Next Article in Journal
Amyloid-Beta Peptides and Activated Astroglia Impairs Proliferation of Nerve Growth Factor Releasing Cells In Vitro: Implication for Encapsulated Cell Biodelivery-Mediated AD Therapy
Next Article in Special Issue
Modulation of FLT3-ITD Localization and Targeting of Distinct Downstream Signaling Pathways as Potential Strategies to Overcome FLT3-Inhibitor Resistance
Previous Article in Journal
Adenosine Receptor Antagonists to Combat Cancer and to Boost Anti-Cancer Chemotherapy and Immunotherapy
Previous Article in Special Issue
Distinct Clinical Impact and Biological Function of Angiopoietin and Angiopoietin-like Proteins in Human Breast Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Role for the Bone Marrow Microenvironment in Drug Resistance of Acute Myeloid Leukemia

by
Seyed Mohammadreza Bolandi
1,2,†,
Mahdi Pakjoo
3,†,
Peyman Beigi
3,
Mohammad Kiani
2,
Ali Allahgholipour
2,
Negar Goudarzi
1,
Jamshid S. Khorashad
4 and
Anna M. Eiring
5,*
1
Department of Immunology, Razi Vaccine and Sera Research Institute, Karaj, Iran
2
Department of Pharmacology, Karaj Branch, Islamic Azad University, Karaj, Iran
3
Department of Hematology, Faculty of Medical Sciences, Tarbiat Modares University, Tehran, Iran
4
Centre for Haematology, Hammersmith Hospital, Imperial College London, London W12 0HS, UK
5
Center of Emphasis in Cancer, Department of Molecular and Translational Medicine, Texas Tech University Health Sciences Center at El Paso, El Paso, TX 79905, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2021, 10(11), 2833; https://doi.org/10.3390/cells10112833
Submission received: 22 September 2021 / Revised: 13 October 2021 / Accepted: 14 October 2021 / Published: 21 October 2021
(This article belongs to the Collection Emerging Cancer Target Genes)

Abstract

:
Acute myeloid leukemia (AML) is a heterogeneous disease with a poor prognosis and remarkable resistance to chemotherapeutic agents. Understanding resistance mechanisms against currently available drugs helps to recognize the therapeutic obstacles. Various mechanisms of resistance to chemotherapy or targeted inhibitors have been described for AML cells, including a role for the bone marrow niche in both the initiation and persistence of the disease, and in drug resistance of the leukemic stem cell (LSC) population. The BM niche supports LSC survival through direct and indirect interactions among the stromal cells, hematopoietic stem/progenitor cells, and leukemic cells. Additionally, the BM niche mediates changes in metabolic and signal pathway activation due to the acquisition of new mutations or selection and expansion of a minor clone. This review briefly discusses the role of the BM microenvironment and metabolic pathways in resistance to therapy, as discovered through AML clinical studies or cell line and animal models.

1. Introduction

Hematopoietic stem cells (HSCs) produce all blood cell types throughout life due to their capacity for self-renewal and differentiation [1,2]. Any disruption of this process can lead to abnormal expansion of cellular clones, which may lead to hematologic malignancies such as acute myeloid leukemia (AML) [2,3,4,5]. AML is a heterogeneous disease with extreme proliferation of myeloblasts (>20%) in the bone marrow (BM) [6,7]. AML is responsible for 1% of all annual new cancer cases and 1.8% of all cancer deaths in the United States (US). AML is a male predominant disease, with a risk ratio of 1.6 for males and 1.2 for females [3]. It is among the top 15 most prevalent cancers, with an average age of 70 years at diagnosis [8]. Morbidity and mortality of AML increase with age [9], and the global AML incidence has progressively increased during the last several decades (from 63,840 cases in 1990 to 119,570 cases in 2017) [10]. In children, AML is the most common leukemia after acute lymphoblastic leukemia (ALL), with a five-year survival rate of 64% [11]. The best prognosis among the AML subtypes is acute promyelocytic leukemia (APL), which harbors the t(15;17) translocation, generating the promyelocytic leukemia (PML)-retinoic acid receptor alpha (RARA) fusion gene, and is curable with arsenic trioxide and all-trans retinoic acid (ATRA) treatment. The worst survival rate among the AML subtypes is in patients with FMS-like tyrosine kinase 3 (FLT3) mutations, monosomy 7, and del 5q [11,12,13]. Moreover, childhood AML prevalence is highest among newborns less than one year of age, with an incidence rate of 18.4 per million [11].
Although a diverse range of treatment options for AML have been introduced over the past several decades, the health care community is still struggling to improve the poor prognosis, especially in elderly patients [14]. The well-known 7+3 induction chemotherapy is the most common approach for non-APL disease, which is based on three days of Anthracyclines (in most cases Daunorubicin) accompanied with seven days of continuous infusion with a pyrimidine analog like Cytarabine [15]. After the achievement of complete remission (CR), hematopoietic stem cell transplantation (HSCT) and/or intermediate to high dose Cytarabine is prescribed as consolidation therapy [16]. However, AML has the shortest overall survival (OS) among the acute leukemias, with a 2-year and 5-year OS of only 32% and 24%, respectively [14]. To be more specific, relapse and primary (initial) refractory AML are indispensable challenges in the treatment of AML. Indeed, 10–40% of younger patients (<45 years) and more than 60% of elderly AML patients (>60 years) are primarily refractory to initial induction chemotherapy. A significant proportion of AML patients relapse, even those who achieve CR. AML relapse is due to various factors, such as dysregulation of the signaling pathways associated with DNA damage response sensing proteins, mutations in cell cycle control genes, changes in programmed cell death (including apoptosis and autophagy), altered anti-cancer drug trafficking, and other mechanisms that still need to be discovered [17,18]. Another important reason why many patients relapse is the inability of most therapies to target the leukemic stem cell (LSC) population [19].
The etiology of AML is not completely understood. AML is generally categorized into three groups: (1) de novo AML (initially diagnosed with AML), (2) secondary AML (myeloid disorders that develop after other diseases, such as myelofibrosis, chronic myeloid leukemia, or myelodysplastic syndromes), and (3) therapy-related AML (t-AML) (following chemical exposure) [20]. AML has been associated with risk factors such as old age, male gender, smoking, chemicals (e.g., benzene and formaldehyde), genetic disorders (e.g., Fanconi anemia (FA) and Bloom syndrome), radiation, AML familial history (mutations in GATA Binding Protein 1 (GATA1), DEAD-box helicase 41 (DDX41), runt-related transcription factor1 (RUNX1), CCAAT/enhancer-binding protein alpha (CEBPA), and Ankyrin repeat domain 26 (ANKRD26)), as well as chemotherapeutic agents (alkylating agents and topoisomerase II inhibitors) [21]. In the present review, we discuss the various mechanisms contributing to drug resistance in AML, including both intrinsic and extrinsic mechanisms that have been discovered through animal models or clinical investigations.

2. Genomic and Immunophenotypic Characteristics

General symptoms of AML include fatigue, shortness of breath, bruising and recurrent infections that are consequences of anemia, thrombocytopenia, and neutropenia [22]. For initial diagnosis, BM aspiration is performed to assess morphology, molecular genetic tests, cytogenetic analysis, cytochemistry (including myeloperoxidase (MPO) activity), and immunophenotyping (e.g., CD34, CD13, CD33, CD113, and CD117) [22]. Metastasis is rarely seen in AML; however, it is mostly related with monocytic lineage infiltration in monoblastic/monocytic AML (AML-M4/M5 FAB category), which may lead to gingival hyperplasia or myeloid sarcoma within the central nervous system (CNS), abdomen, ovaries, muscles, and lungs in AML, especially for patients with the t(8;21) translocation (AML-M2 FAB category) [23,24,25].
Genomic analyses have revolutionized AML diagnosis and prognosis [26]. According to the latest world health organization (WHO) categorization, 85% of AML patients show one or more of the genomic abnormalities presented in Table 1 [27]. During the immunophenotypic analysis of AML, CD34 and CD117 are the antigens commonly used to detect myeloblasts [28]. CD13, CD15, CD33, MPO, and CD16 are myeloid markers commonly used for lineage assignment, along with monocytic differentiation markers such as CD11b, CD64, CD14, and CD4 [28]. Erythroid precursors express CD71, CD105, CD117, CD235a, and CD36, whereas megakaryocytic precursors express CD61 and CD42b [28]. In AML, an increase of the immature myeloid population must be confirmed through diagnosis of at least two markers, including MPO, CD33, CDw65, and CD117 [22]. At least one pan myeloid marker (CD13, CD33, and CDw65) is seen in 95% of cases, whereas all three markers can be found in ~50% of cases. Lymphoid markers such as CD3, CD2, CD4, CD5, CD56, CD22, and CD79a are expressed in almost 25% of cases, whereas the CD7 and CD19 markers can be found in 10–30% and <3% of patients, respectively [22,28].

3. Treatment

According to European Leukemia Net (ELN), AML prognosis using cytogenetic and molecular analysis is divided into four groups, including favorable, intermediate I, intermediate II, and adverse [29]. From this group, patients older than 60 years of age show the worst prognosis [20]. AML treatment is generally associated with poor outcomes, even in young patients using high dose chemotherapy and HSCT [30]. Drug resistance and low five-year survival is a main feature of AML. In patients <70 years of age, the five-year survival is nearly 40%, but in patients older than 70 years, the three-year survival does not go beyond 10% [3,30,31,32,33]. Recent advances in chemotherapy, immunotherapy, HSCT, and targeted therapy have led to improvements in AML treatment [34]. The 7+3 regimen is the first choice of AML therapy, which includes seven days of Daunorubicin or Idarubicin and 3 days of Cytarabine administration [34,35,36]. This regimen is the most effective approach for patients in the favorable prognosis category (below 60 years and/or with Core binding factor (CBF)/Nucleophosmin 1 (NPM1) translocation) [20]. Despite its widespread use, this regimen is unfortunately associated with increased toxicity and often fails to eradicate the LSC population, resulting in many cases of relapsed or refractory AML [31,37]. In addition to conventional therapies for AML, novel agents have been introduced due to the identification of underlying genomic abnormalities, such as Midostaurin in the case of AML patients with FLT3 mutations [20].
HSCT, targeted therapy, or other types of chemotherapy are mainly post-induction treatment strategies based on the patient’s status, AML type, and appropriate HSC donor availability [20]. To perform HSCT, morphologic complete remission (M-CR) must be achieved. M-CR means that blasts in the BM must be less than 5% among at least 200 nucleated cells, there should be no sign of extramedullary or persistent disease, and platelet and neutrophil absolute count must be more than 100,000 and 1000 per microliter, respectively [20]. To monitor minimal residual disease (MRD) and treatment response, methods such as morphologic assessment, multiparameter flow cytometry, digital droplet PCR (ddPCR), real-time quantitative (RTq)–PCR, and next generation sequencing (NGS) are applied [20,38]. For HSCT, standard myeloablative conditioning (MAC-HSCT) regimens in AML include Cyclophosphamide and total body irradiation (TBI) or Cyclophosphamide and Busulfan or Fludarabine and Busulfan [39], which is not recommended in patients older than 70 years due to the possibility of toxicity. Therefore, only a small proportion of patients can benefit from this approach [39,40]. While HSCT is the only definitive cure for AML, it is accompanied by graft-versus-host disease (GVHD) as the most major chronic side effect and the prognosis after HSCT remains poor [40,41,42]. 27–35% of younger patients with de novo AML and 38–62% of patients older than 60 years of age are deprived of HSCT because they fail to achieve M-CR [20].
Poor response to conventional therapies, and the side effects associated with them, have led to diverse therapeutic strategies and novel agents which are hoped to improve survival. Targeted therapy in AML is considered the next game changer of the field when cytogenetic and molecular abnormalities provide an actionable target. The selection of treatment for many cases would be based on the individual characteristics of the disease, indicating personalized medicine as the evolving approach for management of AML cases [43]. Based on this, new inhibitors have been developed according to the known target, such as immunotherapy to target specific intra- or extra-cellular antigens.
Genomic alterations in FLT3, NPM1, DNA methyl transferase 3A (DNMT3A), tumor protein 53 (TP53), TET methyl cytosine dioxygenase 2 (TET2), and isocitrate dehydrogenase (IDH1/2) are frequently observed in AML [44,45]. In recent years, some new medications, including Midostaurin (FLT3 inhibitor), Gilteritinib (FLT3 inhibitor), CPX-351, Gemtuzumab-Ozogamicin (anti-CD33 monoclonal antibody conjugated with calicheamicin), Enasidenib (IDH2 inhibitor), Ivosidenib (IDH1 inhibitor), Venetoclax (B-cell lymphoma 2 (BCL-2) inhibitor), and Glasdegib (Smoothened (SMO) inhibitor), have been approved by the Food and Drug Administration (FDA) to be used for AML treatment [46], all of which are targeted therapies aimed at personalizing the approach to management of AML [8]. In this approach, drugs are administered based on the patient’s individual condition after molecular analysis, age, clinical status, chemotherapy history, and bone marrow dysplastic alterations are identified [8]. Some promising drugs that inhibit specific markers to overcome AML are shown in Table 2.

4. Resistance

Many patients who achieve CR will relapse in less than three years while exhibiting drug resistance and poor prognosis [49]. Relapse is usually diagnosed via clonal expansion of minor pre-existing clones, or through detection of novel mutations acquired by the leukemic cells, which can be more aggressive if they develop in less than six months following treatment [20]. Drug resistance is usually categorized as primary or secondary (acquired) [34]. Primary drug resistance is usually defined as de novo lack of response to treatment and is related to the patient’s leukaemia genotype, availability of the target for the applied drug, or the G0 cell cycle phase of the LSC population. Secondary resistance, on the other hand, indicates a gradual loss of sensitivity to the drug after an initial response. This is associated with disease evolution through the development of escape mechanisms, such as new mutations which lead to recruiting or blocking signaling pathways, or enhanced production of cytokines, interleukins, or growth factors [34,50].
LSCs remain a major obstacle in the way of achieving complete remission in AML [51,52]. Recent studies have revealed that the leukemic niche plays a crucial role in AML persistence by nesting of LSCs and protecting them from both the immune system and therapeutics [53]. LSCs are considered to be responsible for AML initiation, chemotherapy resistance, disease progression, and MRD due to their quiescence and higher self-renewal capabilities [53,54]. LSCs may originate from HSCs or HPCs that acquire the ability of self-renewal upon oncogenic alterations [55]. Generally, abnormal proliferation, disruption of differentiation, and maturation arrest are consequences of events like TET2, NPM1, DNMT3A, IDH1, and IDH2 mutations, which can turn normal HSCs into pre-leukemic cells and finally leukemic cells [5,56,57]. LSCs may reside at the level of the CD34+38 or CD34+38+ cell fraction [55]. The common specificities of stem cells, such as self-renewal capacity, multi-drug resistance, and immaturity, enable them to initiate leukemia in immunosuppressed mouse models of the disease [58,59]. Specific markers of LSCs have not been completely defined due to the similarities with normal HSCs; however, a variety of expressed markers have been identified among AML patients [59,60]. During leukemic transformation, LSCs deploy various molecules and immune suppressor cytokines to alter vital regulatory mechanisms within the BM microenvironment [61], leading to failure of the immune system to maintain normal hematopoiesis [61]. LSCs escape the effects of cytotoxic agents by nesting in hematopoietic niches within the BM microenvironment [53,62].
AML cells can have a negative influence on normal haematopoiesis. In the beginning, initial leukemic stem cells (pre-LSCs) and HSCs are both located in the same microenvironment. However, leukemic cells gradually occupy and change the hematopoietic niche [63]. Kumar et al. indicated that leukemic cells can mediate molecular changes in the BM niche and convert the normal hematopoietic niche into the leukemic niche, which supports leukemic cell survival and growth [64]. In addition, leukemic cells decrease the capacity of the niche to maintain HSCs and block normal hematopoiesis [13,65]. Xenograft models of AML have shown that CXCR4-expressing leukemic cells compete with normal HSCs to bind CXCL12-expressing BM endothelial cells. This causes a reduction in normal hematopoiesis and a decreased response to therapy, indicating an important role for the BM microenvironment in AML therapeutic responses [66,67]. In AML patients, the expression of the Jagged-1, Hes-1, Hes-5, and NOTCH signaling pathways in mesenchymal stem cells (MSCs) was demonstrated to be reduced, and their co-culture with normal HSCs inhibited normal hematopoiesis [68]. Additionally, alterations of transcription factors (TFs) may be responsible for drug resistance in AML LSCs by upregulating ABC transporters, cell cycle progression molecules, and oxidant protection [53,69,70]. Transcription factors that play an important role in AML drug resistance are listed in Table 3.

5. The Normal BM Microenvironment

The bone marrow is a heterogeneous environment that contains various hematopoietic and non-hematopoietic cells, including HSCs and MSCs, also called stromal stem cells (SSCs) (Table 4) [86]. HSCs nest in hematopoietic niches of the BM, but their proliferation and quiescence are under the control of non-hematopoietic niches. However, under stress, they can migrate to different organs like the spleen to continue hematopoiesis [87]. The hematopoietic niche is divided into the endosteal niche and vascular niche (Figure 1) [88]. These two HSC niches differ in many aspects, including calcium levels, oxygen pressure, pH, and cellular variability [88]. Endosteal niches contain quiescent and radiation-resistant HSCs [88], whereas both quiescent and proliferating HSCs can be found within the vascular niche [88]. HSC niches are regulated by non-hematopoietic cells to produce a wide variety of blood cells [87], and MSCs form a primary part of the non-hematopoietic BM niche [89]. These cells are responsible for regulating various functions of HSCs, such as proliferation, differentiation, adhesion, and quiescence through deploying different cytokines, chemokines, and adhesion molecules [89].
In the normal BM microenvironment, HSCs are mostly in a quiescent phase (G0) through the action of factors like stem cell factor (SCF), transforming growth factor β (TGF-β), platelet factor 4 (PF4, CXCL4), angiopoietin-1 (ANGPT1), and thrombopoietin (TPO), and this quiescence is considered a protective mechanism against the destructive effects of the environment and chemotherapy [90]. In addition, SDF-1 (CXCL12) and its receptor CXCR4, both important for HSC nesting, are incorporated with the MSC-secreted cytokines, interleukin (IL)-6 and IL-8, to promote HSC survival [91,92]. Other complementary factors in HSC nesting include VCAM-1, extracellular matrix (ECM), selectins, and hyaluronic acid [91,92]. Finally, NOTCH ligand (NOTCH-L), IL-7, erythropoietin (EPO), and other factors direct the fate and terminal differentiation of cells [93]. Cross-talk and interrelationship between immune cells, dendritic cells (DCs), HSCs, and myeloid-derived suppressor cells (MDSCs) within the bone marrow niche make a regulatory network for apoptosis, proliferation, HSC protection, and homeostasis [61,94]. This cooperation between myeloid and lymphoid lineages regulates HSC differentiation, self-renewal, and proliferation to inhibit leukemia development [61].
Table 4. The function of various cellular components of the BM in normal and AML status.
Table 4. The function of various cellular components of the BM in normal and AML status.
CellNormal Function and ProductsRole in AMLRefs
Adipocyte1. Increases in adulthood
2. Adipokine and Adiponectin
3. Hematopoiesis negative regulation
1. Leukemic cells proliferation
2. Increased adipokinase during leukemia
3. Leukemic cell pro-survival
[44,62,87,89,95]
Endothelial cell1. Notch L
2. E-selectin, P-selectin
3. Vascular cell adhesion molecule 1 (VCAM 1)
4. Intercellular adhesion molecule 1 (ICAM -1)
1. Vascular endothelial growth factor (VEGF) production and Granulocyte-macrophage colony-stimulating factor (GM-CSF) (potential mitogen) stimulation
2. AML progression
[87,89,95,96]
Osteoblast1. N-Cadherin
2. Osteopoietin
3. SCF
4. CXCL12
5. HSC niche establishment
1. Osteogenesis augmentation
2. AML initiation and progression
[44,86,89,97]
CXCL12-abundant reticular cells (CAR cells)1. Stromal cell-derived factor 1(SDF-1)
2. VCAM-1
3. E-/P-Selectin
4. CD44
5. Platelet-derived growth factors (PDFG)
Pro-survival[44,62,87,96]
Regulatory T cells (T-reg)1. IL-10
2. IL-35
3. Inhibits immune reactions against stem cells
1. Up-regulated in AML patients
2. AML leukemic cells induce IL-10 secreting T regulatory (iTreg) cells and natural T regulatory (N-Treg) cells through inducible co-stimulator ligand (ICOSL) expression.
Fibroblast1. Cancer-associated fibroblasts (CAFs)
2. Growth differentiation factor 15 (GDF15)
3. IL-8
Chemotherapy resistance[44,95,98]

6. Role of the BM Microenvironment in AML and Therapy Resistance

Leukemic cells charter a highly disciplined and complex network within the BM microenvironment, especially MSCs, in order to survive and thrive. The BM microenvironment provides leukemic cells with sites to adhere to and tools for suppression of the immune system. Some studies have demonstrated that different aspects of leukemic cell characteristics, such as survival, invasion, growth, angiogenesis, proliferation, apoptosis, and signaling pathways are directly affected by non-hematopoietic cells [52,84,89,93,99,100,101,102,103]. Various cellular components, cytokines, and chemokines that impact AML initiation and therapy resistance at the cellular and molecular level are shown in Table 4 and Table 5.
AML alters the function of the BM stromal cell (BMSC) population to reshape the BM microenvironment, which in return promotes AML tumor cell survival and proliferation. AML cells induce senescence in BMSCs, as demonstrated by increased p16INK4a, β-Galactosidase, and IL-6, and reduced Lamin B [137]. The p16INK4a-driven senescence in BMSC increases the survival and proliferation of AML cells in return [138]. The increased p16INK4a in BMSC seems to be independent of direct cell-cell contact, and is rather due to cytokine secretion. In vivo and in vitro data showed that depletion of non-malignant BMSCs has anti-leukemia activity, and can therefore be considered a therapeutic option [138]. Induction of p16INK4a in BMSCs and subsequent senescence has been shown to be due to superoxide, a type of reactive oxygen species (ROS). The production of ROS by AML cells appears to be through the activity of NADPH oxidase 2 (NOX2) [138].
During leukemic transformation within the BM niche, MSCs are altered to make the entire niche appropriate for leukemogenesis [52]. The close relationship between leukemic cells and the stromal cells of the BM is essential for the development of drug resistance [88]. Stromal cells utilize two mechanisms to induce drug resistance, including soluble factor-mediated drug resistance (SM-DR) and cell adhesion-mediated drug resistance (CAM-DR) [139]. SM-DR includes soluble factors like CXCL12, vascular endothelial growth factor (VEGF), IL-6, fibroblast growth factor (FGF), granulocyte-colony stimulating factor (G-CSF), and other factors mentioned in Table 6. CAM-DR, on the other hand, is caused by direct cell-cell interactions (Table 6) [139]. In vitro assays demonstrated that the co-culture of AML and stromal cells leads to stroma-derived soluble factor (SDSF) secretion, resulting in MAPK/ERK kinase (MEK) pathway activation in leukemic cells and consequently increased survival [104,140]. Additionally, co-culture of apoptosis repressor with caspase recruitment domain (ARC)/IL-1β-expressing MSCs with AML cells upregulates cyclooxygenase-2 (COX-2) and prostaglandin E2 (PGE2) expression in MSCs. The IL-1β-mediated induction of PGE2 secretion from MSCs leads to β-catenin activation and the induction of malignant transformation of HSCs, up-regulation of ARC, and enhanced chemotherapy resistance in AML [141]. Conversely, β-catenin blockage leads to ARC decline and chemo-sensitization [141].
One of the findings in the BM of AML patients is the failure of normal hematopoiesis. BM failure is not due to depletion of HSC numbers, but rather due to failure of the BM to produce sufficient numbers of progenitor cells [152]. The MSCs seem to play a major role in blocking the transition from HSCs to progenitors in the BM of AML patients. Recent data suggest that hypoxia in the BM of AML patients activates hypoxia-associated molecules, such as stanniocalcin1 (STC1), which is secreted from MSCs and increases the stemness of normal HSCs, thereby preventing differentiation [153].
Signaling pathways are another part of this regulatory network, allowing the microenvironment to control leukemia cell behavior and vice versa. Interruptions in any of these pathways may lead to cross-talk imbalance and the development of leukemia [154,155,156]. Dysregulation of various signaling pathways have been shown to be responsible for the aberrant self-renewal in leukemic cells, leading to poor prognosis and chemotherapy resistance in many AML cases [157,158,159]. Some effects of signal pathway disruption are presented in Table 6 and Figure 2.
A recent report by Forte et al. showed the role of nestin-positive (nestin+) MSCs in AML development and resistance to chemotherapy [160], providing a rich niche for the HSCs and LSCs. In contrast with chronic myeloid leukemia (CML), where there is a reduced number of nestin+ MSCs [161], there is an enrichment of nestin+ cells in AML bone marrow, and this enrichment is essential for the viability and proliferation of AML cells in vitro and in vivo [160,162]. In addition to their role in the development of AML, nestin+ MSCs were demonstrated to induce resistance to chemotherapy through enhanced glutathione (GSH)-peroxidase (Gpx) activity. AML LSCs were recently shown to increase their metabolic activity through enhanced oxidative phosphorylation (OXPHO) and increased tricarboxylic acid (TCA) cycle in mitochondria. This increased reliance on mitochondrial activity is further provided by transfer of mitochondria from nestin+ MSCs directly to the AML cells. Increased metabolism leads to increased ROS production, which must be controlled or it is lethal to the cells, and therefore the antioxidant glutathione pathway is induced in AML cells by nestin+ MSCs through activating GSH-Gpx [160].
Indirect connections between leukemic cells and the microenvironment is in part regulated by cellular vesicles which are divided into exosomes, exomers, microvesicles, and apoptotic bodies, based on their size or source [163,164]. Exosomes are secreted by normal and/or leukemic cells, and in contrast to their size (30–100 nm), contain various mRNAs, microRNAs, long non-coding RNAs, and proteins (i.e., cytokines) that play important roles in regulating cell proliferation, differentiation, and apoptosis [165,166]. Exosomes carry factors like Fas Ligand (FAS-L), NPM1, FLT3, Matrix Metallopeptidase 9 (MMP9), insulin-like growth factor type 1 receptor (IGF1-R), CXCR4, and chaperones to alter the BM microenvironment, improve leukemic cell survival, and extrinsically mediate drug resistance in primarily sensitive AML [165,167,168]. The exosomes are identified by markers such as ALG-2 interacting protein X (ALIX), CD63, CD81, CD9, syndecan-1, tumor susceptibility gene 101 (TSG 101), major histocompatibility complex (MHC) molecules, and heat shock protein 70 (HSP 70) [165].
Recent data suggests that other tissue microenvironments may also contribute to drug resistance in AML. For instance, it was reported that the liver niche promotes proliferation of resident leukemic cells and prevents their apoptosis through regulating their polyunsaturated fatty acid (PUFA) metabolism, leading to activation of the ERK pathway to promote the stability of the anti-apoptotic proteins, BCL-2 and BCL-XL. Additionally, infiltrating AML cells caused damage to hepatocytes, resulting in the secretion of cytidine deaminase (CAD) from the damaged hepatic cells. The released CAD destroys chemotherapeutic agents, thereby leading to drug resistance. [169].

7. Metabolic Pathways, AML LSC Survival, and Resistance to Therapy

Venetoclax in combination with hypomethylating agents has been approved for the treatment of both newly diagnosed and relapsed/refractory AML patients [170]; however, 30% of patients show primary resistance and many others develop resistance following treatment [171]. Primary AML cells cannot effectively use common metabolic fuels such as glucose or fatty acids, but have an aberrant reliance on the uptake and catabolism of amino acids to drive the TCA cycle and promote OXPHOS. The combination of Venetoclax and Azacytidine (ven/az) inhibits amino acid metabolism, leading to reduced OXPHOS and LSC death [172]. However, ven/az is ineffective at relapse because the LSCs change their metabolic preferences and requirement for amino acids. At relapse, LSCs increase their energy production and, in addition to amino acids, use fatty acids as sources for the increased activity of the TCA cycle. The enhancement of TCA cycle activity depends on nicotinamide adenine dinucleotide (NAD+)-dependent TCA cycle enzymes, which require higher NAD+ levels for their activity. NAD+ is produced through salvage pathways from nicotinamide during relapse [173]. Primary AML patient cells were found to produce high levels of superoxide, a phenomenon that could be related to cell proliferation [174]. AML LSCs and their progeny have been shown to have a greater mitochondrial mass and higher rates of oxygen consumption compared with normal HSCs. There are increasing amounts of data in the literature showing a significant role for mitochondria in both AML pathogenesis and resistance to therapy. Mitochondria contain complexes that regulate protein levels by eliminating excess or damaged proteins. One of the 15 identified proteases for eliminating damaged proteins in the mitochondria is caseinolytic protease P (ClpP) [175]. ClpP maintains the integrity of OXPHOS, and its inhibition results in an increase of misfolded proteins in the respiratory chain, leading to respiratory dysfunction in AML cells [176]. However, hyperactivation of ClpP can also be toxic to cells. The activation of ClpP by ONC201 and ONC212 was shown to induce apoptosis in primary AML cells with little effect on normal HSCs [177]. Primary AML patients with higher ClpP expression were shown to be more sensitive to ClpP activators compared with samples that have lower-than-average expression levels. Activation of ClpP selectively degrades the respiratory chain similarly in normal HSCs; however, the greater sensitivity of AML cells reflects the enhanced reliance of AML cells on OXPHOS and lower spare reserve capacity in their respiratory chain [177].
Targeting different components of the mitochondria has been suggested as a strategy to overcome resistance in patients treated with ven/az. The caseinolytic peptidase B protein homolog (CLPB) protein, a mitochondrial AAA+ ATPase chaperone, was one of the genes shown to be upregulated in primary AML, and was further upregulated upon acquisition of Venetoclax resistance [178]. Cheng et al. showed that CLPB maintains the mitochondrial cristae structure through its interaction with the cristae-shaping protein, OPA1, and if lost, promotes apoptosis by inducing cristae remodeling and mitochondrial stress responses. This finding suggests that targeting mitochondrial architecture may provide a promising approach to circumvent Venetoclax resistance [178].
In a study by Hole et al., 65% of AML patients showed significantly elevated superoxide production compared with normal controls, which was shown to occur through the function of NOX family members [55]. The enhanced ROS formation promotes cell proliferation and migration and thereby contributes to leukemic cell transformation [179,180]. In normal cells, ROS-induced stress results in activation of stress-activated protein kinase (SPARK). p38MAPK is a SPARK that is activated by ROS, resulting in cell cycle arrest. The high level of ROS in primary AML blasts is associated with defective p38MAPK stress signaling [174]. This means that, in spite of high ROS production, the AML blast cells do not undergo cell cycle arrest. The elevated ROS levels have not been shown to be limited to particular AML subtypes [174]. Among the NOX family, mainly NOX2 expression in primary AML blasts has been shown to be correlated with superoxide production [174]. The generated superoxide by NOX is converted to H2O2 by superoxide dismutase. Primary AML cells constitutively generate H2O2, which promotes the proliferation of both AML blasts and cell lines [174], and therefore NOX2 may be essential for the viability and proliferation of AML cells [181]. However, a different mechanism for oncogenicity of NOX2 in AML was reported by Adane et al., who demonstrated that the NOX2 complex is strongly expressed in LSCs and its expression is important for LSC self-renewal [182]. The role of NOX2 at inducing self-renewal was shown to be through activation of FOXC1. Inhibition of NOX2 in the LSCs of an AML mouse model reduced the dynamic of mitochondrial and glycolytic metabolism, indicating that suppression of NOX2 could reduce the core metabolic pathways in AML cells and be a therapeutic option for eradicating AML LSCs [182].

8. Concluding Thoughts

AML is a heterogeneous disease that has a poor prognosis, especially in older individuals. Both intrinsic and extrinsic factors of leukemic cells and signals from the BM microenvironment play a role in disease pathogenesis and response to therapy. In recent years, many different enzymes, transcription factors, signaling pathways, and components of the microenvironment have been shown to contribute to LSC survival and drug resistance in AML, and thereby represent novel targets for therapy. As a result, several different targeted therapies have been developed for the treatment of AML. Although these types of medications improve the outcome of many AML patients, some still have an unfavorable response, meaning that we have much more to discover in order to cure this incredibly challenging disease. In the future, personalized medicine will be required to eradicate this disease, in which a patient is treated based on their individual mutation status and drug sensitivity. Eradication of AML will rely on the realization of new target inhibitors and the use of multiple drugs in personalized medicine approaches. Finally, the heterogeneity of the disease highlights the importance of personalized medicine and the need for new diagnostic methods.

Author Contributions

The authors confirm contribution to the paper as follows: Conceptualization, writing, and editing of the paper: J.S.K., S.M.B., M.P., P.B., A.A., M.K. and N.G.; graphic design and tables: S.M.B., M.P., M.K., A.A. and N.G.; revised the manuscript: A.M.E. and J.S.K.; supervision: A.M.E. and J.S.K.; project administration: A.M.E. and J.S.K.; S.M.B. and M.P. equally contributed to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Elsa U. Pardee Foundation (A.M.E.) and the National Cancer Institute of the National Institutes of Health under Award Number K22CA216008 (A.M.E.). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaushansky, K.; Zhan, H. The regulation of normal and neoplastic hematopoiesis is dependent on microenvironmental cells. Adv. Biol. Regul. 2018, 69, 11–15. [Google Scholar] [CrossRef] [PubMed]
  2. Korn, C.; Méndez-Ferrer, S. Myeloid malignancies and the microenvironment. Blood 2017, 129, 811–822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Shallis, R.M.; Wang, R.; Davidoff, A.; Ma, X.; Zeidan, A.M. Epidemiology of acute myeloid leukemia: Recent progress and enduring challenges. Blood Rev. 2019, 36, 70–87. [Google Scholar] [CrossRef] [PubMed]
  4. Gordon, J.E.; Wong, J.J.L.; Rasko, J.E. Micro RNA s in myeloid malignancies. Br. J. Haematol. 2013, 162, 162–176. [Google Scholar] [CrossRef]
  5. Potts, K.S.; Bowman, T.V. Modeling myeloid malignancies using zebrafish. Front. Oncol. 2017, 7, 297. [Google Scholar] [CrossRef]
  6. Thomas, D.; Majeti, R. Biology and relevance of human acute myeloid leukemia stem cells. Blood 2017, 129, 1577–1585. [Google Scholar] [CrossRef] [PubMed]
  7. Papayannidis, C.; Sartor, C.; Marconi, G.; Fontana, M.C.; Nanni, J.; Cristiano, G.; Parisi, S.; Paolini, S.; Curti, A. Acute Myeloid Leukemia Mutations: Therapeutic Implications. Int. J. Mol. Sci. 2019, 20, 2721. [Google Scholar] [CrossRef] [Green Version]
  8. Yilmaz, M.; Kadia, T.; Ravandi, F. Identifying effective drug combinations for patients with acute myeloid leukemia. Expert Rev. Anticancer Ther. 2020, 1–11. [Google Scholar] [CrossRef] [PubMed]
  9. National Cancer Institute. Cancer Stat Facts: Leukemia-Acute Myeloid Leukemia (AML). 2017. Available online: https://seer.cancer.gov/statfacts/html/amyl.html (accessed on 17 September 2021).
  10. Yi, M.; Li, A.; Zhou, L.; Chu, Q.; Song, Y.; Wu, K. The global burden and attributable risk factor analysis of acute myeloid leukemia in 195 countries and territories from 1990 to 2017: Estimates based on the global burden of disease study 2017. J. Hematol. Oncol. 2020, 13, 1–16. [Google Scholar] [CrossRef]
  11. Puumala, S.E.; Ross, J.A.; Aplenc, R.; Spector, L.G. Epidemiology of childhood acute myeloid leukemia. Pediatr. Blood Cancer 2013, 60, 728–733. [Google Scholar] [CrossRef] [Green Version]
  12. Chen, X.; Pan, J.; Wang, S.; Hong, S.; Hong, S.; He, S. The Epidemiological Trend of Acute Myeloid Leukemia in Childhood: A Population-Based Analysis. J. Cancer 2019, 10, 4824. [Google Scholar] [CrossRef]
  13. Medyouf, H.; Mossner, M.; Jann, J.-C.; Nolte, F.; Raffel, S.; Herrmann, C.; Lier, A.; Eisen, C.; Nowak, V.; Zens, B. Myelodysplastic cells in patients reprogram mesenchymal stromal cells to establish a transplantable stem cell niche disease unit. Cell Stem cell 2014, 14, 824–837. [Google Scholar] [CrossRef] [Green Version]
  14. Noone, A.; Howlader, N.; Krapcho, M.; Miller, D.; Brest, A.; Yu, M.; Ruhl, J.; Tatalovich, Z.; Mariotto, A.; Lewis, D. Surveillance, Epidemiology, and End Results (SEER) Program Cancer Statistics Review, 1975-2015; National Cancer Institute: Bethesda, MD, USA, 2018.
  15. Short, N.J.; Rytting, M.E.; Cortes, J.E. Acute myeloid leukaemia. Lancet 2018, 392, 593–606. [Google Scholar] [CrossRef]
  16. Magina, K.N.; Pregartner, G.; Zebisch, A.; Wölfler, A.; Neumeister, P.; Greinix, H.T.; Berghold, A.; Sill, H. Cytarabine dose in the consolidation treatment of AML: A systematic review and meta-analysis. Blood 2017, 130, 946–948. [Google Scholar] [CrossRef] [Green Version]
  17. Cree, I.A.; Charlton, P. Molecular chess? Hallmarks of anti-cancer drug resistance. BMC Cancer 2017, 17, 10. [Google Scholar] [CrossRef] [Green Version]
  18. Gabra, M.M.; Salmena, L. microRNAs and acute myeloid leukemia chemoresistance: A mechanistic overview. Front. Oncol. 2017, 7, 255. [Google Scholar] [CrossRef]
  19. van Gils, N.; Denkers, F.; Smit, L. Escape From Treatment; the Different Faces of Leukemic Stem Cells and Therapy Resistance in Acute Myeloid Leukemia. Front. Oncol. 2021, 11, 659253. [Google Scholar] [CrossRef] [PubMed]
  20. Saultz, J.N.; Garzon, R. Acute Myeloid Leukemia: A Concise Review. J. Clin. Med. 2016, 5, 33. [Google Scholar] [CrossRef] [Green Version]
  21. Narayanan, D.; Weinberg, O.K. How I investigate acute myeloid leukemia. Int. J. Lab. Hematol. 2020, 42, 3–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Khwaja, A.; Bjorkholm, M.; Gale, R.E.; Levine, R.L.; Jordan, C.T.; Ehninger, G.; Bloomfield, C.D.; Estey, E.; Burnett, A.; Cornelissen, J.J. Acute myeloid leukaemia. Nat. Rev. Dis. Primers 2016, 2, 1–22. [Google Scholar] [CrossRef] [PubMed]
  23. Matarraz, S.; Almeida, J.; Flores-Montero, J.; Lécrevisse, Q.; Guerri, V.; López, A.; Bárrena, S.; Van Der Velden, V.H.; Te Marvelde, J.G.; Van Dongen, J.J. Introduction to the diagnosis and classification of monocytic-lineage leukemias by flow cytometry. Cytometry B Clin. Cytom. 2017, 92, 218–227. [Google Scholar] [CrossRef] [Green Version]
  24. Demirer, S.; Özdemir, H.; Şencan, M.; Marakoḡlu, I. Gingival hyperplasia as an early diagnostic oral manifestation in acute monocytic leukemia: A case report. Eur. J. Dent. 2007, 1, 111. [Google Scholar] [CrossRef] [Green Version]
  25. Reikvam, H.; Hatfield, K.J.; Kittang, A.O.; Hovland, R.; Bruserud, Ø. Acute myeloid leukemia with the t (8; 21) translocation: Clinical consequences and biological implications. J. Biomed. Biotechnol. 2011, 2011, 104631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Bullinger, L.; Döhner, K.; Döhner, H. Genomics of acute myeloid leukemia diagnosis and pathways. J. Clin. Oncol. 2017, 35, 934–946. [Google Scholar] [CrossRef]
  27. Papaemmanuil, E.; Gerstung, M.; Bullinger, L.; Gaidzik, V.I.; Paschka, P.; Roberts, N.D.; Potter, N.E.; Heuser, M.; Thol, F.; Bolli, N. Genomic classification and prognosis in acute myeloid leukemia. N. Engl. J. Med. 2016, 374, 2209–2221. [Google Scholar] [CrossRef]
  28. Galera, P.K.; Jiang, C.; Braylan, R. Immunophenotyping of Acute Myeloid Leukemia. Methods Mol. Biol. 2019, 2032, 281–296. [Google Scholar] [PubMed]
  29. Döhner, H.; Estey, E.; Grimwade, D.; Amadori, S.; Appelbaum, F.R.; Büchner, T.; Dombret, H.; Ebert, B.L.; Fenaux, P.; Larson, R.A. Diagnosis and management of AML in adults: 2017 ELN recommendations from an international expert panel. Blood 2017, 129, 424–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Isidori, A.; Salvestrini, V.; Ciciarello, M.; Loscocco, F.; Visani, G.; Parisi, S.; Lecciso, M.; Ocadlikova, D.; Rossi, L.; Gabucci, E. The role of the immunosuppressive microenvironment in acute myeloid leukemia development and treatment. Expert Rev. Hematol. 2014, 7, 807–818. [Google Scholar] [CrossRef]
  31. Bose, P.; Vachhani, P.; Cortes, J.E. Treatment of relapsed/refractory acute myeloid leukemia. Curr. Treat. Options Oncol. 2017, 18, 17. [Google Scholar] [CrossRef]
  32. Kampen, K.R.; Ter Elst, A.; de Bont, E.S. Vascular endothelial growth factor signaling in acute myeloid leukemia. Cell. Mol. Life Sci. 2013, 70, 1307–1317. [Google Scholar] [CrossRef] [Green Version]
  33. American Cancer Society. Cancer Facts and Figures 2005; American Cancer Society: Atlanta, GA, USA, 2005. [Google Scholar]
  34. Zhang, J.; Gu, Y.; Chen, B. Mechanisms of drug resistance in acute myeloid leukemia. Onco Targets Ther. 2019, 12, 1937. [Google Scholar] [CrossRef] [Green Version]
  35. Briot, T.; Roger, E.; Thépot, S.; Lagarce, F. Advances in treatment formulations for acute myeloid leukemia. Drug Discov. Today 2018, 23, 1936–1949. [Google Scholar] [CrossRef] [Green Version]
  36. De Kouchkovsky, I.; Abdul-Hay, M. Acute myeloid leukemia: A comprehensive review and 2016 update. Blood Cancer J. 2016, 6, e441. [Google Scholar] [CrossRef]
  37. Acheampong, D.O.; Adokoh, C.K.; Asante, D.-B.; Asiamah, E.A.; Barnie, P.A.; Bonsu, D.O.; Kyei, F. Immunotherapy for acute myeloid leukemia (AML): A potent alternative therapy. Biomed. Pharmacother. 2018, 97, 225–232. [Google Scholar] [CrossRef]
  38. Forghieri, F.; Comoli, P.; Marasca, R.; Potenza, L.; Luppi, M. Minimal/Measurable Residual Disease Monitoring in NPM1-Mutated Acute Myeloid Leukemia: A Clinical Viewpoint and Perspectives. Int. J. Mol. Sci. 2018, 19, 3492. [Google Scholar] [CrossRef] [Green Version]
  39. Kassim, A.A.; Savani, B.N. Hematopoietic stem cell transplantation for acute myeloid leukemia: A review. Hematol. Oncol. Stem Cell Ther. 2017, 10, 245–251. [Google Scholar] [CrossRef] [PubMed]
  40. Craddock, C.; Raghavan, M. Which patients with acute myeloid leukemia in CR1 can be spared an allogeneic transplant? Curr. Opin. Hematol. 2019, 26, 58–64. [Google Scholar] [CrossRef] [PubMed]
  41. Kumar, L. Leukemia: Management of relapse after allogeneic bone marrow transplantation. J. Clin. Oncol. 1994, 12, 1710–1717. [Google Scholar] [CrossRef] [PubMed]
  42. Yee, G.; McGuire, T. Allogeneic bone marrow transplantation in the treatment of hematologic diseases. Clin. Pharm. 1985, 4, 149–160. [Google Scholar]
  43. Cerrano, M.; Itzykson, R. New treatment options for acute myeloid leukemia in 2019. Curr. Oncol. Rep. 2019, 21, 16. [Google Scholar] [CrossRef]
  44. Ladikou, E.E.; Sivaloganathan, H.; Pepper, A.; Chevassut, T. Acute Myeloid Leukaemia in Its Niche: The Bone Marrow Microenvironment in Acute Myeloid Leukaemia. Curr. Oncol. Rep. 2020, 22, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Xiao, Y.; Geng, Z.; Deng, T.; Wang, D.; Jiang, L. Tumor Necrosis Factor Receptor Type 1-Associated Death Domain Protein Is a Potential Prognostic Biomarker in Acute Myeloid Leukemia. Am. J. Med. Sci. 2019, 357, 111–115. [Google Scholar] [CrossRef] [PubMed]
  46. Tiong, I.S.; Wei, A.H. New drugs creating new challenges in acute myeloid leukemia. Genes Chromosomes Cancer 2019, 58, 903–914. [Google Scholar] [CrossRef] [Green Version]
  47. Elshoury, A.; Przespolewski, A.; Baron, J.; Wang, E.S. Advancing treatment of acute myeloid leukemia: The future of FLT3 inhibitors. Expert Rev. Anticancer Ther. 2019, 19, 273–286. [Google Scholar] [CrossRef]
  48. Kopmar, N.E.; Estey, E.H. New Drug Approvals in Acute Myeloid Leukemia: An Unprecedented Paradigm Shift. Clin. Adv. Hematol. Oncol. 2019, 17, 569–575. [Google Scholar]
  49. Döhner, H.; Estey, E.H.; Amadori, S.; Appelbaum, F.R.; Büchner, T.; Burnett, A.K.; Dombret, H.; Fenaux, P.; Grimwade, D.; Larson, R.A. Diagnosis and management of acute myeloid leukemia in adults: Recommendations from an international expert panel, on behalf of the European LeukemiaNet. Blood 2010, 115, 453–474. [Google Scholar] [CrossRef]
  50. Lovly, C.M.; Shaw, A.T. Molecular pathways: Resistance to kinase inhibitors and implications for therapeutic strategies. Clin. Cancer Res. 2014, 20, 2249–2256. [Google Scholar] [CrossRef] [Green Version]
  51. Schofield, R. The relationship between the spleen colony-forming cell and the haemopoietic stem cell. Blood Cells 1978, 4, 7–25. [Google Scholar]
  52. Tabe, Y.; Konopleva, M. Role of microenvironment in resistance to therapy in AML. Curr. Hematol. Malig. Rep. 2015, 10, 96–103. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, A.; Zhong, H. Roles of the bone marrow niche in hematopoiesis, leukemogenesis, and chemotherapy resistance in acute myeloid leukemia. Hematology 2018, 23, 729–739. [Google Scholar] [CrossRef] [Green Version]
  54. Chopra, M.; Bohlander, S.K. The cell of origin and the leukemia stem cell in acute myeloid leukemia. Genes Chromosomes Cancer 2019, 58, 850–858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Testa, U.; Labbaye, C.; Castelli, G.; Pelosi, E. Oxidative stress and hypoxia in normal and leukemic stem cells. Exp. Hematol. 2016, 44, 540–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Corces-Zimmerman, M.R.; Hong, W.-J.; Weissman, I.L.; Medeiros, B.C.; Majeti, R. Preleukemic mutations in human acute myeloid leukemia affect epigenetic regulators and persist in remission. Proc. Natl. Acad. Sci. USA. 2014, 111, 2548–2553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Chan, S.M.; Majeti, R. Role of DNMT3A, TET2, and IDH1/2 mutations in pre-leukemic stem cells in acute myeloid leukemia. Int. J. Hematol. 2013, 98, 648–657. [Google Scholar] [CrossRef]
  58. Lapidot, T.; Sirard, C.; Vormoor, J.; Murdoch, B.; Hoang, T.; Caceres-Cortes, J.; Minden, M.; Paterson, B.; Caligiuri, M.A.; Dick, J.E. A cell initiating human acute myeloid leukaemia after transplantation into SCID mice. Nature 1994, 367, 645–648. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, X.; Huang, S.; Chen, J.L. Understanding of leukemic stem cells and their clinical implications. Mol. Cancer 2017, 16, 2. [Google Scholar] [CrossRef] [Green Version]
  60. Hanekamp, D.; Cloos, J.; Schuurhuis, G.J. Leukemic stem cells: Identification and clinical application. Int. J. Hematol. 2017, 105, 549–557. [Google Scholar] [CrossRef]
  61. Camacho, V.; McClearn, V.; Patel, S.; Welner, R.S. Regulation of normal and leukemic stem cells through cytokine signaling and the microenvironment. Int. J. Hematol. 2017, 105, 566–577. [Google Scholar] [CrossRef]
  62. Behrmann, L.; Wellbrock, J.; Fiedler, W. The bone marrow stromal niche: A therapeutic target of hematological myeloid malignancies. Expert Opin. Ther. Targets 2020, 24, 451–462. [Google Scholar] [CrossRef]
  63. Lane, S.W.; Wang, Y.J.; Celso, C.L.; Ragu, C.; Bullinger, L.; Sykes, S.M.; Ferraro, F.; Shterental, S.; Lin, C.P.; Gilliland, D.G. Differential niche and Wnt requirements during acute myeloid leukemia progression. Blood 2011, 118, 2849–2856. [Google Scholar] [CrossRef] [Green Version]
  64. Kumar, B.; Garcia, M.; Weng, L.; Jung, X.; Murakami, J.; Hu, X.; McDonald, T.; Lin, A.; Kumar, A.; DiGiusto, D. Acute myeloid leukemia transforms the bone marrow niche into a leukemia-permissive microenvironment through exosome secretion. Leukemia 2018, 32, 575–587. [Google Scholar] [CrossRef]
  65. Schepers, K.; Pietras, E.M.; Reynaud, D.; Flach, J.; Binnewies, M.; Garg, T.; Wagers, A.J.; Hsiao, E.C.; Passegué, E. Myeloproliferative neoplasia remodels the endosteal bone marrow niche into a self-reinforcing leukemic niche. Cell Stem Cell 2013, 13, 285–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Glait-Santar, C.; Desmond, R.; Feng, X.; Bat, T.; Chen, J.; Heuston, E.; Mizukawa, B.; Mulloy, J.C.; Bodine, D.M.; Larochelle, A. Functional niche competition between normal hematopoietic stem and progenitor cells and myeloid leukemia cells. Stem Cells 2015, 33, 3635–3642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Pitt, L.A.; Tikhonova, A.N.; Hu, H.; Trimarchi, T.; King, B.; Gong, Y.; Sanchez-Martin, M.; Tsirigos, A.; Littman, D.R.; Ferrando, A.A. CXCL12-producing vascular endothelial niches control acute T cell leukemia maintenance. Cancer Cell 2015, 27, 755–768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Kim, J.-A.; Shim, J.-S.; Lee, G.-Y.; Yim, H.W.; Kim, T.-M.; Kim, M.; Leem, S.-H.; Lee, J.-W.; Min, C.-K.; Oh, I.-H. Microenvironmental remodeling as a parameter and prognostic factor of heterogeneous leukemogenesis in acute myelogenous leukemia. Cancer Res. 2015, 75, 2222–2231. [Google Scholar] [CrossRef] [Green Version]
  69. Kobayashi, S.S.; Takei, H. Transcription factor-based therapies for acute myeloid leukemia. Rinsho Ketsueki 2018, 59, 922–931. [Google Scholar]
  70. Karathedath, S.; Rajamani, B.M.; Musheer Aalam, S.M.; Abraham, A.; Varatharajan, S.; Krishnamurthy, P.; Mathews, V.; Velayudhan, S.R.; Balasubramanian, P. Role of NF-E2 related factor 2 (Nrf2) on chemotherapy resistance in acute myeloid leukemia (AML) and the effect of pharmacological inhibition of Nrf2. PLoS ONE 2017, 12, e0177227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Takei, H.; Kobayashi, S.S. Targeting transcription factors in acute myeloid leukemia. Int. J. Hematol. 2019, 109, 28–34. [Google Scholar] [CrossRef] [Green Version]
  72. Seipel, K.; Marques, M.T.; Bozzini, M.-A.; Meinken, C.; Mueller, B.U.; Pabst, T. Inactivation of the p53–KLF4–CEBPA Axis in Acute Myeloid Leukemia. Clin. Cancer Res. 2016, 22, 746–756. [Google Scholar] [CrossRef] [Green Version]
  73. Wahlin, A. Accumulating evidence for a role of p53 in multiple drug resistant Acute Myeloid Leukemia. Leuk. Lymphoma 2008, 49, 383–384. [Google Scholar] [CrossRef]
  74. Pan, X.-N.; Chen, J.-J.; Wang, L.-X.; Xiao, R.-Z.; Liu, L.-L.; Fang, Z.-G.; Liu, Q.; Long, Z.-J.; Lin, D.-J. Inhibition of c-Myc overcomes cytotoxic drug resistance in acute myeloid leukemia cells by promoting differentiation. PLoS ONE 2014, 9, e105381. [Google Scholar] [CrossRef]
  75. Gleixner, K.V.; Schneeweiss, M.; Eisenwort, G.; Berger, D.; Herrmann, H.; Blatt, K.; Greiner, G.; Byrgazov, K.; Hoermann, G.; Konopleva, M. Combined targeting of STAT3 and STAT5: A novel approach to overcome drug resistance in chronic myeloid leukemia. Haematologica 2017, 102, 1519–1529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Mesbahi, Y.; Zekri, A.; Ghaffari, S.H.; Tabatabaie, P.S.; Ahmadian, S.; Ghavamzadeh, A. Blockade of JAK2/STAT3 intensifies the anti-tumor activity of arsenic trioxide in acute myeloid leukemia cells: Novel synergistic mechanism via the mediation of reactive oxygen species. Eur. J. Pharmacol. 2018, 834, 65–76. [Google Scholar] [CrossRef] [PubMed]
  77. Morris, V.A.; Cummings, C.L.; Korb, B.; Boaglio, S.; Oehler, V.G. Deregulated KLF4 expression in myeloid leukemias alters cell proliferation and differentiation through microRNA and gene targets. Mol. Cell. Biol. 2016, 36, 559–573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Safa, M.; Mousavizadeh, K.; Noori, S.; Pourfathollah, A.; Zand, H. cAMP protects acute promyelocytic leukemia cells from arsenic trioxide-induced caspase-3 activation and apoptosis. Eur. J. Pharmacol. 2014, 736, 115–123. [Google Scholar] [CrossRef] [PubMed]
  79. Shankar, D.B.; Cheng, J.C.; Sakamoto, K.M. Role of cyclic AMP response element binding protein in human leukemias. Cancer 2005, 104, 1819–1824. [Google Scholar] [CrossRef]
  80. Mitton, B.; Chae, H.-D.; Hsu, K.; Dutta, R.; Aldana-Masangkay, G.; Ferrari, R.; Davis, K.; Tiu, B.C.; Kaul, A.; Lacayo, N. Small molecule inhibition of cAMP response element binding protein in human acute myeloid leukemia cells. Leukemia 2016, 30, 2302–2311. [Google Scholar] [CrossRef]
  81. Mueller, B.U.; Pabst, T.; Osato, M.; Asou, N.; Johansen, L.M.; Minden, M.D.; Behre, G.; Hiddemann, W.; Ito, Y.; Tenen, D.G. Heterozygous PU. 1 mutations are associated with acute myeloid leukemia. Blood 2002, 100, 998–1007. [Google Scholar] [CrossRef] [Green Version]
  82. Goyama, S.; Schibler, J.; Cunningham, L.; Zhang, Y.; Rao, Y.; Nishimoto, N.; Nakagawa, M.; Olsson, A.; Wunderlich, M.; Link, K.A.; et al. Transcription factor RUNX1 promotes survival of acute myeloid leukemia cells. J. Clin. Investig. 2013, 123, 3876–3888. [Google Scholar] [CrossRef] [Green Version]
  83. Darwish, N.H.; Sudha, T.; Godugu, K.; Bharali, D.J.; Elbaz, O.; El-ghaffar, H.A.A.; Azmy, E.; Anber, N.; Mousa, S.A. Novel targeted nano-parthenolide molecule against NF-kB in Acute Myeloid Leukemia. Molecules 2019, 24, 2103. [Google Scholar] [CrossRef] [Green Version]
  84. Zhou, J.; Ching, Y.Q.; Chng, W.J. Aberrant nuclear factor-kappa B activity in acute myeloid leukemia: From molecular pathogenesis to therapeutic target. Oncotarget 2015, 6, 5490–5500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Kagoya, Y.; Yoshimi, A.; Kataoka, K.; Nakagawa, M.; Kumano, K.; Arai, S.; Kobayashi, H.; Saito, T.; Iwakura, Y.; Kurokawa, M. Positive feedback between NF-κB and TNF-α promotes leukemia-initiating cell capacity. J. Clin. Investig. 2014, 124, 528–542. [Google Scholar] [CrossRef]
  86. Asada, N.; Takeishi, S.; Frenette, P.S. Complexity of bone marrow hematopoietic stem cell niche. Int. J. Hematol. 2017, 106, 45–54. [Google Scholar] [CrossRef] [Green Version]
  87. Crane, G.M.; Jeffery, E.; Morrison, S.J. Adult haematopoietic stem cell niches. Nat. Rev. Immunol. 2017, 17, 573. [Google Scholar] [CrossRef]
  88. Ghiaur, G.; Wroblewski, M.; Loges, S. Acute myelogenous leukemia and its microenvironment: A molecular conversation. Semin. Hematol. 2015, 52, 200–206. [Google Scholar]
  89. Shafat, M.S.; Gnaneswaran, B.; Bowles, K.M.; Rushworth, S.A. The bone marrow microenvironment–Home of the leukemic blasts. Blood Rev. 2017, 31, 277–286. [Google Scholar] [CrossRef] [Green Version]
  90. Bakker, S.T.; Passegué, E. Resilient and resourceful: Genome maintenance strategies in hematopoietic stem cells. Exp. Hematol. 2013, 41, 915–923. [Google Scholar] [CrossRef] [Green Version]
  91. Ostanin, A.; Petrovskii, Y.L.; Shevela, E.Y.; Chernykh, E. Multiplex analysis of cytokines, chemokines, growth factors, MMP-9 and TIMP-1 produced by human bone marrow, adipose tissue, and placental mesenchymal stromal cells. Bull. Exp. Biol. Med. 2011, 151, 133–141. [Google Scholar] [CrossRef]
  92. Kondo, M.; Wagers, A.J.; Manz, M.G.; Prohaska, S.S.; Scherer, D.C.; Beilhack, G.F.; Shizuru, J.A.; Weissman, I.L. Biology of hematopoietic stem cells and progenitors: Implications for clinical application. Annu. Rev. Immunol. 2003, 21, 759–806. [Google Scholar] [CrossRef] [PubMed]
  93. Schepers, K.; Campbell, T.B.; Passegué, E. Normal and leukemic stem cell niches: Insights and therapeutic opportunities. Cell Stem Cell 2015, 16, 254–267. [Google Scholar] [CrossRef] [Green Version]
  94. Riether, C.; Schürch, C.; Ochsenbein, A. Regulation of hematopoietic and leukemic stem cells by the immune system. Cell Death Differ. 2015, 22, 187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Behrmann, L.; Wellbrock, J.; Fiedler, W. Acute myeloid leukemia and the bone marrow niche—Take a closer look. Front. Oncol. 2018, 8, 444. [Google Scholar] [CrossRef] [Green Version]
  96. Yu, V.; Scadden, D. Hematopoietic stem cell and its bone marrow niche. Curr. Top. Dev. Biol. 2016, 118, 21–44. [Google Scholar]
  97. Le, P.M.; Andreeff, M.; Battula, V.L. Osteogenic niche in the regulation of normal hematopoiesis and leukemogenesis. Haematologica 2018, 103, 1945–1955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Mangialardi, G.; Cordaro, A.; Madeddu, P. The bone marrow pericyte: An orchestrator of vascular niche. Regen. Med. 2016, 11, 883–895. [Google Scholar] [CrossRef] [Green Version]
  99. Tabe, Y.; Konopleva, M. Leukemia stem cells microenvironment. In Stem Cell Microenvironments and Beyond, Springer: 2017; pp 19-32.
  100. Yilmaz, Ö.H.; Valdez, R.; Theisen, B.K.; Guo, W.; Ferguson, D.O.; Wu, H.; Morrison, S.J. Pten dependence distinguishes haematopoietic stem cells from leukaemia-initiating cells. Nature 2006, 441, 475–482. [Google Scholar] [CrossRef] [Green Version]
  101. Sansone, P.; Bromberg, J. Targeting the interleukin-6/Jak/stat pathway in human malignancies. J. Clin. Oncol. 2012, 30, 1005–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Susek, K.H.; Karvouni, M.; Alici, E.; Lundqvist, A. The role of CXC chemokine receptors 1–4 on immune cells in the tumor microenvironment. Front. Immunol. 2018, 9, 2159. [Google Scholar] [CrossRef]
  103. Cho, B.-S.; Kim, H.-J.; Konopleva, M. Targeting the CXCL12/CXCR4 axis in acute myeloid leukemia: From bench to bedside. Korean J. Intern. Med. 2017, 32, 248. [Google Scholar] [CrossRef]
  104. Rashidi, A.; Uy, G.L. Targeting the microenvironment in acute myeloid leukemia. Curr. Hematol. Malig. Rep. 2015, 10, 126–131. [Google Scholar] [CrossRef] [PubMed]
  105. Gruszka, A.M.; Valli, D.; Restelli, C.; Alcalay, M. Adhesion deregulation in acute myeloid leukaemia. Cells 2019, 8, 66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Schlesinger, M.; Bendas, G. Contribution of very late antigen-4 (VLA-4) integrin to cancer progression and metastasis. Cancer Metastasis Rev. 2015, 34, 575–591. [Google Scholar] [CrossRef] [PubMed]
  107. Dong-Feng, Z.; Ting, L.; Yong, Z.; Cheng, C.; Xi, Z.; Pei-Yan, K. The TPO/c-MPL pathway in the bone marrow may protect leukemia cells from chemotherapy in AML patients. Pathol. Oncol. Res. 2014, 20, 309–317. [Google Scholar] [CrossRef] [PubMed]
  108. Yang, J.G.; Wang, L.L.; Ma, D.C. Effects of vascular endothelial growth factors and their receptors on megakaryocytes and platelets and related diseases. Br. J. Haematol. 2018, 180, 321–334. [Google Scholar] [CrossRef] [PubMed]
  109. Baaten, B.J.; Li, C.-R.; Deiro, M.F.; Lin, M.M.; Linton, P.J.; Bradley, L.M. CD44 regulates survival and memory development in Th1 cells. Immunity 2010, 32, 104–115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Binder, S.; Luciano, M.; Horejs-Hoeck, J. The cytokine network in acute myeloid leukemia (AML): A focus on pro-and anti-inflammatory mediators. Cytokine Growth Factor Rev. 2018, 43, 8–15. [Google Scholar] [CrossRef] [PubMed]
  111. Sacchetti, B.; Funari, A.; Remoli, C.; Giannicola, G.; Kogler, G.; Liedtke, S.; Cossu, G.; Serafini, M.; Sampaolesi, M.; Tagliafico, E.; et al. No Identical “Mesenchymal Stem Cells” at Different Times and Sites: Human Committed Progenitors of Distinct Origin and Differentiation Potential Are Incorporated as Adventitial Cells in Microvessels. Stem Cell Rep. 2016, 6, 897–913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Boraschi, D.; Tagliabue, A. The interleukin-1 receptor family. Sem. Immunol. 2013, 25, 394–407. [Google Scholar]
  113. Shapouri-Moghaddam, A.; Mohammadian, S.; Vazini, H.; Taghadosi, M.; Esmaeili, S.A.; Mardani, F.; Seifi, B.; Mohammadi, A.; Afshari, J.T.; Sahebkar, A. Macrophage plasticity, polarization, and function in health and disease. J. Cell. Physiol. 2018, 233, 6425–6440. [Google Scholar] [CrossRef] [PubMed]
  114. Yazdi, A.S.; Ghoreschi, K. The interleukin-1 family. In Regulation of Cytokine Gene Expression in Immunity and Diseases; Springer: Berlin, Germany, 2016; pp. 21–29. [Google Scholar]
  115. Estrov, Z.; Kurzrock, R.; Talpaz, M. Role of Interleukin-1 Inhibitory Molecules in Therapy of Acute and Chronic Myelogenous Leukemia. Leuk. Lymphoma 1993, 10, 407–418. [Google Scholar] [CrossRef] [PubMed]
  116. Arranz, L.; del Mar Arriero, M.; Villatoro, A. Interleukin-1β as emerging therapeutic target in hematological malignancies and potentially in their complications. Blood Rev. 2017, 31, 306–317. [Google Scholar] [CrossRef] [PubMed]
  117. Mehta, A.K.; Gracias, D.T.; Croft, M. TNF activity and T cells. Cytokine 2018, 101, 14–18. [Google Scholar] [CrossRef] [PubMed]
  118. Maleknia, M.; Valizadeh, A.; Pezeshki, S.; Saki, N. Immunomodulation in leukemia: Cellular aspects of anti-leukemic properties. Clin. Transl. Oncol. 2020, 22, 1–10. [Google Scholar] [CrossRef]
  119. Medler, J.; Wajant, H. Tumor necrosis factor receptor-2 (TNFR2): An overview of an emerging drug target. Expert Opin. Ther. Targets 2019, 23, 295–307. [Google Scholar] [CrossRef] [PubMed]
  120. Zhou, X.; Li, Z.; Zhou, J. Tumor necrosis factor α in the onset and progression of leukemia. Exp. Hematol. 2017, 45, 17–26. [Google Scholar] [CrossRef] [PubMed]
  121. Londino, J.D.; Gulick, D.L.; Lear, T.B.; Suber, T.L.; Weathington, N.M.; Masa, L.S.; Chen, B.B.; Mallampalli, R.K. Post-translational modification of the interferon-gamma receptor alters its stability and signaling. Biochem. J. 2017, 474, 3543–3557. [Google Scholar] [CrossRef] [Green Version]
  122. Kursunel, M.A.; Esendagli, G. The untold story of IFN-γ in cancer biology. Cytokine Growth Factor Rev. 2016, 31, 73–81. [Google Scholar] [CrossRef]
  123. De Weerd, N.A.; Nguyen, T. The interferons and their receptors—distribution and regulation. Immunol. Cell Biol. 2012, 90, 483–491. [Google Scholar] [CrossRef]
  124. Nursal, A.F.; Pehlivan, M.; Sahin, H.H.; Pehlivan, S. The Associations of IL-6, IFN-c, TNF-a, IL-10, and TGF-b1 Functional Variants with Acute Myeloid Leukemia in Turkish Patients. Genet. Test. Mol. Biomark. 2016, 20, 544–551. [Google Scholar] [CrossRef]
  125. Cook, R.S.; Jacobsen, K.M.; Wofford, A.M.; DeRyckere, D.; Stanford, J.; Prieto, A.L.; Redente, E.; Sandahl, M.; Hunter, D.M.; Strunk, K.E. MerTK inhibition in tumor leukocytes decreases tumor growth and metastasis. J. Clin. Investig. 2013, 123, 3231–3242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Wu, G.; Ma, Z.; Cheng, Y.; Hu, W.; Deng, C.; Jiang, S.; Li, T.; Chen, F.; Yang, Y. Targeting Gas6/TAM in cancer cells and tumor microenvironment. Mol. Cancer 2018, 17, 20. [Google Scholar] [CrossRef] [Green Version]
  127. Lo, W.-J.; Chang, W.-S.; Hsu, H.-F.; Ji, H.-X.; Hsiao, C.-L.; Tsai, C.-W.; Yeh, S.-P.; Chen, C.-M.; Bau, D.-T. Significant association of interleukin-10 polymorphisms with childhood leukemia susceptibility in Taiwan. In Vivo 2016, 30, 265–269. [Google Scholar]
  128. Carson, W.E.; Lindemann, M.J.; Baiocchi, R.; Linett, M.; Tan, J.C.; Chou, C.-C.; Narula, S.; Caligiuri, M. The functional characterization of interleukin-10 receptor expression on human natural killer cells. Blood 1995, 85, 3577–3585. [Google Scholar] [CrossRef] [Green Version]
  129. Bruserud, Ø.; Tore, B.G.; Brustugun, O.T.; Bassøe, C.; Nesthus, I.; Espen, P.A.; Bühring, H.; Pawelec, G. Effects of interleukin 10 on blast cells derived from patients with acute myelogenous leukemia. Leukemia 1995, 9, 1910–1920. [Google Scholar]
  130. Wu, Y.; Su, M.; Zhang, S.; Cheng, Y.; Liao, X.Y.; Lin, B.Y.; Chen, Y.Z. Abnormal expression of TGF-beta type II receptor isoforms contributes to acute myeloid leukemia. Oncotarget 2017, 8, 10037–10049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Arend, W.P. Interleukin-1 Receptor Antagonist. Adv. Immunol. 1993, 54, 167–227. [Google Scholar]
  132. Arend, W.P.; Guthridge, C.J. Biological role of interleukin 1 receptor antagonist isoforms. Ann. Rheum. Dis. 2000, 59, i60–i64. [Google Scholar] [CrossRef] [PubMed]
  133. Zhang, J.; Zhang, Y.; Li, C.; Zhang, X.; Xiong, H.; Deng, H. The Role of IL-35 in Regulating Tumor Immunity. Adv. Mod. Oncol. Res. 2018, 4, 8–16. [Google Scholar]
  134. Ok, C.Y.; Young, K.H. Checkpoint inhibitors in hematological malignancies. J. Hematol. Oncol. 2017, 10, 103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Vandsemb, E.N.; Kim, T.K.; Zeidan, A.M. Will deeper characterization of the landscape of immune checkpoint molecules in acute myeloid leukemia bone marrow lead to improved therapeutic targeting? Cancer 2019, 125, 1410. [Google Scholar] [CrossRef] [PubMed]
  136. Kikushige, Y.; Miyamoto, T.; Yuda, J.; Jabbarzadeh-Tabrizi, S.; Shima, T.; Takayanagi, S.-i.; Niiro, H.; Yurino, A.; Miyawaki, K.; Takenaka, K. A TIM-3/Gal-9 autocrine stimulatory loop drives self-renewal of human myeloid leukemia stem cells and leukemic progression. Cell Stem Cell 2015, 17, 341–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Giannopoulos, K. Targeting immune signaling checkpoints in acute myeloid leukemia. J. Clin. Med. 2019, 8, 236. [Google Scholar] [CrossRef] [Green Version]
  138. Abdul-Aziz, A.M.; Sun, Y.; Hellmich, C.; Marlein, C.R.; Mistry, J.; Forde, E.; Piddock, R.E.; Shafat, M.S.; Morfakis, A.; Mehta, T.; et al. Acute myeloid leukemia induces protumoral p16INK4a-driven senescence in the bone marrow microenvironment. Blood 2019, 133, 446–456. [Google Scholar] [CrossRef] [Green Version]
  139. Dias, S.; Choy, M.; Alitalo, K.; Rafii, S. Vascular endothelial growth factor (VEGF)–C signaling through FLT-4 (VEGFR-3) mediates leukemic cell proliferation, survival, and resistance to chemotherapy. Blood 2002, 99, 2179–2184. [Google Scholar] [CrossRef]
  140. Yang, X.; Sexauer, A.; Levis, M. Bone marrow stroma-mediated resistance to FLT 3 inhibitors in FLT 3-ITD AML is mediated by persistent activation of extracellular regulated kinase. Br. J. Haematol. 2014, 164, 61–72. [Google Scholar] [CrossRef] [PubMed]
  141. Carter, B.Z.; Mak, P.Y.; Wang, X.; Tao, W.; Ruvolo, V.; Mak, D.; Mu, H.; Burks, J.K.; Andreeff, M. An ARC-Regulated IL1β/Cox-2/PGE2/β-Catenin/ARC Circuit Controls Leukemia–Microenvironment Interactions and Confers Drug Resistance in AML. Cancer Res. 2019, 79, 1165–1177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Sha, C.; Jia, G.; Jingjing, Z.; Yapeng, H.; Zhi, L.; Guanghui, X. miR-486 is involved in the pathogenesis of acute myeloid leukemia by regulating JAK-STAT signaling. Naunyn Schmiedebergs Arch. Pharmacol. 2020, 394, 177–187. [Google Scholar] [CrossRef]
  143. Venugopal, S.; Bar-Natan, M.; Mascarenhas, J.O. JAKs to STATs: A tantalizing therapeutic target in acute myeloid leukemia. Blood Rev. 2020, 40, 100634. [Google Scholar] [CrossRef] [PubMed]
  144. Takam Kamga, P.; Collo, G.D.; Resci, F.; Bazzoni, R.; Mercuri, A.; Quaglia, F.M.; Tanasi, I.; Delfino, P.; Visco, C.; Bonifacio, M. Notch Signaling Molecules as Prognostic Biomarkers for Acute Myeloid Leukemia. Cancers 2019, 11, 1958. [Google Scholar] [CrossRef] [Green Version]
  145. Xu, X.; Zhao, Y.; Xu, M.; Dai, Q.; Meng, W.; Yang, J.; Qin, R. Activation of Notch signal pathway is associated with a poorer prognosis in acute myeloid leukemia. Med. Oncol. 2011, 28, 483–489. [Google Scholar] [CrossRef]
  146. Terao, T.; Minami, Y. Targeting hedgehog (Hh) pathway for the acute myeloid leukemia treatment. Cells 2019, 8, 312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. McCubrey, J.A.; Steelman, L.S.; Chappell, W.H.; Abrams, S.L.; Wong, E.W.; Chang, F.; Lehmann, B.; Terrian, D.M.; Milella, M.; Tafuri, A. Roles of the Raf/MEK/ERK pathway in cell growth, malignant transformation and drug resistance. Biochim. Biophys. Acta 2007, 1773, 1263–1284. [Google Scholar] [CrossRef] [Green Version]
  148. Knight, T.; Irving, J.A. Ras/Raf/MEK/ERK Pathway Activation in Childhood Acute Lymphoblastic Leukemia and Its Therapeutic Targeting. Front. Oncol. 2014, 4, 160. [Google Scholar] [CrossRef]
  149. Nepstad, I.; Hatfield, K.J.; Grønningsæter, I.S.; Reikvam, H. The PI3K-Akt-mTOR Signaling Pathway in Human Acute Myeloid Leukemia (AML) Cells. Int. J. Mol. Sci. 2020, 21, 2907. [Google Scholar] [CrossRef] [Green Version]
  150. Gruszka, A.M.; Valli, D.; Alcalay, M. Wnt signalling in acute myeloid leukaemia. Cells 2019, 8, 1403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Ashihara, E.; Takada, T.; Maekawa, T. Targeting the canonical Wnt/β-catenin pathway in hematological malignancies. Cancer Sci. 2015, 106, 665–671. [Google Scholar] [CrossRef] [Green Version]
  152. Miraki-Moud, F.; Anjos-Afonso, F.; Hodby, K.A.; Griessinger, E.; Rosignoli, G.; Lillington, D.; Jia, L.; Davies, J.K.; Cavenagh, J.; Smith, M.; et al. Acute myeloid leukemia does not deplete normal hematopoietic stem cells but induces cytopenias by impeding their differentiation. Proc. Natl. Acad. Sci. USA 2013, 110, 13576–13581. [Google Scholar] [CrossRef] [Green Version]
  153. Waclawiczek, A.; Hamilton, A.; Rouault-Pierre, K.; Abarrategi, A.; Albornoz, M.G.; Miraki-Moud, F.; Bah, N.; Gribben, J.; Fitzgibbon, J.; Taussig, D.; et al. Mesenchymal niche remodeling impairs hematopoiesis via stanniocalcin 1 in acute myeloid leukemia. J. Clin. Investig. 2020, 130, 3038–3050. [Google Scholar] [CrossRef]
  154. Pelullo, M.; Zema, S.; Nardozza, F.; Checquolo, S.; Screpanti, I.; Bellavia, D. Wnt, Notch, and TGF-β pathways impinge on Hedgehog signaling complexity: An open window on cancer. Front. Genet. 2019, 10. [Google Scholar] [CrossRef] [Green Version]
  155. Konopleva, M.Y.; Jordan, C.T. Leukemia stem cells and microenvironment: Biology and therapeutic targeting. J. Clin. Oncol. 2011, 29, 591–599. [Google Scholar] [CrossRef] [Green Version]
  156. Bakker, E.; Qattan, M.; Mutti, L.; Demonacos, C.; Krstic-Demonacos, M. The role of microenvironment and immunity in drug response in leukemia. Biochim. Biophys. Acta 2016, 1863, 414–426. [Google Scholar] [CrossRef]
  157. Bertacchini, J.; Heidari, N.; Mediani, L.; Capitani, S.; Shahjahani, M.; Ahmadzadeh, A.; Saki, N. Targeting PI3K/AKT/mTOR network for treatment of leukemia. Cell Mol. Life Sci. 2015, 72, 2337–2347. [Google Scholar] [CrossRef]
  158. Sakamoto, K.M.; Grant, S.; Saleiro, D.; Crispino, J.D.; Hijiya, N.; Giles, F.; Platanias, L.; Eklund, E.A. Targeting novel signaling pathways for resistant acute myeloid leukemia. Mol. Genet. Metab. 2015, 114, 397–402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Arrigoni, E.; Del Re, M.; Galimberti, S.; Restante, G.; Rofi, E.; Crucitta, S.; Baratè, C.; Petrini, M.; Danesi, R.; Di Paolo, A. Concise review: Chronic myeloid leukemia: Stem cell niche and response to pharmacologic treatment. Stem Cells Transl. Med. 2018, 7, 305–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Forte, D.; Garcia-Fernandez, M.; Sanchez-Aguilera, A.; Stavropoulou, V.; Fielding, C.; Martin-Perez, D.; Lopez, J.A.; Costa, A.S.H.; Tronci, L.; Nikitopoulou, E.; et al. Bone Marrow Mesenchymal Stem Cells Support Acute Myeloid Leukemia Bioenergetics and Enhance Antioxidant Defense and Escape from Chemotherapy. Cell Metab. 2020, 32, 829–843.e9. [Google Scholar] [CrossRef] [PubMed]
  161. Arranz, L.; Sanchez-Aguilera, A.; Martin-Perez, D.; Isern, J.; Langa, X.; Tzankov, A.; Lundberg, P.; Muntion, S.; Tzeng, Y.S.; Lai, D.M.; et al. Neuropathy of haematopoietic stem cell niche is essential for myeloproliferative neoplasms. Nature 2014, 512, 78–81. [Google Scholar] [CrossRef] [PubMed]
  162. Hanoun, M.; Zhang, D.; Mizoguchi, T.; Pinho, S.; Pierce, H.; Kunisaki, Y.; Lacombe, J.; Armstrong, S.A.; Duhrsen, U.; Frenette, P.S. Acute myelogenous leukemia-induced sympathetic neuropathy promotes malignancy in an altered hematopoietic stem cell niche. Cell Stem Cell 2014, 15, 365–375. [Google Scholar] [CrossRef] [Green Version]
  163. Kouzi, F.; Zibara, K.; Bourgeais, J.; Picou, F.; Gallay, N.; Brossaud, J.; Dakik, H.; Roux, B.; Hamard, S.; Le Nail, L.-R. Disruption of gap junctions attenuates acute myeloid leukemia chemoresistance induced by bone marrow mesenchymal stromal cells. Oncogene 2020, 39, 1198–1212. [Google Scholar] [CrossRef]
  164. Nehrbas, J.; Butler, J.T.; Chen, D.-W.; Kurre, P. Extracellular vesicles and chemotherapy resistance in the AML microenvironment. Front. Oncol. 2020, 10, 90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Mudgapalli, N.; Nallasamy, P.; Chava, H.; Chava, S.; Pathania, A.S.; Gunda, V.; Gorantla, S.; Pandey, M.K.; Gupta, S.C.; Challagundla, K.B. The role of exosomes and MYC in therapy resistance of acute myeloid leukemia: Challenges and opportunities. Mol. Aspects Med. 2019, 70, 21–32. [Google Scholar] [CrossRef]
  166. Pando, A.; Reagan, J.L.; Quesenberry, P.; Fast, L.D. Extracellular vesicles in leukemia. Leuk. Res. 2018, 64, 52–60. [Google Scholar] [CrossRef]
  167. Boyiadzis, M.; Whiteside, T.L. Exosomes in acute myeloid leukemia inhibit hematopoiesis. Curr. Opin. Hematol. 2018, 25, 279–284. [Google Scholar] [CrossRef] [PubMed]
  168. Huan, J.; Hornick, N.I.; Shurtleff, M.J.; Skinner, A.M.; Goloviznina, N.A.; Roberts, C.T., Jr.; Kurre, P. RNA trafficking by acute myelogenous leukemia exosomes. Cancer Res. 2013, 73, 918–929. [Google Scholar] [CrossRef] [Green Version]
  169. Ye, H.; Minhajuddin, M.; Krug, A.; Pei, S.; Chou, C.H.; Culp-Hill, R.; Ponder, J.; De Bloois, E.; Schniedewind, B.; Amaya, M.L.; et al. The Hepatic Microenvironment Uniquely Protects Leukemia Cells through Induction of Growth and Survival Pathways Mediated by LIPG. Cancer Discov. 2021, 11, 500–519. [Google Scholar] [CrossRef]
  170. DiNardo, C.D.; Pratz, K.; Pullarkat, V.; Jonas, B.A.; Arellano, M.; Becker, P.S.; Frankfurt, O.; Konopleva, M.; Wei, A.H.; Kantarjian, H.M.; et al. Venetoclax combined with decitabine or azacitidine in treatment-naive, elderly patients with acute myeloid leukemia. Blood 2019, 133, 7–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Konopleva, M.; Pollyea, D.A.; Potluri, J.; Chyla, B.; Hogdal, L.; Busman, T.; McKeegan, E.; Salem, A.H.; Zhu, M.; Ricker, J.L.; et al. Efficacy and Biological Correlates of Response in a Phase II Study of Venetoclax Monotherapy in Patients with Acute Myelogenous Leukemia. Cancer Discov. 2016, 6, 1106–1117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Jones, C.L.; Stevens, B.M.; D’Alessandro, A.; Reisz, J.A.; Culp-Hill, R.; Nemkov, T.; Pei, S.; Khan, N.; Adane, B.; Ye, H.; et al. Inhibition of Amino Acid Metabolism Selectively Targets Human Leukemia Stem Cells. Cancer Cell 2018, 34, 724–740 e4. [Google Scholar] [CrossRef] [Green Version]
  173. Jones, C.L.; Stevens, B.M.; Pollyea, D.A.; Culp-Hill, R.; Reisz, J.A.; Nemkov, T.; Gehrke, S.; Gamboni, F.; Krug, A.; Winters, A.; et al. Nicotinamide Metabolism Mediates Resistance to Venetoclax in Relapsed Acute Myeloid Leukemia Stem Cells. Cell Stem Cell 2020, 27, 748–764.e4. [Google Scholar] [CrossRef] [PubMed]
  174. Hole, P.S.; Zabkiewicz, J.; Munje, C.; Newton, Z.; Pearn, L.; White, P.; Marquez, N.; Hills, R.K.; Burnett, A.K.; Tonks, A.; et al. Overproduction of NOX-derived ROS in AML promotes proliferation and is associated with defective oxidative stress signaling. Blood 2013, 122, 3322–3330. [Google Scholar] [CrossRef] [PubMed]
  175. Corydon, T.J.; Bross, P.; Holst, H.U.; Neve, S.; Kristiansen, K.; Gregersen, N.; Bolund, L. A human homologue of Escherichia coli ClpP caseinolytic protease: Recombinant expression, intracellular processing and subcellular localization. Biochem. J. 1998, 331, 309–316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Cole, A.; Wang, Z.; Coyaud, E.; Voisin, V.; Gronda, M.; Jitkova, Y.; Mattson, R.; Hurren, R.; Babovic, S.; Maclean, N.; et al. Inhibition of the Mitochondrial Protease ClpP as a Therapeutic Strategy for Human Acute Myeloid Leukemia. Cancer Cell 2015, 27, 864–876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Ishizawa, J.; Zarabi, S.F.; Davis, R.E.; Halgas, O.; Nii, T.; Jitkova, Y.; Zhao, R.; St-Germain, J.; Heese, L.E.; Egan, G.; et al. Mitochondrial ClpP-Mediated Proteolysis Induces Selective Cancer Cell Lethality. Cancer Cell 2019, 35, 721–737.e9. [Google Scholar] [CrossRef]
  178. Chen, X.; Glytsou, C.; Zhou, H.; Narang, S.; Reyna, D.E.; Lopez, A.; Sakellaropoulos, T.; Gong, Y.; Kloetgen, A.; Yap, Y.S.; et al. Targeting Mitochondrial Structure Sensitizes Acute Myeloid Leukemia to Venetoclax Treatment. Cancer Discov. 2019, 9, 890–909. [Google Scholar] [CrossRef]
  179. Reddy, M.M.; Fernandes, M.S.; Salgia, R.; Levine, R.L.; Griffin, J.D.; Sattler, M. NADPH oxidases regulate cell growth and migration in myeloid cells transformed by oncogenic tyrosine kinases. Leukemia 2011, 25, 281–289. [Google Scholar] [CrossRef] [Green Version]
  180. Godfrey, R.; Arora, D.; Bauer, R.; Stopp, S.; Muller, J.P.; Heinrich, T.; Bohmer, S.A.; Dagnell, M.; Schnetzke, U.; Scholl, S.; et al. Cell transformation by FLT3 ITD in acute myeloid leukemia involves oxidative inactivation of the tumor suppressor protein-tyrosine phosphatase DEP-1/ PTPRJ. Blood 2012, 119, 4499–4511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Mason, C.C.; Fiol, C.R.; Baker, M.J.; Nadal-Melsio, E.; Yebra-Fernandez, E.; Bicalho, L.; Chowdhury, A.; Albert, M.; Reid, A.G.; Claudiani, S.; et al. Identification of genetic targets in acute myeloid leukaemia for designing targeted therapy. Br. J. Haematol. 2021, 192, 137–145. [Google Scholar] [CrossRef] [PubMed]
  182. Adane, B.; Ye, H.; Khan, N.; Pei, S.; Minhajuddin, M.; Stevens, B.M.; Jones, C.L.; D’Alessandro, A.; Reisz, J.A.; Zaberezhnyy, V.; et al. The Hematopoietic Oxidase NOX2 Regulates Self-Renewal of Leukemic Stem Cells. Cell Rep. 2019, 27, 238–254.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The endosteal and vascular bone marrow niche. The endosteal niche hosts quiescent or self-renewing HSCs. The vascular niche hosts differentiating HSCs using cell-cell interactions and secreted molecules. This figure is adopted from [98]. CAR cells, CXCL12-abundant reticular cells; HSC, Hematopoietic stem cells; MSC, mesenchymal stem cells; MPC, Myeloid progenitor cells; PGE2, Prostaglandin E2; SCF, Stem Cell Factor; SNO cell, spindle-shaped N-cadherin+CD45- osteoblastic cell; TNF-α, Tumors Necrosis Factor α; TPO, Thrombopoietin.
Figure 1. The endosteal and vascular bone marrow niche. The endosteal niche hosts quiescent or self-renewing HSCs. The vascular niche hosts differentiating HSCs using cell-cell interactions and secreted molecules. This figure is adopted from [98]. CAR cells, CXCL12-abundant reticular cells; HSC, Hematopoietic stem cells; MSC, mesenchymal stem cells; MPC, Myeloid progenitor cells; PGE2, Prostaglandin E2; SCF, Stem Cell Factor; SNO cell, spindle-shaped N-cadherin+CD45- osteoblastic cell; TNF-α, Tumors Necrosis Factor α; TPO, Thrombopoietin.
Cells 10 02833 g001
Figure 2. Activation of different signaling pathways in a leukemic stem cell. AKT, PKB (Protein kinase B); BCL2, B-cell lymphoma 2; GPCR, G-protein-coupled receptor; JAK, Janus kinase; LRP, lipoprotein receptor-related protein; mTOR, mechanistic target of rapamycin; N-ICD, Notch-intracellular domain; NF-κB, Nuclear factor-kappaB; PI3K, Phosphoinositide 3-kinases; PKC, Protein kinase C; RTK, Receptor tyrosine kinases; STAT, Signal transducer and activator of transcription.
Figure 2. Activation of different signaling pathways in a leukemic stem cell. AKT, PKB (Protein kinase B); BCL2, B-cell lymphoma 2; GPCR, G-protein-coupled receptor; JAK, Janus kinase; LRP, lipoprotein receptor-related protein; mTOR, mechanistic target of rapamycin; N-ICD, Notch-intracellular domain; NF-κB, Nuclear factor-kappaB; PI3K, Phosphoinositide 3-kinases; PKC, Protein kinase C; RTK, Receptor tyrosine kinases; STAT, Signal transducer and activator of transcription.
Cells 10 02833 g002
Table 1. WHO classification of AML subtypes [27].
Table 1. WHO classification of AML subtypes [27].
NumberGenomic Classification of AMLRate
1NPM1-mutated AML27%
2AML with mutated chromatin and/or RNA-splicing genes which include (RUNX1, MLL, SRSF2, ASXL1, STAG2)18%
3AML with TP53 mutations and/or chromosomal aneuploidy13%
4AML with inv (16) (p13.1q22) or t(16;16) (p13.1; q22); CBFB–MYH115%
5AML with biallelic CEBPA mutations4%
6AML with t (15;17) (q22; q12); PML–RARA4%
7AML with t (8;21) (q22; q22); RUNX1–RUNX1T14%
8AML with MLL fusion genes; t(x;11) (x; q23)3%
9AML with inv (3) (q21q26.2) or t (3;3) (q21; q26.2); GATA2, MECOM (EVI1)1%
10AML with IDH2R172 mutations and no other class-defining lesions1%
11AML with t (6;9) (p23; q34); DEK–NUP2141%
Table 2. Medications with the purpose of AML targeted therapy.
Table 2. Medications with the purpose of AML targeted therapy.
FunctionNameTargetMechanismFDA ApprovedRefs
IDH1 inhibitorIvosidenibIDH1Myeloblast differentiation induction through isocitrate dehydrogenase 1 (IDH1) inhibition and 2-hydroxyglutarate (2-HG) blockageYes[46]
IDH2 inhibitorEnasidenibIDH2Myeloblast differentiation induction through isocitrate dehydrogenase 2 (IDH2) inhibition and 2-HG blockageYes[46]
FLT3 inhibitorGilteritinibFLT3-TKD
  • FLT3-I inhibition
  • AXL receptor tyrosine kinase inhibition
  • FLT3-TKD and FLT3-D835 TKD receptor antagonist
Yes[47]
QuizartinibFLT3-ITD
  • FLT3 second generation inhibitor
  • Tumor cell apoptosis inducer
No[47,48]
Antibody drug conjugate (ABDC)Gemtuzumab ozogamicin (GO)CD33Anti-CD33 monoclonal antibody conjugated with cytotoxin Yes[46]
Selective E-selectin antagonistUproleselan (GMI-1271)E-selectinChemotherapy sensitizerNo[46]
Table 3. Transcription factor roles in AML.
Table 3. Transcription factor roles in AML.
TFEffectsTherapeuticsRefs
NF-E2 related factor-2 (NRF2)1. Reactive oxygen species (ROS) neutralization
2. Chemotherapy resistant
3. Antioxidant response element (ARE) up-regulation
Brusatol[70,71]
CCAAT/enhancer binding protein alpha (C/EBPα)1. Tumor suppressor
2. Activated by TP53-KLF4
3. Down-regulated in AML due to TP53 down-regulation
4. Drug resistance
5. CSF3R, MPO, and ELANE up-regulation
ICCB280
NSC23766
OICR-9429
C/EBPA-siRNA
[71,72]
TP531. Tumor suppressor
2. Down-regulated in AML
3. Severe drug resistance
4. BAX and CDKN1A up-regulation
PRIMA-1
PRIMA-1MET
SAR405838
AM-8553
AMG232
MK-8242
DS-3032b
CGM097
[71,73]
c-MYCUp-regulated in AML
1. Leukemic cells proliferation enhancement
2. Chemotherapy resistance
3. BCL-2, CDKN1A and CCND1 up-regulation
IIA6B17
NY2267
MYRA-A
10074-G5
Mycro3
JQ-1
[71,74]
STAT3Up-regulated in AML
1. Chemotherapy resistance
2. Pro-survival
3. Proliferation enhancement
4. Anti-apoptotic
5. BCL-2, BCL-XL, Mc1-1, cyclin D1, and c-MYC up-regulation
Galiellalactone[71,75,76]
Krüppel-like factor 4 (KLF4)1. Tumor suppressor
2. Cell cycle arrest by CDKN1A suppression
3. Down-regulated in AML (NPM1-mutant)
4. Down-regulation is correlated with chemoresistance
5. P21, P27 up-regulation
6. Suppressed by metal-regulatory transcription factor 1 (MTF-1)
APTO-253[69,71,72,77]
cAMP response element-binding protein (CREB)Up-regulated in AML
1. Pro-survival
2. Anti-apoptotic
3. Chemotherapy resistance
4. Up-regulates BCL-2
5. Up regulates transcription of numerous gens such as c-fos, junB, and egr-1
STF-017794
STF-038533
STF-046536
STF-046728
STF-055910
[69,71,78,79,80]
PU.1Up-regulated in AML
1. Up-regulates CSF1R, IL7R, CD11b, M-CSFR, GM-CSFR, G-CSFR
2. Hematopoiesis defect in AML
DB2313
DB2115
DB1976
[71,81]
Runt-related transcription factor 1 (RUNX1)Up-regulated in AML
1. Up-regulates C/EBPα, PU.1, and cell cycle progression
2. Down-regulates TP53
Chb-M
Chb-50
[71,82]
NF-κBUp-regulated in AML
Poor prognostic factor
1. Up-regulates BCL-2 and BCL-XL
2. Pro-survival
3. Feed-back positive effect with TNF-α in AML
Bortezomib (FDA)[83,84,85]
Table 5. BM cytokine and chemokine network interrelationship in AML.
Table 5. BM cytokine and chemokine network interrelationship in AML.
ReceptorCell(s)LigandLigand SourceNormal FunctionExpression in AMLRefs
CXCR41. Most immune cells
2. AML leukemic cells
SDF-1 (CXCL12)1. MSC
2. Leukemic cells
1.Chemotaxis
2. Migration
3. Pro-survival
1. Chemotherapy resistance
2. Pro-survivalthrough PI3K/AKT and MEK/ERK activation
[44,95,102,103,104]
VCAM-1 (CD106, fibronectin)Stromal cellsVery late antigen 4 (VLA-4)1. HSC and hematopoietic progenitors
2. Monocytes (MO)
3. Leukemic cells
4. Myeloid cells
5. Immature dendritic cells
6. Neutrophils
7. Eosinophils
8. Immature mast cells
9. Endothelial cells
1. Adhesion
2. Pro-survival
3. Proliferation
1. Pro-survival
2. Proliferative
3. NF-κB activation
4. Chemotherapy resistance
5. MRD and relapse
[62,95,105,106]
RANKNK cellRANKL or Tumor necrosis factor-receptor (TNF-R)1. Stromal cells
2. Osteoblast
3. Activated lymphocyte
4. Leukemic cells
Bone remodelingNK cell inhibitory[44]
c-MPL (CD 110)1. HSC
2. Megakaryocyte (MK)
3. Chronic myeloid leukemia (CML)
4. AML leukemic cells
TPO1. Liver
2. Kidney
1. HSC quiescence
2. Thrombopoiesis
Chemotherapy resistance[87,107]
Vascular endothelial growth factor receptor(VEGFR)1. MO
2. MQ
3. Vascularendothelial cells (VEC)
4. Lymphoid endothelial cells (LEC)
5. HSC
1. VEGF
2. PIGF
1. Stromal cell
2. MK
3. HSC
4. Leukemic cells
1. GM-CSF stimulation
2. Angiogenesis
3. Metabolichomeostasis
4. Proliferation
5. Migration
6. Tubulogenesis
1. Anti-apoptotic
2. Chemotherapy resistance
[32,95,108]
E-Selectin1. Endothelial cells
2. Stromal cell
CD441. HSC and Hematopoietic progenitors
2. T cells
3. Leukemic stem cells
4. Stromal cells
1. HSC pro-survival
2. Proliferation of HSCs
1. E-selectin: chemotherapy resistance
2. CD44:
Pro-survival
[95,104,105,109]
IL-1R11. Most hematopoietic and non-hematopoietic cells
2. AML leukemic cells
IL-1β1. Myeloid lineage
2. Leukemic cells
3. EC
4. MSC
5. MQ
1. Pro-inflammatory
2. Hematopoiesis regulation
1. Pro-survival
2. Pro-proliferative
3. Sometimes feedback positive
4. Association with endogenous IL-1β related to apoptosis resistance
[110,111,112,113,114,115,116]
TNFαRI
(p55 or p60)
A broad spectrum of different cell types like AML cellsTNF-α1. CD8/ CD4 T cell
2. NKT cells
3. Neutrophils
4. Macrophage 1 (MQ1)
5. LSCs
6. MSCs
Pro-inflammatory1. Pro-survival
2. Chemotherapy resistance
3. NF-κB activation
[44,110,113,117,118,119,120]
IFNGR1,21. Widelydistributed on various cell types
2. LSCs
IFN-ϒMost immune cellsPro-inflammatory1. Anti-leukemic
2. Anti-proliferative
3. Antigen presentation through MHC I/II augment
4. Nitric oxide (NO) and reactive oxygen species (ROS) mediators, NADPH, and inducible nitric oxide synthase (INOS) production
[110,118,121,122,123]
IL-10R1. AML leukemic cells
2. T cells
3. B cells
4. NK cells
5. Epithelial cells
6. Endothelial cells
7. Plasmacytoid DCs
8. Peripheral blood mononuclear cells (PBMCs)
IL-101. T helper 2 (TH 2)
2. BM-MSCs
3. Macrophage 2 (MQ2)
4. T-reg
5. B cells
6. MO
7. Thymocytes
Anti-inflammatory TH1 suppressor1. Growth arrest-specific gene 6 (Gas6) up-regulation
2. Pro-survival
3. Chemotherapy resistance
[118,123,124,125,126,127,128,129]
TGF-βR1. T cell
2. Hematopoietic progenitor cells
3. AML leukemic cells
TGF-β1. T-reg
2. MQ2
3. MSC
4. Endothelial cells
5. Platelets
6. PBMCs
1. Anti-inflammatory
2. Proliferation
3. Migration
4. Pro-survival
5. Growth and differentiation inhibition of hematopoietic progenitor cells
1. Anti-proliferative
2. IL-1β, IL-6, GM-CSF, and granulocyte colony-stimulating factor (G-CSF) production
3. Reduction in AML
[110,118,126,130]
IL1R11. Most hematopoietic and non-hematopoietic cells
2. AML leukemic cells
IL-1Ra1. MQ 2
2. MO
4. Neu
6. Fibroblasts
7. Chondrocytes
1. Anti-inflammatory
2. IL-1 antagonist
Leukemic cell colonization inhibitor[110,112,131,132]
IL-35R1. Effector T cells
2. CD4+ T-reg
3. AML leukemic cells
IL-351. T-reg
2. DCs
3. B-reg
4. sometimes in endothelial cells, monocytes and smooth muscle cells
1. Anti-inflammatory
2. Inhibits T cell proliferation
3. Transformation of T cells to iTreg
1. Anti-apoptotic
2. Proliferation
3. Weak prognosis
4. AML progression
[110,118,133]
PD1
(CD279)
LymphocytesProgrammed death-ligand 1 (PDL1) (CD274) (B7-H1)1. T-reg
2. Follicular T cells (FTC)
3.MQ
4. Dendritic cell (DC)
5. placental syncytiotrophoblasts
6. MO
7. AML leukemic cells
T cell activation and proliferation inhibitor1. Pro-survival
2. Weak prognosis
[118,134]
Lymphocyte activation gene-3
(LAG3)
T cellMHC IIAPCsT cells inhibitory 1. Correlation with programmed death-1 (PD1)
2. Increased activity of leukemic cells
[118,135]
Galectin-9 (Gal-9)1. AML LSC
2. Lymphocytes
3. Spleen
4. Thymus
T-cell immunoglobin mucin-3
(TIM-3)
1. AML leukemic cells
2. MO
3. DC
4. Some of T cells
5. NK cells
6. Myeloid pre-leukemic progenitors
Not in normal HSCs
1. TH1 inhibitory
2. DC maturation
3. TNF-α secretion from monocytes
4. Innate immune system activation
Strong
self-renewal signaling through TIM-3/Gal-9 autocrine loop, NF-κβ and β-catenin signaling
Up-regulated in pre-leukemic disorders
[136]
Cytotoxic T-lymphocyte antigen-4 (CTLA-4) or (CD152) 1. T cells
2. AML leukemic cells
Β7-1
Β7-2
Antigen-presenting cells (APCs)T-cell inhibitory and tolerance induction 1. AML relapse and MRD
2. Immune evasion
Blockage leads to sensitivity to cytotoxic T lymphocytes (CTL)
[134,137]
Table 6. Signaling pathways related to AML drug resistance.
Table 6. Signaling pathways related to AML drug resistance.
Signaling PathwayLeukemic EffectMechanismTherapeuticsActivator Ligand (L)
Receptor (R)
Mediators (M)
Target (T)
Refs
JAK/STATChemo-therapy resistance1. Proliferation
2. Pro-survival
1. Ruxolitinib (FDA)
2. Ruxolitinib
3. Pacritinib
4. Lestaurtinib
5. Fedratinib
6. Momelotinib
L: TPO/MPL/G-CSF
R: Cytokine receptor superfamily
M: JAK2, STAT3, STAT5, TYK2
T: p21, Mcl-1, PIM1, BCL-2, BCL-XL
[142,143]
Notch11. Poor prognosis
2. Chemotherapy resistance
1. Rb phosphorylation
2. C-MYC and BCL-2 up-regulation
3. Pro-survival
4. Proliferation
5. Connection to Delta-1 leads to NF-κB pathway activation
GSIs
(GSI-IX and GSI-XII)
L: Deltalike1,4 Jagged1
R: NOTCH1
M: Notch intracellular domain of Notch (N-ICN)
T:
1. CSL activity
Hes family: HES1, HES5
Hes-related repressor proteins (Herps) family: HERP2
2. DELTEX1
[144,145]
Hedgehog (Hh)1. Poor prognosis
2. Chemotherapy resistance
Activated in AML through GLI1 and SMO up-regulation1. LDE225 (Sonidegib)
2. PF-04449913 (Glasdegib)
3. Vismodegib (GDC-0449)
4. BMS-833923 (XL139)
5. GANT-61
L: Hh proteins
R: PTCH1 and SMO
M: GLI1
T: BCL-2, SNAIL, RAS, TGF-β, c-MYC
[146]
Ras/Raf/MEK/ERK1. Chemotherapy resistance
2. Leukemic cell survival
1. Anti-apoptotic
2. Pro-survival through Raf-1 downstream molecule phosphorylation
1. L-779,450
2. ZM 336372
3. Bay 43-9006
4. Geldanamycin
5. Coumermycin
5. Dasatinib
6. PD98059
7. U0126
8. PD184352
9. ARRY142886
L:
1. Ras proteins (Ha-Ras, N-Ras, Ki-Ras 4A, Ki-Ras 4B)
2. Protein kinase C (PKC)
R: Receptor tyrosine kinases (RTK)
M:
Raf-1, A-Raf and B-Raf
T:
1. Transcription factors, including Ets-1, c-Jun
and c-MYC
CREB
NF-κB
2. Bad, Bim, Mcl-1, caspase 9, BCL-2
[147,148]
Phosphatidy-linositol 3-kinase (PI3K)/Akt/mTOR1. Poor prognosis
2. Chemotherapy resistance
1. Glycolysis up-regulation
2. Proliferation
3. Pro-survival
1. Ridaforolimus
2. Sirolimus (Rapamycin)
3. Everolimus
4. Temsirolimus
L: Wide variety of extracellular stimuli
R: G-protein-coupled receptors (GPCRs)
RTK, various integrins, B and T cell receptors
M: Akt, mTOR
T: p70S6K, S6RP, 4EBP1
[53,149]
Wnt1. Poor prognosis
2. Chemotherapy resistance
1. LSC self- renewal
2. AML progression
1. Celecoxib
2. CWP232291
3. LY2090314
4. PRI-724
5. Sulindac
L: Wnt1
Wnt3a, PCP
R: Frizzled (FZD) and lipoprotein receptor-related protein (LRP)
M: β-catenin, Ca2+
T: cyclin D1, c-MYC, Hox genes, MLL/ENL

[150,151]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bolandi, S.M.; Pakjoo, M.; Beigi, P.; Kiani, M.; Allahgholipour, A.; Goudarzi, N.; Khorashad, J.S.; Eiring, A.M. A Role for the Bone Marrow Microenvironment in Drug Resistance of Acute Myeloid Leukemia. Cells 2021, 10, 2833. https://doi.org/10.3390/cells10112833

AMA Style

Bolandi SM, Pakjoo M, Beigi P, Kiani M, Allahgholipour A, Goudarzi N, Khorashad JS, Eiring AM. A Role for the Bone Marrow Microenvironment in Drug Resistance of Acute Myeloid Leukemia. Cells. 2021; 10(11):2833. https://doi.org/10.3390/cells10112833

Chicago/Turabian Style

Bolandi, Seyed Mohammadreza, Mahdi Pakjoo, Peyman Beigi, Mohammad Kiani, Ali Allahgholipour, Negar Goudarzi, Jamshid S. Khorashad, and Anna M. Eiring. 2021. "A Role for the Bone Marrow Microenvironment in Drug Resistance of Acute Myeloid Leukemia" Cells 10, no. 11: 2833. https://doi.org/10.3390/cells10112833

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