripheral T-cell lymphomas: 2017. Am. J. Hematol. 2017, 92, 706–715. [Google Scholar] [CrossRef] [Green Version]
  373. Iqbal, J.; Wright, G.; Wang, C.; Rosenwald, A.; Gascoyne, R.D.; Weisenburger, D.D.; Greiner, T.C.; Smith, L.; Guo, S.; Wilcox, R.A.; et al. Gene expression signatures delineate biological and prognostic subgroups in peripheral T-cell lymphoma. Blood 2014, 123, 2915–2923. [Google Scholar] [CrossRef]
  374. Berendsen, M.R.; Stevens, W.B.C.; Brand, M.V.D.; Van Krieken, J.H.; Scheijen, B. Molecular Genetics of Relapsed Diffuse Large B-Cell Lymphoma: Insight into Mechanisms of Therapy Resistance. Cancers 2020, 12, 3553. [Google Scholar] [CrossRef] [PubMed]
  375. Lacy, S.E.; Barrans, S.L.; Beer, P.A.; Painter, D.; Smith, A.G.; Roman, E.; Cooke, S.L.; Ruiz, C.; Glover, P.; Van Hoppe, S.J.L.; et al. Targeted sequencing in DLBCL, molecular subtypes, and outcomes: A Haematological Malignancy Research Network report. Blood 2020, 135, 1759–1771. [Google Scholar] [CrossRef]
  376. Matolcsy, A.; Casali, P.; Warnke, R.A.; Knowles, D.M. Morphologic transformation of follicular lymphoma is associated with somatic mutation of the translocated Bcl-2 gene. Blood 1996, 88, 3937–3944. [Google Scholar] [CrossRef] [Green Version]
  377. Lo Coco, F.; Gaidano, G.; Louie, D.C.; Offit, K.; Chaganti, R.S.; Dalla-Favera, R. p53 mutations are associated with histologic trans-formation of follicular lymphoma. Blood 1993, 82, 2289–2295. [Google Scholar] [CrossRef] [Green Version]
  378. Correia, C.; Schneider, P.A.; Dai, H.; Dogan, A.; Maurer, M.J.; Church, A.K.; Novak, A.J.; Feldman, A.L.; Wu, X.; Ding, H.; et al. BCL2 mutations are associated with increased risk of transformation and shortened survival in follicular lymphoma. Blood 2015, 125, 658–667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  379. Bouska, A.; Zhang, W.; Gong, Q.; Iqbal, J.; Scuto, A.; Vose, J.; Ludvigsen, M.; Fu, K.; Weisenburger, D.D.; Greiner, T.C.; et al. Combined copy number and mutation analysis identifies oncogenic pathways associated with transformation of follicular lymphoma. Leukemia 2017, 31, 83–91. [Google Scholar] [CrossRef] [Green Version]
  380. Ricci, J.E.; Chiche, J. Metabolic Reprogramming of Non-Hodgkin’s B-Cell Lymphomas and Potential Therapeutic Strategies. Front. Oncol. 2018, 8, 556. [Google Scholar] [CrossRef]
  381. Nabrinsky, E.; Danilov, A.V.; Koller, P.B. High-Risk Mantle Cell Lymphoma in the Era of Novel Agents. Curr. Hematol. Malign- Rep. 2021, 16, 8–18. [Google Scholar] [CrossRef] [PubMed]
  382. Rosenthal, A.; Younes, A. High grade B-cell lymphoma with rearrangements of MYC and BCL2 and/or BCL6: Double hit and triple hit lymphomas and double expressing lymphoma. Blood Rev. 2017, 31, 37–42. [Google Scholar] [CrossRef]
  383. Miyaoka, M.; Kikuti, Y.Y.; Carreras, J.; Ikoma, H.; Hiraiwa, S.; Ichiki, A.; Kojima, M.; Ando, K.; Yokose, T.; Sakai, R.; et al. Clinicopathological and genomic analysis of double-hit follicular lymphoma: Comparison with high-grade B-cell lymphoma with MYC and BCL2 and/or BCL6 rearrangements. Mod. Pathol. 2018, 31, 313–326. [Google Scholar] [CrossRef]
  384. Chung, C. Current targeted therapies in lymphomas. Am. J. Health Syst. Pharm. 2019, 76, 1825–1834. [Google Scholar] [CrossRef]
  385. Klener, P.; Klanova, M. Drug Resistance in Non-Hodgkin Lymphomas. Int. J. Mol. Sci. 2020, 21, 2081. [Google Scholar] [CrossRef] [Green Version]
  386. Xue, E.A.W.; Zhang, M. Updating targets for natural killer/T-cell lymphoma immunotherapy. Cancer Biol. Med. 2021, 18, 52–62. [Google Scholar] [CrossRef] [PubMed]
  387. Wang, H.; Fu, B.-B.; Gale, R.P.; Liang, Y. NK-/T-cell lymphomas. Leukemia 2021, 35, 2460–2468. [Google Scholar] [CrossRef]
  388. Xu, R.-H.; Pelicano, H.; Zhou, Y.; Carew, J.S.; Feng, L.; Bhalla, K.N.; Keating, M.J.; Huang, P. Inhibition of glycolysis in cancer cells: A novel strategy to overcome drug resistance associated with mitochondrial respiratory defect and hypoxia. Cancer Res. 2005, 65, 613–621. [Google Scholar] [PubMed]
  389. Caro, P.; Kishan, A.U.; Norberg, E.; Stanley, I.A.; Chapuy, B.; Ficarro, S.B.; Polak, K.; Tondera, D.; Gounarides, J.; Yin, H.; et al. Metabolic Signatures Uncover Distinct Targets in Molecular Subsets of Diffuse Large B Cell Lymphoma. Cancer Cell 2012, 22, 547–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  390. Kusao, I.; Troelstrup, D.; Shiramizu, B. Possible Mitochondria-Associated Enzymatic Role in Non-Hodgkin Lymphoma Residual Disease. Cancer Growth Metastasis 2008, 1, 3–8. [Google Scholar] [CrossRef] [Green Version]
  391. Zeng, A.G.X.; Leung, A.C.Y.; Brooks-Wilson, A.R. Somatic Mitochondrial DNA Mutations in Diffuse Large B-Cell Lymphoma. Sci. Rep. 2018, 8, 3623. [Google Scholar] [CrossRef] [Green Version]
  392. Kikushige, Y. Pathogenesis of chronic lymphocytic leukemia and the development of novel therapeutic strategies. J. Clin. Exp. Hematop. 2020, 60, 146–158. [Google Scholar] [CrossRef]
  393. Strati, P.; Abruzzo, L.V.; Wierda, W.G.; O’Brien, S.; Ferrajoli, A.; Keating, M.J. Second Cancers and Richter Transformation Are the Leading Causes of Death in Patients with Trisomy 12 Chronic Lymphocytic Leukemia. Clin. Lymphoma Myeloma Leuk. 2015, 15, 420–427. [Google Scholar] [CrossRef] [Green Version]
  394. Hallek, M. Chronic lymphocytic leukemia: 2020 update on diagnosis, risk stratification and treatment. Am. J. Hematol. 2019, 94, 1266–1287. [Google Scholar] [CrossRef] [Green Version]
  395. Nadeu, F.; Delgado, J.; Royo, C.; Baumann, T.; Stankovic, T.; Pinyol, M.; Jares, P.; Navarro, A.; Garcia, D.M.; Beà, S.; et al. Clinical impact of clonal and subclonal TP53, SF3B1, BIRC3, NOTCH1, and ATM mutations in chronic lymphocytic leukemia. Blood 2016, 127, 2122–2130. [Google Scholar] [CrossRef]
  396. Jitschin, R.; Hofmann, A.D.; Bruns, H.; Gießl, A.; Bricks, J.; Berger, J.; Saul, D.; Eckart, M.J.; Mackensen, A.; Mougiakakos, D. Mitochondrial metabolism contributes to oxidative stress and reveals therapeutic targets in chronic lymphocytic leukemia. Blood 2014, 123, 2663–2672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  397. Vazquez, G.G.; Aloyz, R. Metabolic rewiring beyond Warburg in chronic lymphocytic leukemia: How much do we actually know? Crit. Rev. Oncol. 2019, 134, 65–70. [Google Scholar] [CrossRef]
  398. Sahin, E.; Colla, S.; Liesa, M.; Moslehi, J.; Muller, F.; Guo, M.; Cooper, M.; Kotton, D.; Fabian, A.J.; Walkey, C.; et al. Telomere dysfunction induces metabolic and mitochondrial compromise. Nature 2011, 470, 359–365. [Google Scholar] [CrossRef] [Green Version]
  399. Hosnijeh, F.S.; Lan, Q.; Rothman, N.; Liu, C.S.; Cheng, W.-L.; Nieters, A.; Guldberg, P.; Tjonneland, A.; Campa, D.; Martino, A.; et al. Mitochondrial DNA copy number and future risk of B-cell lymphoma in a nested case-control study in the prospective EPIC cohort. Blood 2014, 124, 530–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  400. Zhang, X.; Rastogi, P.; Shah, B.; Zhang, L. B lymphoblastic leukemia/lymphoma: New insights into genetics, molecular aberrations, subclassification and targeted therapy. Oncotarget 2017, 8, 66728–66741. [Google Scholar] [CrossRef] [Green Version]
  401. Mullighan, C.G. The molecular genetic makeup of acute lymphoblastic leukemia. Hematol. Am. Soc. Hematol. Educ. Program. 2012, 2012, 389–396. [Google Scholar] [CrossRef] [Green Version]
  402. Woo, J.S.; Alberti, M.O.; Tirado, C.A. Childhood B-acute lymphoblastic leukemia: A genetic update. Exp. Hematol. Oncol. 2014, 3, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  403. Inaba, H.; Mullighan, C.G. Pediatric acute lymphoblastic leukemia. Haematologica 2020, 105, 2524–2539. [Google Scholar] [CrossRef]
  404. Rytelewski, M.; Harutyunyan, K.; Baran, N.; Mallampati, S.; Zal, M.A.; Cavazos, A.; Butler, J.M.; Konoplev, S.; El Khatib, M.; Plunkett, S.; et al. Inhibition of Oxidative Phosphorylation Reverses Bone Marrow Hypoxia Visualized in Imageable Syngeneic B-ALL Mouse Model. Front. Oncol. 2020, 10, 991. [Google Scholar] [CrossRef] [PubMed]
  405. Madden, S.K.; de Araujo, A.D.; Gerhardt, M.; Fairlie, D.P.; Mason, J.M. Taking the Myc out of cancer: Toward therapeutic strategies to directly inhibit c-Myc. Mol. Cancer 2021, 20, 3. [Google Scholar] [CrossRef]
  406. Wade, M.; Wahl, G.M. c-Myc, Genome Instability, and Tumorigenesis: The Devil Is in the Details. Curr. Top. Microbiol. Immunol. 2006, 302, 169–203. [Google Scholar] [CrossRef]
  407. 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]
  408. Vafa, O.; Wade, M.; Kern, S.; Beeche, M.; Pandita, T.K.; Hampton, G.M.; Wahl, G.M. c-Myc Can Induce DNA Damage, Increase Reactive Oxygen Species, and Mitigate p53 Function: A Mechanism for Oncogene-Induced Genetic Instability. Mol. Cell 2002, 9, 1031–1044. [Google Scholar] [CrossRef]
  409. Eertink, J.J.; Arens, A.I.J.; Huijbregts, J.E.; Celik, F.; de Keizer, B.; Stroobants, S.; de Jong, D.; Wiegers, S.E.; Zwezerijnen, G.J.C.; Burggraaff, C.N.; et al. Aberrant patterns of PET response during treatment for DLBCL patients with MYC gene rearrangements. Eur. J. Nucl. Med. Mol. Imaging 2021. Epub ahead of print. [Google Scholar] [CrossRef] [PubMed]
  410. Vaupel, P.; Multhoff, G. The Warburg Effect: Historical Dogma Versus Current Rationale. Adv. Exp. Med. Biol. 2021, 1269, 169–177. [Google Scholar] [CrossRef]
  411. Osthus, R.C.; Shim, H.; Kim, S.; Li, Q.; Reddy, R.; Mukherjee, M.; Xu, Y.; Wonsey, D.; Lee, L.A.; Dang, C.V. Deregulation of Glucose Transporter 1 and Glycolytic Gene Expression by c-Myc. J. Biol. Chem. 2000, 275, 21797–21800. [Google Scholar] [CrossRef] [Green Version]
  412. Goetzman, E.S.; Prochownik, E.V. The Role for Myc in Coordinating Glycolysis, Oxidative Phosphorylation, Glutaminolysis, and Fatty Acid Metabolism in Normal and Neoplastic Tissues. Front. Endocrinol. 2018, 9, 129. [Google Scholar] [CrossRef]
  413. Casciano, J.C.; Perry, C.; Cohen-Nowak, A.J.; Miller, K.D.; Voorde, J.V.; Zhang, Q.; Chalmers, S.; Sandison, M.E.; Liu, Q.; Hedley, A.; et al. MYC regulates fatty acid metabolism through a multigenic program in claudin-low triple negative breast cancer. Br. J. Cancer 2020, 122, 868–884. [Google Scholar] [CrossRef]
  414. Kim, J.; Lee, J.-H.; Iyer, V.R. Global Identification of Myc Target Genes Reveals Its Direct Role in Mitochondrial Biogenesis and Its E-Box Usage In Vivo. PLoS ONE 2008, 3, e1798. [Google Scholar] [CrossRef] [Green Version]
  415. Rosenwald, A.; Bens, S.; Advani, R.; Barrans, S.; Copie-Bergman, C.; Elsensohn, M.H.; Natkunam, Y.; Calaminici, M.; Sander, B.; Baia, M.; et al. Prognostic Significance of MYC Rearrangement and Translocation Partner in Diffuse Large B-Cell Lymphoma: A Study by the Lunenburg Lymphoma Biomarker Consortium. J. Clin. Oncol. 2019, 37, 3359–3368. [Google Scholar]
  416. Xia, Y.; Zhang, X. The Spectrum of MYC Alterations in Diffuse Large B-Cell Lymphoma. Acta Haematol. 2020, 143, 520–528. [Google Scholar] [CrossRef]
  417. Xu-Monette, Z.Y.; Deng, Q.; Manyam, G.C.; Tzankov, A.; Li, L.; Xia, Y.; Wang, X.-X.; Zou, D.; Visco, C.; Dybkær, K.; et al. MYC mutation profiling and prognostic significance in de novo diffuse large B-cell lymphoma. Clin. Cancer Res. 2016, 22, 3593–3605. [Google Scholar] [CrossRef] [Green Version]
  418. Ozsvari, B.; Sotgia, F.; Lisanti, M.P. A new mutation-independent approach to cancer therapy: Inhibiting oncogenic RAS and MYC, by targeting mitochondrial biogenesis. Aging 2017, 9, 2098–2116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  419. D’Aquila, P.; Ronchetti, D.; Cantafio, M.G.; Todoerti, K.; Taiana, E.; Fabiani, F.; Montesanto, A.; Neri, A.; Passarino, G.; Viglietto, G.; et al. Epigenetic Regulation of Mitochondrial Quality Control Genes in Multiple Myeloma: A Sequenom MassARRAY Pilot Investigation on HMCLs. J. Clin. Med. 2021, 10, 1295. [Google Scholar] [CrossRef] [PubMed]
  420. Gebhard, A.W.; Jain, P.; Nair, R.R.; Emmons, M.F.; Argilagos, R.F.; Koomen, J.M.; McLaughlin, M.L.; Hazlehurst, L.A. MTI-101 (cyclized HYD1) binds a CD44 containing complex and induces necrotic cell death in multiple myeloma. Mol. Cancer Ther. 2013, 12, 2446–2458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  421. Vandenabeele, P.; Galluzzi, L.; Berghe, T.V.; Kroemer, G. Molecular mechanisms of necroptosis: An ordered cellular explosion. Nat. Rev. Mol. Cell Biol. 2010, 11, 700–714. [Google Scholar] [CrossRef]
  422. Zhan, X.; Yu, W.; Franqui-Machin, R.; Bates, M.; Nadiminti, K.; Cao, H.; Amendt, B.A.; Jethava, Y.; Frech, I.; Zhan, F.; et al. Alteration of mitochondrial biogenesis promotes disease progression in multiple myeloma. Oncotarget 2017, 8, 111213–111224. [Google Scholar] [CrossRef]
  423. Marlein, C.R.; Piddock, R.E.; Mistry, J.J.; Zaitseva, L.; Hellmich, C.; Horton, R.H.; Zhou, Z.; Auger, M.J.; Bowles, K.M.; Rushworth, S.A. CD38-Driven Mitochondrial Trafficking Promotes Bioenergetic Plasticity in Multiple Myeloma. Cancer Res. 2019, 79, 2285–2297. [Google Scholar] [CrossRef] [Green Version]
  424. Katsuoka, F.; Yamamoto, M. Small Maf proteins (MafF, MafG, MafK): History, structure and function. Gene 2016, 586, 197–205. [Google Scholar] [CrossRef] [Green Version]
  425. Takahashi, S. Functional analysis of large MAF transcription factors and elucidation of their relationships with human diseases. Exp. Anim. 2021, 70, 264–271. [Google Scholar] [CrossRef]
  426. Perumal, D.; Kuo, P.-Y.; Leshchenko, V.V.; Jiang, Z.; Divakar, S.K.A.; Cho, H.J.; Chari, A.; Brody, J.; Reddy, M.R.; Zhang, W.; et al. Dual Targeting of CDK4 and ARK5 Using a Novel Kinase Inhibitor ON123300 Exerts Potent Anticancer Activity against Multiple Myeloma. Cancer Res. 2016, 76, 1225–1236. [Google Scholar] [CrossRef] [Green Version]
  427. Murakami, Y.I.; Yatabe, Y.; Sakaguchi, T.; Sasaki, E.; Yamashita, Y.; Morito, N.; Yoh, K.; Fujioka, Y.; Matsuno, F.; Hata, H.; et al. c-Maf Expression in Angioimmunoblastic T-cell Lymphoma. Am. J. Surg. Pathol. 2007, 31, 1695–1702. [Google Scholar] [CrossRef] [PubMed]
  428. Missiroli, S.; Perrone, M.; Genovese, I.; Pinton, P.; Giorgi, C. Cancer metabolism and mitochondria: Finding novel mechanisms to fight tumours. EBioMedicine 2020, 59, 102943. [Google Scholar] [CrossRef]
  429. Condelli, V.; Crispo, F.; Pietrafesa, M.; Lettini, G.; Matassa, D.S.; Esposito, F.; Landriscina, M.; Maddalena, F. HSP90 Molecular Chaperones, Metabolic Rewiring, and Epigenetics: Impact on Tumor Progression and Perspective for Anticancer Therapy. Cells 2019, 8, 532. [Google Scholar] [CrossRef] [Green Version]
  430. Al-Azawi, A.; Sulaiman, S.; Arafat, K.; Yasin, J.; Nemmar, A.; Attoub, S. Impact of Sodium Dichloroacetate Alone and in Combination Therapies on Lung Tumor Growth and Metastasis. Int. J. Mol. Sci. 2021, 22, 12553. [Google Scholar] [CrossRef]
  431. Xintaropoulou, C.; Ward, C.; Wise, A.; Marston, H.; Turnbull, A.; Langdon, S.P. A comparative analysis of inhibitors of the glycolysis pathway in breast and ovarian cancer cell line models. Oncotarget 2015, 6, 25677–25695. [Google Scholar] [CrossRef] [Green Version]
  432. Zhang, L.; Zhang, J.; Ye, Z.; Townsend, D.M.; Tew, K.D. Pharmacology of ME-344, a novel cytotoxic isoflavone. Adv. Cancer Res. 2019, 142, 187–207. [Google Scholar] [CrossRef] [PubMed]
  433. Kim, E.-Y.; Chung, T.-W.; Han, C.W.; Park, S.Y.; Park, K.H.; Jang, S.B.; Ha, K.-T. A Novel Lactate Dehydrogenase Inhibitor, 1-(Phenylseleno)-4-(Trifluoromethyl) Benzene, Suppresses Tumor Growth through Apoptotic Cell Death. Sci. Rep. 2019, 9, 3969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  434. Veseli, B.E.; Perrotta, P.; Van Wielendaele, P.; Lambeir, A.; Abdali, A.; Bellosta, S.; Monaco, G.; Bultynck, G.; Martinet, W.; De Meyer, G.R.Y. Small molecule 3PO inhibits glycolysis but does not bind to 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase-3 (PFKFB3). FEBS Lett. 2020, 594, 3067–3075. [Google Scholar] [CrossRef] [PubMed]
  435. Shi, L.; Pan, H.; Liu, Z.; Xie, J.; Han, W. Roles of PFKFB3 in cancer. Signal Transduct. Target. Ther. 2017, 2, 17044. [Google Scholar] [CrossRef] [PubMed]
  436. Camarda, R.; Zhou, A.Y.; Kohnz, R.A.; Balakrishnan, S.; Mahieu, C.; Anderton, B.; Eyob, H.; Kajimura, S.; Tward, A.; Krings, G.; et al. Inhibition of fatty acid oxidation as a therapy for MYC-overexpressing triple-negative breast cancer. Nat. Med. 2016, 22, 427–432. [Google Scholar] [CrossRef]
  437. Thupari, J.N.; Pinn, M.L.; Kuhajda, F.P. Fatty Acid Synthase Inhibition in Human Breast Cancer Cells Leads to Malonyl-CoA-Induced Inhibition of Fatty Acid Oxidation and Cytotoxicity. Biochem. Biophys. Res. Commun. 2001, 285, 217–223. [Google Scholar] [CrossRef]
  438. Fu, Y.; Zou, T.; Shen, X.; Nelson, P.J.; Li, J.; Wu, C.; Yang, J.; Zheng, Y.; Bruns, C.; Zhao, Y.; et al. Lipid metabolism in cancer progression and therapeutic strategies. MedComm 2020, 2, 27–59. [Google Scholar] [CrossRef] [PubMed]
  439. Casara, P.; Davidson, J.; Claperon, A.; Le Toumelin-Braizat, G.; Vogler, M.; Bruno, A.; Chanrion, M.; Lysiak-Auvity, G.; Le Diguarher, T.; Starck, J.-B.; et al. S55746 is a novel orally active BCL-2 selective and potent inhibitor that impairs hematological tumor growth. Oncotarget 2018, 9, 20075–20088. [Google Scholar] [CrossRef] [Green Version]
  440. Zhu, J.; Xin, M.; Xu, C.; He, Y.; Zhang, W.; Wang, Z.; Zhuang, C. Ligand-based substituent-anchoring design of selective receptor-interacting protein kinase 1 necroptosis inhibitors for ulcerative colitis therapy. Acta Pharm. Sin. B 2021, 11, 3193–3205. [Google Scholar] [CrossRef]
  441. Nakamura, A.; Arita, T.; Tsuchiya, S.; Donelan, J.; Chouitar, J.; Carideo, E.; Galvin, K.; Okaniwa, M.; Ishikawa, T.; Yoshida, S. Antitumor Activity of the Selective Pan-RAF Inhibitor TAK-632 in BRAF Inhibitor-Resistant Melanoma. Cancer Res. 2013, 73, 7043–7055. [Google Scholar] [CrossRef] [Green Version]
  442. Dai, W.; Wang, G.; Chwa, J.; Oh, M.E.; Abeywardana, T.; Yang, Y.; Wang, Q.A.; Jiang, L. Mitochondrial division inhibitor (mdivi-1) decreases oxidative metabolism in cancer. Br. J. Cancer 2020, 122, 1288–1297. [Google Scholar] [CrossRef] [Green Version]
  443. Wu, D.; Dasgupta, A.; Chen, K.; Neuber-Hess, M.; Patel, J.; Hurst, T.E.; Mewburn, J.D.; Lima, P.D.A.; Alizadeh, E.; Martin, A.; et al. Identification of novel dynamin-related protein 1 (Drp1) GTPase inhibitors: Therapeutic potential of Drpitor1 and Drpitor1a in cancer and cardiac ischemia-reperfusion injury. FASEB J. 2020, 34, 1447–1464. [Google Scholar] [CrossRef] [Green Version]
  444. Guan, Y.; Wang, Y.; Li, B.; Shen, K.; Li, Q.; Ni, Y.; Huang, L. Mitophagy in carcinogenesis, drug resistance and anticancer therapeutics. Cancer Cell Int. 2021, 21, 350. [Google Scholar] [CrossRef] [PubMed]
  445. Liu, Y.; Liu, T.; Lei, T.; Zhang, D.; Du, S.; Girani, L.; Qi, D.; Lin, C.; Tong, R.; Wang, Y. RIP1/RIP3-regulated necroptosis as a target for multifaceted disease therapy (Review). Int. J. Mol. Med. 2019, 44, 771–786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The adaptation of mitochondrial metabolism in cancer cells. Black arrows: biochemical processes under normal circumstances. Red text and arrows: the mechanism of dysregulation in tumor tissue. Green text and arrows: possible therapeutic targets. Abbreviations: ATP: adenosine trisphosphate, ROS: reactive oxygen species, ETC: Electron transport chain. PDH: pyruvate dehydrogenase, KGDH: α-ketoglutarate dehydrogenase, IDH: isocitrate dehydrogenase, α-KG: α -ketoglutarate.
Figure 1. The adaptation of mitochondrial metabolism in cancer cells. Black arrows: biochemical processes under normal circumstances. Red text and arrows: the mechanism of dysregulation in tumor tissue. Green text and arrows: possible therapeutic targets. Abbreviations: ATP: adenosine trisphosphate, ROS: reactive oxygen species, ETC: Electron transport chain. PDH: pyruvate dehydrogenase, KGDH: α-ketoglutarate dehydrogenase, IDH: isocitrate dehydrogenase, α-KG: α -ketoglutarate.
Life 11 01351 g001
Figure 2. The regulation of mtDNA replication, transcription and translation and their role as therapeutic targets. Blue arrows: mtDNA replication, transcription, translation, and their regulators. Red text and arrows: inhibitors. Green text and arrow: mtDNA repair mechanisms.
Figure 2. The regulation of mtDNA replication, transcription and translation and their role as therapeutic targets. Blue arrows: mtDNA replication, transcription, translation, and their regulators. Red text and arrows: inhibitors. Green text and arrow: mtDNA repair mechanisms.
Life 11 01351 g002
Figure 3. Mitochondrial fission-fusion, mitophagy and their regulation. Red arrows: inhibition. Green arrows: activation. Abbreviations: Drp1: Dynamin-related protein 1, Fis1: fission protein 1, Mff: mitochondrial fission factor, Mief: mitochondrial elongation factors, MOMP: Mitochondrial outer membrane potential, OXPHOS: oxidative phosphorylation, Mfn1,2: mitofusin 1,2, Opa1: Optic atrophy 1, PINK1: phosphatase and tensin homologue (PTEN)-induced putative kinase 1, Nix: Nip3-like protein X, Bnip3: adenovirus E1B 19 kDa-interacting protein 3, Ambra-1: activating molecule in BECLIN1-regulated autophagy, Fundc1: FUN14 Domain Containing 1.
Figure 3. Mitochondrial fission-fusion, mitophagy and their regulation. Red arrows: inhibition. Green arrows: activation. Abbreviations: Drp1: Dynamin-related protein 1, Fis1: fission protein 1, Mff: mitochondrial fission factor, Mief: mitochondrial elongation factors, MOMP: Mitochondrial outer membrane potential, OXPHOS: oxidative phosphorylation, Mfn1,2: mitofusin 1,2, Opa1: Optic atrophy 1, PINK1: phosphatase and tensin homologue (PTEN)-induced putative kinase 1, Nix: Nip3-like protein X, Bnip3: adenovirus E1B 19 kDa-interacting protein 3, Ambra-1: activating molecule in BECLIN1-regulated autophagy, Fundc1: FUN14 Domain Containing 1.
Life 11 01351 g003
Figure 4. The interaction between mitochondria, epigenetic regulator and transcriptional factor mutations in hematologic malignancies. Blue arrows: connections in physiological function, Red arrows: effect of mutations, red dashed arrow: possible or not yet fully proved effect of mutation, green boxes/arrows: possible therapies. Abbreviations: DNMT: DNA-methyltransferases, TET: Ten eleven translocation enzymes, IDH: isocitrate dehydrogenases, α-KG: α-ketoglutarate, 5-MC: 5-methylcytosine, 5-hMC: 5-hydroxymethylcytosine.
Figure 4. The interaction between mitochondria, epigenetic regulator and transcriptional factor mutations in hematologic malignancies. Blue arrows: connections in physiological function, Red arrows: effect of mutations, red dashed arrow: possible or not yet fully proved effect of mutation, green boxes/arrows: possible therapies. Abbreviations: DNMT: DNA-methyltransferases, TET: Ten eleven translocation enzymes, IDH: isocitrate dehydrogenases, α-KG: α-ketoglutarate, 5-MC: 5-methylcytosine, 5-hMC: 5-hydroxymethylcytosine.
Life 11 01351 g004
Figure 5. The pathogenetic role of IDH1/2. Blue and yellow arrows, black text: biochemical processes. Green text and arrows: cofactors. Red lines and arrows: inhibition.IDH mutations in MDS and MDS/AML. The frequency of IDH1/2 mutations has been reported between 4 and 12% in MDS patients [156,157,158,159]. IDH2 mutations were particularly enriched in the RAEB subtype of MDS and were mutually exclusive with TET2 and SF3B1 mutations but were frequently associated with SRSF2 mutations. Many authors found that IDH1/2 mutations were associated with poor prognosis, particularly in low-risk MDS [160,161]. The proportion of IDH2 mutation, however, was higher than IDH1 mutation in high-risk MDS. Both IDH1 and IDH2 mutations showed a significantly shorter progression time to AML when associated with low-risk mutations, such as GATA2, NRAS, KRAS, RUNX1, STAG2, and ASXL1. Another study found that only IDH1 but not IDH2 mutations are associated with shortened leukemia-free survival [162]. In addition, the presence of certain subtypes of IDH2 mutations (for example IDH2-R172) is a predictor of poor response to chemotherapy [163]. Both IDH1 and IDH2 mutations are found in AMLs, with lower frequency in the pediatric patients than in the adult ones, the latter being about 20%. In therapy-related AML (t-AML), its frequency is around 7%, with IDH1 mutations being more frequent [159,164,165,166,167,168].
Figure 5. The pathogenetic role of IDH1/2. Blue and yellow arrows, black text: biochemical processes. Green text and arrows: cofactors. Red lines and arrows: inhibition.IDH mutations in MDS and MDS/AML. The frequency of IDH1/2 mutations has been reported between 4 and 12% in MDS patients [156,157,158,159]. IDH2 mutations were particularly enriched in the RAEB subtype of MDS and were mutually exclusive with TET2 and SF3B1 mutations but were frequently associated with SRSF2 mutations. Many authors found that IDH1/2 mutations were associated with poor prognosis, particularly in low-risk MDS [160,161]. The proportion of IDH2 mutation, however, was higher than IDH1 mutation in high-risk MDS. Both IDH1 and IDH2 mutations showed a significantly shorter progression time to AML when associated with low-risk mutations, such as GATA2, NRAS, KRAS, RUNX1, STAG2, and ASXL1. Another study found that only IDH1 but not IDH2 mutations are associated with shortened leukemia-free survival [162]. In addition, the presence of certain subtypes of IDH2 mutations (for example IDH2-R172) is a predictor of poor response to chemotherapy [163]. Both IDH1 and IDH2 mutations are found in AMLs, with lower frequency in the pediatric patients than in the adult ones, the latter being about 20%. In therapy-related AML (t-AML), its frequency is around 7%, with IDH1 mutations being more frequent [159,164,165,166,167,168].
Life 11 01351 g005
Table 1. Key oncogenic events in the pathogenesis of AML.
Table 1. Key oncogenic events in the pathogenesis of AML.
ClassKey Oncogenic Events
I.
Proliferative and survival advantages
FLT3, KIT, RAS, PNP11, JAK2, CBL, ERG, BAALC
II.
Alterations of cellular differentiation, apoptosis
PML-RARA, RUNX1, RUNX1T1, CBFB, MYH11, MLL, CEBPA, NPM1, TP53
III.
No classification (recently described, mainly epigenetic modulators)
DNMT3a, TET2, IDH1/2, ASXL1, WT1
Table 2. The effects of transcription factors on mitochondria. Transcription factors described in detail in this review are summarized here with their main effects on mitochondrial metabolism, dynamics, and apoptosis. In addition, the involvement in various hematologic malignancies is also summarized. ↑: increase, ↓: decrease. Abbreviations: AML: Acute myeloid leukemia, BL: Burkitt lymphoma, DLBCL: Diffuse large B-cell lymphoma, Drp1: dynamin-related protein1, GLS2: mitochondrial glutaminase 2, MCL: Mantle-cell lymphoma, Mff: mitochondrial fission factor, lMAFs: large Musculoaponeurotic fibrosarcoma (MAF) proteins, MDS: Myelodysplastic syndrome, MM: Multiple myeloma, OXPHOS: oxidative phosphorylation, ROS: reactive oxygen species, sMAF: small Musculoaponeurotic fibrosarcoma (MAF) proteins.
Table 2. The effects of transcription factors on mitochondria. Transcription factors described in detail in this review are summarized here with their main effects on mitochondrial metabolism, dynamics, and apoptosis. In addition, the involvement in various hematologic malignancies is also summarized. ↑: increase, ↓: decrease. Abbreviations: AML: Acute myeloid leukemia, BL: Burkitt lymphoma, DLBCL: Diffuse large B-cell lymphoma, Drp1: dynamin-related protein1, GLS2: mitochondrial glutaminase 2, MCL: Mantle-cell lymphoma, Mff: mitochondrial fission factor, lMAFs: large Musculoaponeurotic fibrosarcoma (MAF) proteins, MDS: Myelodysplastic syndrome, MM: Multiple myeloma, OXPHOS: oxidative phosphorylation, ROS: reactive oxygen species, sMAF: small Musculoaponeurotic fibrosarcoma (MAF) proteins.
GeneCoded ProteinEffect on Nuclear DNAMetabolic RoleMitochondrial RoleOncogenic Effect
MYC family: MYC
MYCN
MYCL
c-Mycp53 activationExpression of GLUT1, 3↑
Hexokinase activity ↑
Expression of lactate, glutamine transporters↑
Nucleotid synthesis ↑
Amino acid synthesis ↑
Fatty acid synthesis ↑
β-oxidation ↑
Glycolysis ↑
Glutaminolysis ↑
ROS generation ↑
OXPHOS ↑/↓
Expression of mitochondrial trafficking proteins (RHOT1, RHOT2, TRAK2, and Kif5B) ↑
Fission (Drp1, Mff) ↑
Mitochondrial recruitment to cortical cytoskeleton ↑
Organelle movements ↑
Mitochondrial biogenesis and replication (PLOG, PLOG2, and NRF1 expression) ↑
BL, Double hit/triple hit DLBCL, MM, MCL, AML
RUNX1RUNX1/
RUNX1T1 fusion protein
Local remodeling of chromatinCell differentiation ↓
Self-renewal ↑
(unknown mechanism)
Apoptosis ↓
Fission (Drp1-dependent) ↑
ROS production ↑
Granulopoiesis,
Dysplasia- MDS phenotype,
Development of post MDS-AML
TP53Tumor protein 53MYC-TP53 cross talk
Topoisomerase IIα stabilization,
Transcriptional inhibition of PFKFB3 and PFKFB4 genes,
PDK2 transcription ↓
Autophagy ↑
PUMA, NOXA ↑
GLUT 1,3,4 ↓
Hexokinase 2, Phosphoglycerate mutase 1 ↓
Parkin ↑
HIF1α degradation
Fructose-2,6-bisphosphate expression ↓
=promoting glycolysis
Pentose-phosphate-pathway ↓
Fatty acid synthesis ↓
Lipid synthesis ↑
Apoptosis ↑:
Bcl2/BclIX ↓
Bax/Bak ↑
Cytochrome C release ↑
Opa1 cleaving ↑
OXPHOS ↑:
Mitochondrial GLS2 ↑
Drp1 blocking-highly interconected mitichondria ↑
PKA activation-mitochondrial elongation
Fission ↑
ROS synthesis ↓
Li-Fraumeni syndrome,
MDS, AML
RAS family:
HRAS
NRAS
KRAS
small GTPasesReplication stress
RAS/MYC functional connection
Accelerated transcription via TBP
Major downstream RAS effector pathways:
phosphatidylinositiol-3-kinase (PI3K),
mitogen-activated protein kinase (MAPK), Ras-like (Ral) small GPTase signaling pathways
Glycolysis ↑ (via STAT3)
OXPHOS ↓(via STAT3)
ROS synthesis ↑
Fission (Drp1-dependent) ↑
AML, MDS
MAFMAF proteins
lMAFs (oncogenes): MAFA, MAFB, c-MAF, NRL
sMAFs
(transcriptional factors): MAFF, MAFG, MAFK
lMAFs: CCDN2, FOS, JUN, CREB, MTORC2, ATF expression modulation
sMAFs: inhibiting transcription
insulin secretion ↑
cell glucose uptake ↑
glutamine metabolism ↑
OXPHOS ↑
Apoptosis ↑
MM,
Angioimmunoblastic T-cell lymphoma
Table 3. Summary of drugs targeting mitochondria and glycolysis.
Table 3. Summary of drugs targeting mitochondria and glycolysis.
Group of DrugDrug/sMechanism of Action
Mitochondrial metabolism:ETC inhibitorsMetformin
IACS10759
BAY87-2243
MitoTam
Respiratory Complex I inhibition [7,428]
MitoVESRespiratory Complex I-II inhibition [428]
LonidamineRespiratory Complex II inhibition [428]
ME344Respiratoy Complex I-IV inhibition [7,432]
VLX600ETC inhibitor [428]
Mitochonrial metabolism:TCA enzyme inhibitorsCPI-613 (devimistat)PDH and KGDH inhibition [7,428]
Enasibenib
Ivosibenib
IDH inhibition (mutated IDH) [7,428]
Metabolism: glycolysis and other pathway inhibitorsDichloroacetate (DCA)PDH kinase inhibitor [428,430]
WZB117
STF31
Phloretin
Quercetin
GLUT1 inhibitors [431]
2-DG (2-Deoxy-D-glucose)
2-FDG (2- fluorodeoxy-D- glucose)
Competitors for binding hexokinase (converting glucose to glucose-6-phosphate) [428]
3-BromopyruvateHexokinase inhibitor [431]
Diclofenac
Lumiracoxib
Anti-inflammatory drugs with glycolysis inhibition [428]
3PO
PFK-158
Phosphofructokinase inhibition (phosphofructokinase converts fructose-6-P to fructose-1,6-bisP), inhibiting glycolysis [431,434,435]
Oxamic acid
NHI1
1-(Phenylseleno)-4-(trifluoromethyl)benzene
Gossypol
Lactate dehydrogenase inhibitor (lactate dehydrogenase converts pyruvate to lactate), inhibiting glycolysis [428,431,433]
CB-839
Compound 27
968
BPTES (bis-2-(5-phenylacetamido-1,2,4-thiadiazol-2-yl)ethyl sulfide)
Glutaminase inhibitor (converts glutamine to glutamate) [7,14]
L-AsparginaseGlutamine depletion [14]
Sulfasalazine
Erastin
Glutamine-cystine antiporter (xCT, a heterodimer of SLC7A11 and SCL3A2) inhibitor [14]
GamitrinibHeat shock protein 90 (HSP90) inhibitor, inhibiting various metabolic pathways [428,429]
Etomixircarnitine palmitoyltransferase-1 (CPT-1) inhibition (fatty acid oxydation inhibitor) [437]
C75
Cerulein
Orlistat
Triclosan
Amentoflavone
EGCG
TVB-3166
Fatty acid synthase inhibitors [438]
Increased ROS productionCisplatin
5-FU
Paclitaxel
Procarbazine
ROS production [7,20]
mtDNA transcription and translation inhibitionIMT1
IMT1B
Mitochondrial RNA polymerase (POLMRT) inhibition (transcription inhibition), interfering with ETC protein transcription [29]
TigecyclinemtDNA translation inhibition [7,30,31]
Apoptosis inductorsVenetoclax (ABT-199)
BCL201 (aka S55746)
Bcl-2 inhibitors [45,439]
BTSA1Bax activator [43]
Necroptosis inhibitionTAK-632
Ponatinib
RIP1 and RIP3 inhibitiors [440,445] (with TAK-632 also being a pan-RAF inhibitor [441])
Nec-1
GSK2982772 (Compound 5)
RIPA-56 (Compound 56)
7-oxo-2,4,5,7-tetrahydro-6H-pyrazolo [3,4-c]pyridine (Compound 22)
Tozasertib (a.k.a. VX-680 and MK-0457)
Pazopanib
RIP1 inhibitors [445]
DabrafenibRIP3 and B-Raf inhibitor [445]
Drugs altering mitochondrial dynamicsMitochondrial division inhibitor-1 (Mdivi-1)
Drpitor1 and Drpitor1a
Mitochondrial fission inhibitors [442,443]
Melatonin
Mdivi-1
Liensinine
Mitophagy inhibitors (helping downregulating drug resistance in certain tumors) [444]
Ketaconazole
Sorafenib
Mito-CP and Mito-Metformin
Mitophagy inducers (leading to apoptosis due to insufficient number of mitochondria) [444]
Abbreviations: ETC: electron transport chain, IDH: isocitrate dehydrogenase, KGDH: α-ketoglutarate dehydrogenase, mtDNA: mitochondrial DNA, PDH: pyruvate dehydrogenase, ROS: reactive oxygen species, TCA: tricarboxylic acid cycle.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Czegle, I.; Gray, A.L.; Wang, M.; Liu, Y.; Wang, J.; Wappler-Guzzetta, E.A. Mitochondria and Their Relationship with Common Genetic Abnormalities in Hematologic Malignancies. Life 2021, 11, 1351. https://doi.org/10.3390/life11121351

AMA Style

Czegle I, Gray AL, Wang M, Liu Y, Wang J, Wappler-Guzzetta EA. Mitochondria and Their Relationship with Common Genetic Abnormalities in Hematologic Malignancies. Life. 2021; 11(12):1351. https://doi.org/10.3390/life11121351

Chicago/Turabian Style

Czegle, Ibolya, Austin L. Gray, Minjing Wang, Yan Liu, Jun Wang, and Edina A. Wappler-Guzzetta. 2021. "Mitochondria and Their Relationship with Common Genetic Abnormalities in Hematologic Malignancies" Life 11, no. 12: 1351. https://doi.org/10.3390/life11121351

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