Next Article in Journal
Enhanced NAMPT-Mediated NAD Salvage Pathway Contributes to Psoriasis Pathogenesis by Amplifying Epithelial Auto-Inflammatory Circuits
Next Article in Special Issue
Antileukemic Natural Product Induced Both Apoptotic and Pyroptotic Programmed Cell Death and Differentiation Effect
Previous Article in Journal
SIRT3 Overexpression Ameliorates Asbestos-Induced Pulmonary Fibrosis, mt-DNA Damage, and Lung Fibrogenic Monocyte Recruitment
Previous Article in Special Issue
Non-Coding RNA Signatures of B-Cell Acute Lymphoblastic Leukemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Hypoxic Bone Marrow Microenvironment in Acute Myeloid Leukemia and Future Therapeutic Opportunities

1
Department of Experimental, Diagnostic and Specialty Medicine, University of Bologna, 40138 Bologna, Italy
2
Istituto di Ematologia “Seràgnoli”, IRCCS Azienda Ospedaliero, Universitaria di Bologna, 40138 Bologna, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(13), 6857; https://doi.org/10.3390/ijms22136857
Submission received: 25 May 2021 / Revised: 18 June 2021 / Accepted: 22 June 2021 / Published: 25 June 2021
(This article belongs to the Special Issue Advances in Molecular Biology and Targeted Therapy of Leukemias 3.0)

Abstract

:
Acute myeloid leukemia (AML) is a hematologic malignancy caused by a wide range of alterations responsible for a high grade of heterogeneity among patients. Several studies have demonstrated that the hypoxic bone marrow microenvironment (BMM) plays a crucial role in AML pathogenesis and therapy response. This review article summarizes the current literature regarding the effects of the dynamic crosstalk between leukemic stem cells (LSCs) and hypoxic BMM. The interaction between LSCs and hypoxic BMM regulates fundamental cell fate decisions, including survival, self-renewal, and proliferation capacity as a consequence of genetic, transcriptional, and metabolic adaptation of LSCs mediated by hypoxia-inducible factors (HIFs). HIF-1α and some of their targets have been associated with poor prognosis in AML. It has been demonstrated that the hypoxic BMM creates a protective niche that mediates resistance to therapy. Therefore, we also highlight how hypoxia hallmarks might be targeted in the future to hit the leukemic population to improve AML patient outcomes.

1. Introduction

Acute myeloid leukemia (AML) is a highly heterogeneous hematologic malignancy that results from a wide range of alterations of myeloid precursors responsible for the uncontrolled proliferation, block of differentiation, and escape from apoptosis. The resulting expansion of leukemic-initiating cells induces bone marrow (BM) failure, leading to normal hematopoietic functions. AML represents the most common acute type of leukemia in adults, affecting approximately 1% of the population [1]. Although AML can occur at any age, the disease incidence increases sharply with age, showing a median age at diagnosis of 68 years.
AML originates from the oncogenic transformation of hematopoietic stem cells (HSCs) that may undergo further genetic alterations leading to the progression of the disease, with multiple coexisting clones that evolve over time [2,3]. The disease phenotype and its clinical course are strikingly influenced by several recurrent mutations, which define the prognostic stratification of patients. Some recurrently mutated genes are now potential targets for a series of newly approved or investigational treatments [4,5,6]. Since 1970, the standard of care for AML patients has been based on chemotherapeutic treatment (the ‘3 + 7’ regimen, combining daunorubicin and cytarabine), resulting in 5-year survival rates of 40–45% for patients up to 55 years, that decrease to 30–35% for patients up to 60 and may become as low as 10–15% in older patients [7,8]. Moreover, the relapsed disease after complete remission still represents the most common cause of death among AML patients, and it is associated with increased molecular complexity that often contributes to treatment failure [9,10]. Therefore, unraveling the cytogenetic and molecular profile of AML has improved the therapeutic opportunities leading to the development of new targeted therapies.
In recent years, however, it has been acknowledged that AML development and progression are dependent on cell-intrinsic oncogenic alterations and are also strongly influenced by the components of the bone marrow microenvironment (BMM) where normal and leukemic stem cells (LSCs) reside. It has been demonstrated that, physiologically, the BMM can be distinguished in two distinct niches, defined endosteal and vascular niche [11,12,13]. The endosteal HSC niche is mainly composed of osteoblasts that support hematopoiesis [11,14], while the HSC vascular niche mainly contains sinusoidal endothelial cells [12], which support HSCs proliferation and differentiation [15,16]. It has also been demonstrated that low pO2 levels characterize the BMM, although highly vascularized [17,18,19]. In particular, it has been shown that O2 levels in BM range from 1% to 4% [17,20], decreasing from vessels to the endosteum—hence the term “hypoxic niche.” Tumor hypoxia has been first described as a pathogenic condition in most solid cancers. It has been associated with increased tumor aggressiveness, immunosuppression, and decreased sensitivity to radiotherapy and chemotherapy [21]. The earliest evidence about the presence of a hypoxic niche arose from both in vivo studies, which confirmed that HSCs reside in hypoxic BM areas with diminished blood perfusion [22,23,24,25], and from in vitro hypoxic cultures (1%–3% O2) that elucidated the role of hypoxia in promoting quiescence and self-renewal of HSCs [26,27,28], as well as their differentiation into erythroid, megakaryocytic, and granulocytic-monocytic progenitors [29,30,31,32]. The peculiar property of hypoxic BMM is that it represents a physiologic condition to control HSCs homeostasis. Still, in the case of leukemic transformation, it becomes inhospitable to normal HSCs creating a ”malignant niche” that sustains survival and proliferation of LSCs [33]. A growing number of studies have provided valuable insights into the crucial role of hypoxia in the pathophysiology of AML. The increasing interest in the hypoxic niche arises from its dual role in regulating fundamental cell fate decisions for HSCs and LSCs. Therefore, in this article, we will review the current understanding of the pathogenetic impact of hypoxic BMM in AML, and we will outline some hypoxia-targeted treatment strategies that could be further developed in the future.

2. Regulation of Response to Hypoxia

Hypoxia-inducible factor (HIFs) proteins mediate the major adaptive response to hypoxia by introducing many genes required for adaptation to hypoxia. The HIF family includes three O2-sensitive α-subunits (HIF-1α, HIF-2α, and HIF-3α) and one stable β-subunit (HIF-1β/ARNT). They work as a heterodimer through the interaction between α and β subunits. In normal O2 conditions, the two α subunits are hydroxylated by the interaction with prolyl hydroxylases (PHDs) located on von Hippel-Lindau-associated E3-ubiquitin ligase (VHL), and then are degraded by the proteasome to ensure a rapid turnover. On the contrary, under hypoxic conditions, the PHDs are inactive because they require molecular oxygen as a cofactor. Thus, the binding of α subunits to VHL is inhibited. The final consequence is stabilizing the oxygen-sensitive subunit and its translocation into the nucleus, where it can dimerize with HIF-1β. This heterodimer, in turn, interacts with the transcriptional co-activator p300/CREB-binding protein (CBP) to recognize the hypoxia-responsive elements (HREs), leading to the genetic expression designed to boost the adaptive response to hypoxia [34] (Figure 1). The expression of HIF downstream genes has been associated with many cellular functions, including angiogenesis, energetic metabolism (glycolysis, tricarboxylic acid cycle, and oxidative phosphorylation), proliferation and survival, autophagy, proteolysis, and pH regulation [35,36,37].

3. Effects of Hypoxia and HIFs on HSCs

Since the concept of the hypoxic microenvironment was proposed, many studies have investigated the role of hypoxia role on HSCs functions, building from previous observations on solid cancers that demonstrated the crucial impact of HIF proteins on cellular transcriptional programs. Several in vivo imaging studies have demonstrated that HSCs preferentially localize in poorly perfused endosteal niches [38,39] that promote quiescence and pluripotent state of HSCs [40] and support their expansion in response to damage signals [23,41,42]. In contrast, it was observed that HSCs residing in the vascular niche are short-term proliferating cells responsible for replenishing circulating cells [16,38] (Figure 2). Furthermore, an immune-protective property of hypoxic niches has been documented mediated by regulatory T cells (Treg) that preferentially localize in the endosteal niche where they shelter HSCs from CD4 and CD8 T cell immune attack, enabling transplanted allo-HSCs to escape from allogeneic rejection [43,44].
Among the three α subunits, HIF-1α and HIF-2α have been more comprehensively studied, while less is known about HIF-3α. HIF-1α expression in HSCs is transcriptionally controlled by the homeodomain transcription factor Meis1 [45], while HIF-2α is a target of Stat5 [46]. In accordance with the hypoxic environment of the BM, it has been reported that normal HSCs show enrichment of HIF-1α [45,47] and that its genetic deletion results in loss of quiescence and reduction of repopulating activity of HSCs [47,48]. Moreover, through in vivo and in vitro studies, it has been demonstrated that HIF-1α accumulation obtained by hypoxic culture conditions, as well as its genetic stabilization deriving from VHL deletion or functional inhibition of the PHD domain, is enough to induce HSC quiescence and to enhance their self-renewal potential [49,50,51]. Loss-of-function studies indicated that HSCs isolated from HIF-1α-deficient mice showed increased expression of the hallmark genes of senescent stem cells p16Ink4a and p19Arf [47]. Therefore, much evidence supports the crucial role of HIF-1α as a master regulator of quiescence, self-renewal capacity, and survival of HSCs [52]. As previously mentioned, most of the adaptive changes induced by HIF-1α have been associated with the expression of several genes involved in numerous cellular processes, making them important players for HSC function. They include those encoding for the vascular endothelial growth factor (VEGF), stromal cell-derived factor 1 (SDF1), stem cell factor (SCF), angiopoietin 2 (ANGPT2), angiopoietin-like proteins (ANGPTLs), insulin-like growth factor 2 (IGF-2), insulin-like growth factor-binding proteins (IGFBPs), forkhead box O (FOXOs) and the cyclin-dependent kinase inhibitors p16, p19 and p21 [35,36,37,47,53]. In addition, HIF-1α has been deeply characterized as a metabolic regulator, able to favor the switch from oxidative to glycolytic metabolism, thereby limiting the mitochondrial potential to allow the quiescence state of HSCs [54,55,56]. Indeed, HIF-1α activates the transcription of glucose transporters, glycolytic enzymes, and glycolytic inducing factors [54,57]. Its conditional deletion in HSCs results in a metabolic shift to oxidative metabolism, increasing oxidative stress [58].
While the crucial role of HIF-1α on HSC maintenance has been clarified, some incertitude remains regarding the exact role of HIF-2α. HIF-2α deletion does not change HSC number or steady-state hematopoiesis [59]. However, it induces increased reactive oxygen species (ROS) production, which in turn stimulates endoplasmic reticulum stress and apoptosis together with a considerable engraftment reduction during xenotransplantation [60].
Collectively, these studies demonstrated that the hypoxic BMM sustains the cellular homeostasis of HSCs through the regulation of the gene expression programs mostly mediated by HIFs proteins.

4. HIFs in AML

The peculiar property of hypoxic BMM is related to its ambivalent role in physiologic and pathologic conditions. Therefore, understanding the complex interplay between the hypoxic microenvironment and the different cellular populations that reside in the BM could contribute to identifying pathway alterations leading to leukemic transformation. Thus far, many studies have demonstrated the involvement of BMM in the initiation and propagation of leukemia. Several pieces of evidence concur to demonstrate that the hematopoietic niches sustain survival, proliferation, and differentiation of LSCs [12,13,14,61,62,63]. A series of studies have also shown that the crosstalk between LSCs and the BM niche plays a dual role, being involved in both leukemia-induced microenvironment reprogramming and microenvironment-induced leukemogenesis mechanisms. Indeed, it has reported that engraftment and proliferation of CD34 + LSCs in BM induce alterations in the stromal microenvironment, creating a “malignant niche” inhospitable to normal HSCs [64,65].
Regarding the expression of HIF-1α in AML, it has been demonstrated that out of 84 cases of normal karyotype AML, about 20% displayed leukemic cells expressing high levels of cytoplasmic HIF-1α [66]. In another study, Wang et al. investigated HIF-1α expression in AML, showing that HIF-1α is up-regulated in CD34 + CD38-human AML primary cells. Moreover, they demonstrated that HIF-1α silencing and its inhibition by echinomycin induced apoptosis in LSCs and impaired their ability to reconstitute AML into xenotransplanted mice [67]. As far as the studies that have explored the expression levels of HIF-2α are concerned, it has been reported that in a cohort of 33 AML cases, 10 displayed higher levels of HIF-2α compared to normal CD34 + cells [60]. Moreover, the same study showed that HIF-2α silencing in leukemic blasts significantly reduced the leukemic engraftment into immunodeficient mice [60]. Finally, Vukovic et al. observed that HIF-2α deletion accelerated leukemia initiation in mixed-lineage leukemia fusion driven (MLL-AF9) mice models and that this acceleration was further potentiated by HIF-1α co-deletion [68].
This evidence indicates that understanding the crosstalk between LSCs and BM niche is essential to better understand the leukemogenic process and some therapy resistance/evasion mechanisms adopted by LSCs. It provides the rationale to develop novel therapeutic approaches to target the cytoprotective mechanisms deriving from the “malignant niche.”

5. Metabolic Alterations Induced by Hypoxia

One of the best-described alterations induced by hypoxia is the regulation of energetic metabolism, which is susceptible to a marked change mostly mediated by the transcriptional regulation of several key enzymes. Hypoxia and HIF family proteins have already been demonstrated to affect cellular metabolism through the switch from mitochondrial oxidative phosphorylation (OXPHOS) to glycolysis to avoid oxidative stress related to OXPHOS function in low O2 levels conditions. Metabolic analyses performed in AML cell lines and primary blasts confirmed that the major metabolic changes associated with the hypoxic microenvironment include those involved in energy production. However, the available data in AML models are still controversial, suggesting there might be a certain degree of dependence on specific cell-intrinsic factors [69]. On the one hand, Lodi and colleagues [70] described similar metabolic effects induced by hypoxia in two different models of leukemia, the AML cell line KG-1 and the chronic myeloid leukemia (CML) cell line K562. They observed that, although the two cell lines have different metabolic profiles in normoxia, the metabolic changes were strikingly similar when cultured under hypoxia (1% O2), suggesting a common adaptive response to hypoxia. In particular, the study described that hypoxia induces enrichment of the glycolytic pathway with increased levels of lactate and alanine and citrate, fumarate, and succinate, which are tricarboxylic acid (TCA) metabolites [70]. Moreover, both cell lines showed increased concentration of metabolites of the hexosamine pathway such as uridine diphosphate N-acetylglucosamine (UDP-GlcNAc) and uridine diphosphate N-acetylgalactosamine (UDP-GalNAc) [70], which was previously associated with response to several cellular stresses, including hypoxia [71].
On the other hand, we compared the metabolomic alterations induced by hypoxia on OCI-AML3 and KASUMI-1 AML cell lines and showed that hypoxic adaption might rather be dependent on the genomic background [72]. After 20 h of hypoxia, OCI-AML3 cells showed reduced pyruvate levels. They increased 2-hydroxyglutarate (2-HG), fumarate, succinate, lactate, and alanine levels, whereas KASUMI-1 exhibited increased levels of alanine, glutamine, glutamate, and fatty acids with little impact on the TCA cycle intermediates [72]. Lastly, the study published by Goto et al. analyzed the differences in growth and survival of the acute monocytic leukemia THP-1 and the acute promyelocytic leukemia NB-4 cell lines under hypoxic culture conditions. They reported that whereas NB4 cells grew slowly and showed increased ROS production after 48 h of hypoxia, THP-1 cells quickly adapted to hypoxic culture conditions, avoiding ROS production. They did so by changing their energy metabolism from OXPHOS to glycolysis through the up-regulation of pyruvate dehydrogenase kinase 1 (PDK1), a key glycolytic enzyme that inhibits the conversion of pyruvate to acetyl-coenzyme A.
Moreover, the two cell lines also showed different adaptation mechanisms against ROS-induced damage. NB4 cells cultured for longer periods (7 days) increased the ratio between reduced and oxidized glutathione (GSH/GSSG) in order to avoid oxidative stress, while hypoxic THP-1 treated with a glycolysis inhibitor to force ROS production, showed only a slight reduction of growth without accumulation of ROS, given their ability to exchange the cytochrome C oxidase subunit 4 isoform 1 (COX 4-1) subunit to COX 4-2 [73]. Thus, these two studies conducted in the AML cell lines showed that the specific adaptive changes differed strikingly among the different AML cellular models, highlighting a potential role of the genomic background during adaption to hypoxia.
The metabolic reprogramming of AML cells and their heavy dependence on glucose consumption were confirmed in primary leukemic cells [69,74]. A study by Chen et al. compared the metabolomic profiling between primary cells isolated from 400 AML patients and those from 446 healthy controls. The study identified a metabolic signature ascribable to hypoxia, exhibiting significant alterations in six metabolites including lactate, 2-oxoglutarate, pyruvate, 2HG, and citrate, without any significant differences among different WHO AML subtypes [75]. The authors showed that the increased levels of the metabolites mentioned above, except for citrate, were found to be negatively associated with the overall survival (OS) and event-free survival (EFS) of AML patients [75].
All this evidence supports the crucial influence of metabolic switch for LSC survival in the hypoxic BMM and draws attention to the adaptive differences among the different AML subtypes. This highlights the need to further investigate and deeply understand whether the genomic background can drive the metabolic adaption of LSCs to hypoxia. Figure 3 illustrates the most relevant metabolic alterations driven by hypoxia in leukemic cells.

5.1. Effects of Altered Metabolism on Epigenetic Regulation

The metabolic switch induced by hypoxia is particularly relevant in light of the critical role of some TCA intermediates as epigenetic regulators. As showed in Figure 4, increased levels of 2-HG, succinate, and fumarate could be effectively used as a fuel for cancer cells to promote their growth and proliferation. Since isocitrate dehydrogenase (IDH) gene mutations were identified in AML, their impact on metabolism became evident. Indeed, mutant IDH1 or IDH2 lose their canonical function for converting isocitrate to α-ketoglutarate (α-KG), acquiring the neomorphic activity to catalyze the production of 2-HG [76]. This oncometabolite is structurally similar to α-KG; thereby, it competitively inhibits α-KG-dependent enzymes, deregulating the TCA cycle and inducing histone- and DNA-hypermethylation inhibition of epigenetic regulators [77,78]. The final consequence of 2-HG over-production in IDH-mutant AML is impaired cellular growth and differentiation arrest [79,80]. Concerning 2-HG production observed under hypoxia, it has been reported that in this condition, malate dehydrogenase, lactate dehydrogenase, and phosphoglycerate dehydrogenase could generate 2-HG, in the absence of IDH mutations, via enzymatic reduction of α-KG [81]. Succinate and fumarate were also classified as oncometabolite. Both were identified as regulators of some HIF targets via ten-eleven translocation (TET) inhibition [82] and as inhibitors of PHDs [83,84], suggesting a role in the maintenance of hypoxia response as well as in histone and DNA demethylation. Moreover, the succinate dehydrogenase (SDH) complex enzyme, also called mitochondrial complex II, catalyzes the oxidation of succinate to fumarate in the TCA cycle. It represents a metabolic hub for mitochondrial metabolism, also catalyzing the reduction of ubiquinone (UQ) to ubiquinol (UQH2), thereby linking the TCA cycle to the electron transport chain.

5.2. Role of Altered Metabolites in Cellular Signaling

Emerging evidence demonstrates that several metabolic intermediates could function as signaling molecules that control signals and stress response (Figure 4). For example, accumulation of the glycolytic end-product lactate under hypoxia is required for N-Myc downstream-regulated 3 (NDRG3) protein function. This, in turn, mediated the activation of RAF/extracellular-signal-regulated kinase 1 and 2 (ERK1/2) signaling pathways to promote angiogenesis and cell proliferation [85,86]. Additionally, increased levels of succinate and α-KG have been linked with signaling functions given their unexpected ability to act as ligands for the renin-angiotensin system through the G-protein-coupled receptors (GPR91) [87]. The activation of GPR91 mediated by succinate stimulated the proliferation of HSCs through the induction of ERK1/2 and induced multilineage blood cell recovery after chemotherapy in mice models [88]. Furthermore, the oncometabolite 2-HG also acts as a signaling metabolite by inhibiting the activity of ATP synthase. Thus, it reduces the mitochondrial respiration and mammalian target of rapamycin (mTOR) signaling [89]. These pieces of evidence promote the view that mitochondrial metabolites have a dual role in the cell: on one side, they are involved in energy metabolism; on the other side, they exhibit important signaling functions.

6. Gene Expression Regulation Mediated by HIFs to Sustain AML

The role of HIFs proteins as transcription factors has also been associated with their ability to regulate the expression of genes that, in turn, support leukemic transformation and progression. Thus, the communication between BMM and LSCs acts on different cellular functions to sustain AML development and progression and is recognized as a major cause of relapse.
VEGF represents one of the first identified targets of HIF-1α. It is a signal protein mainly involved in both physiologic and pathologic angiogenesis, and its up-regulation has been associated with several cancers, including leukemia. In particular, two different studies have confirmed that HIF-1α and HIF-2α are significantly associated with VEGF-A expression and BM angiogenesis, which is crucial for the pathophysiology of AML [90,91]. Moreover, it has been reported that LSCs also express the VEGFR-2 receptor in vitro and in vivo, leading to an autocrine loop that increases cellular proliferation [92].
Another well-documented HIF-1α target is represented by C-X-C motif chemokine ligand 12 (CXCL12) and its cognate receptor C-X-C motif chemokine receptor 4 (CXCR4), that are mainly involved in cellular homing of HSCs and LSCs in the BM niche [93]. It was demonstrated that CXCR4 over-expression contributes to chemoresistance of leukemic cells [94] and predicts poor prognosis in both wild-type and FMS-like tyrosine kinase 3 (FLT3)-mutated AML [95,96].
Additionally, it has been demonstrated that HIF-1α induces the expression of macrophage migration factor (MIF) [97], with drives interleukin-8 (IL-8) production by the BM mesenchymal stromal cells (BM-MSC), thereby supporting AML cell survival and proliferation [98]. Recently, HIF-1α has been shown to directly induce the expression of IL-8 in AML cell lines and in primary AML blasts, which in turn supports survival and proliferation of AML cells [99].
Severe hypoxia, at 1% but not at 6% O2, down-regulated FLT3 expression in normal CD34 + hematopoietic progenitors in AML primary samples independently from the mutational state of FLT3 [100]. Under such conditions, the survival mechanisms of LSCs were based on phosphoinositide 3-kinase (PI3K)/protein kinase B (PKB, also known as AKT) signaling that it could be activated through signal transducer and activator of transcription 5 (STAT5)-mediated up-regulation of the tyrosine kinase receptor AXL [101].

7. Clinical Assessment and Prognostic Value of Hypoxia

As mentioned in the above sections, the alterations induced by hypoxic BMM play a pivotal role in AML, also suggesting the potential impact of HIFs on clinical disease features. Many studies have demonstrated a prognostic role of HIF proteins and some of their targets in recent years. HIF-1α has emerged as a key regulator in tumor progression and is associated with poor prognosis in many cancers, including AML.
Immunohistochemistry analysis of primary cells isolated from 84 normal karyotype AML patients revealed that high HIF-1α levels were associated with poorer OS and EFS independently from FLT3 internal tandem duplication (ITD) or NPM1 mutations [66]. The higher expression levels of HIF-1α and their association with prognosis were confirmed by mRNA quantification when newly diagnosed AML were compared to healthy controls [102].
More recently, it has been reported that lactate dehydrogenase (LDH) levels could be used as a biomarker to assess the risk of transplantation in AML [103].
As mentioned above, severe hypoxia (1% O2) reduces the proliferative activity of primary blasts entailing a cell cycle block in the G0 phase. Thus, the quiescence state of LSCs might influence their susceptibility toward chemotherapy, which is highly dependent on the S phase for antileukemic activity. In agreement with this observation, Drolle and colleagues demonstrated that the efficacy of cytarabine arabinoside (Ara-C) was strongly reduced in AML cell lines cultured under hypoxic conditions. Moreover, they showed that hypoxic AML cells displayed higher apoptosis-inhibitory protein X-linked inhibitors of apoptosis (XIAP), together with the activation of the pro-survival PI3K pathway, demonstrating that the protective mechanism of hypoxia against Ara-C could be abrogated by PI3K inhibition [104]. Hypoxic HIF-1α stabilization decreased the in vitro sensitivity of AML cell lines to Ara-C, which was reversed by blocking O2 consumption and ROS production with the mitochondrial complex-III inhibitor antimycin [105,106]. Accordingly, it has been reported that AML chemo-resistant LSCs preferentially localize in the hypoxic endosteal regions of the mouse BM where they are protected from the pro-apoptotic effect of Ara-C [107].
Lastly, recent data suggest that HIF-1α could be involved in promyelocytic leukemia/retinoic acid receptor-α (PML-RARα)-dependent resistance. Indeed, in vitro and in vivo studies performed in acute promyelocytic leukemia (APL) models demonstrated that HIF-1α levels increased upon the treatment with all-trans retinoic acid (ATRA). In addition, the pharmacologic inhibition of HIF-1α using EZN-2968 displayed a synergistic effect on the impairment of APL engraftment and progression [108].

8. Therapeutic Opportunities: Targeting Hypoxia in AML

In light of the crucial role of hypoxia in AML development, progression and therapy response, many studies have been assessing its potential for antileukemic therapies. Two different approaches are being explored. The first one focuses on targeting downstream targets of HIFs required for survival of hypoxic LSCs; the second approach is based on hypoxia-activated prodrugs (HAPs), including some selectively activated drugs hypoxia. Therefore, in the following sections, we provide an update on these promising new therapies (Table 1).

8.1. Targeting HIF Downstream Genes

BL8040 is a new-generation CXCR4 inhibitor able to induce the mobilization of AML cells from the protective niche into the circulation and to promote AML differentiation and apoptosis through down-regulation of B-cell lymphoma 2 (BCL-2), myeloid cell leukemia 1 (MCL-1) and cyclin D1 as well as the AKT/ERK signaling pathway [109]. In addition, in vivo studies have shown that BL8040 synergizes either with BCL-2 or FLT3 inhibitors improving survival and reducing minimal residual disease in mice [109]. Furthermore, the safety and efficacy of BL8040 combined with high-dose cytarabine were assessed in a phase IIa study in relapsed and refractory AML, demonstrating a significantly extended OS [110] (ClinicalTrials.gov Identifier: NCT01838395). Moreover, the BATTLE study, a phase Ib/II multicenter ongoing study in AML patients aged 60 years or older, evaluates the safety, tolerability, and efficacy of BL8040 in combination with the programmed death-ligand 1 (PD-L1) inhibitor Atezolizumab (ClinicalTrials.gov Identifier: NCT03154827). Specifically, this combined approach aims to reduce the MRD status of AML patients further and prolong the period of remission.
The neutralizing monoclonal antibody (mAb) IMC-1C11 is a specific inhibitor of human VEGFR-2. Preclinical studies performed in AML models have shown that it can inhibit leukemic cell survival in vitro. Injection of IMC-1C11 in mice xenotransplanted with primary leukemic cells or AML cell lines significantly increased survival by inhibiting the proliferation of leukemic cells [92].

8.2. Targeting Altered Metabolism

IACS-010759 is a small-molecule inhibitor of complex I of the mitochondrial electron transport chain highly effective against OXPHOS. Preclinical studies have shown that IACS-010759 inhibits proliferation and induces apoptosis in leukemic cells, reducing tumor growth in AML mice models [111]. Based on these results, IACS-010759 was moved into phase I clinical evaluation in relapsed or refractory AML patients (ClinicalTrials.gov Identifier: NCT02882321).
Another interesting class of compounds is represented by monocarboxylate transporter (MCT) inhibitors. AR-C155858 and syrosingopine act on lactate metabolism by blocking MCT1 and MCT4 transporters, which remove the excess lactate produced by cancer cells. The lactate transporter MCT4 has been identified as a transcriptional target of HIF-1α [112] and a study from Saulle et al. demonstrated that in AML patients MCT4 over-expression is associated with poor prognosis [113]. In addition, AR-C155858 and syrosingopine exert anti-proliferative effects on AML cell lines that showed increased intracellular lactate level and enhanced the sensitivity of AML cell lines to Ara-C treatment [113].

8.3. Hypoxia-Activated Prodrugs (HAP)

The hypoxia-activated prodrugs approach is based on the ability of such molecules to function as O2 sensors so that they are activated by enzymatic reduction, specifically in hypoxic cells. Typically, a nontoxic prodrug can be activated by the enzymatic addition of one electron, which initiates the formation of DNA reactive species. This process can be inhibited by molecular oxygen. Therefore, HAPs were designed to provide targeted release of active drugs into the tumor microenvironment.
PR104 is a water-soluble phosphate ester pre-prodrug rapidly and systematically hydrolyzed to the corresponding alcohol, PR104A. PR104A, in turn, is activated in hypoxic cells resulting in hydroxylamine (PR-104H) and amine (PR-104M) metabolites, which are nitrogen mustards that act through the formation of DNA cross-links [114,115]. However, a hypoxia-independent mechanism of PR104 activation has also been reported mediated by the enzyme aldo-keto reductase 1C3 (AKR1C3), highly expressed in AML blasts [116,117]. A phase I/II study on relapsed and refractory AML reported severe gastrointestinal toxicity, myelosuppression, neutropenia, and infection [118] (ClinicalTrials.gov Identifier: NCT01037556). These observations suggest toxicity against normal hematopoiesis. Therefore, further studies will need to be performed to determine whether the combination of lower doses of PR-104 with low-dose chemotherapy could potentially maintain efficacy while reducing toxicity.
TH-302 is a 2-nitroimidazole-linked prodrug activated under hypoxia, producing the potent cytotoxin bromo-isophosphoramide mustard (Br-iPM), responsible for DNA cross-link formation. TH-302 was tested in vitro and in vivo, demonstrating hypoxia-selective activity against AML by decreasing HIF-1α expression, reducing the proliferation of leukemic cells, and enhancing double-strand breaks on DNA [119,120]. In AML xenograft models, TH-302 blocked disease progression and prolonged OS of mice [119]. Moreover, TH-302 showed synergistic antileukemic effects in FLT3-ITD AML models when combined with the tyrosine kinase inhibitor Sorafenib [120]. Based on these findings, a phase I study has been conducted in patients with relapsed or refractory AML (ClinicalTrials.gov Identifier: NCT01149915). The study confirmed that TH-302 treatment reduced the expression levels of HIF-1α. However, it showed a limited activity for the drug as a single agent [121].

9. Conclusions

Multiple studies are clarifying the role of hypoxic BMM in the pathogenesis of AML. Experimental and clinical data support the idea that the hypoxic microenvironment sustains LSC survival mainly through the activation of HIFs. Although the impact of BMM has not been well characterized in AML, it has been demonstrated that hypoxia regulates many relevant biological hallmarks of LSCs, summarized in Figure 5. The currently available literature highlights the fundamental importance of hypoxic BMM in supporting leukemic transformation and progression through the regulation of quiescence, self-renewal capacity, and cellular homing of LSCs. Moreover, the fitness of LSCs is enhanced by the metabolic shift and epigenetic modifications, and signaling alterations. Importantly, all these alterations could be crucial in resistance to therapy response, which is dependent on the biology of LSCs within the hypoxic BMM. Several mechanisms of hypoxia-mediated drug resistance have been described in AML, pointing to new attractive strategies based on targeting the alterations induced by hypoxia. Collectively, the data herein reviewed support the importance of a deeper understanding of the interactions between LSCs and the BMM as the sine-qua-non condition to design promising general strategies aimed at eradicating the ‘roots’ of AML, irrespective of the specific genetic alterations and sub-clonal complexity of this insidious form of leukemia.

Author Contributions

Conceptualization, S.B. and S.S.; writing-original draft preparation, S.B.; writing-review and editing, M.M., S.D.S. and C.M.; visualization: S.S.; supervision: S.S. and M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors acknowledge the support from AIL Bologna ODV.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AMLacute myeloid leukemia
BMMbone marrow microenvironment
LSCs leukemic stem cells
HIFhypoxia inducible factor
BMbone marrow
HSCshematopoietic stem cells
PHDprolyl hydroxylases
VHLvon Hippel-Lindau
CBPCREB-binding protein
HREhypoxia responsive elements
Tregregulatory T cells
STAT5signal transducer and activator of transcription 5
VEGFvascular endothelial growth factor
SDF1stromal cell-derived factor 1
SCFstem cell factor
Ang2angiopoietin 2
Angptlangiopoietin-likes
IGF-2insulin-like growth factor 2
IGFBPinsulin-like growth factor-binding protein
FOXOforkhead box O
CDKNcyclin-dependent kinase inhibitor
ROSreactive oxygen species
OXPHOSoxidative phosphorylation
CMLchronic myeloid leukemia
TCAtricarboxylic acid
UDP-GlcNAcuridine diphosphate N-acetylglucosamine
UDP-GalNAcuridine diphosphate N-acetylgalactosamine
2-HG2-hydroxyglutarate
PDK1pyruvate dehydrogenase 1
GSHreduced glutathione
GSSGoxidized glutathione
COX 4-1cytochrome C oxidase subunit 4 isoform 1
COX 4-2cytochrome C oxidase subunit 4 isoform 2
WHOworld health organization
OSoverall survival
EFSevent free survival
IDHisocitrate dehydrogenase
α-KGα-ketoglutarate
TETten-eleven translocation
SDHsuccinate dehydrogenase
UQubiquinone
UQH2ubiquinol
NDRG3N-Myc downstream-regulated 3
ERK1/2extracellular-signal-regulated kinase 1/2
GPR91G-protein-coupled receptor
mTORmammalian target of rapamycin
CXCL12C-X-C motif chemokine ligand 12
CXCR4C-X-C motif chemokine receptor 4
FLT3FMS-like tyrosine kinase 3
MIFmacrophage migration factor
IL-8interleukin-8
MSCmesenchymal stromal cells
PI3Kphosphoinositide 3-kinase
PKB or AKTprotein kinase B
AXLreceptor tyrosine kinase
ITDinternal tandem duplication
NPM1nucleophosmin
LDHlactate dehydrogenase
Ara-Ccytarabine arabinoside
XIAPX-linked inhibitor of apoptosis
PMLpromyelocytic leukemia
RAR- αretinoic acid receptor-α
APLacute promyelocytic leukemia
ATRAtrans retinoic acid
HAPhypoxia-activated prodrugs
BCL-2B-cell lymphoma 2
MCL-1myeloid cell lymphoma 1
PD-L1programmed death ligand 1
mAbmonoclonal antibody
MCTmonocarboxylate transporter
AKR1C3aldo-keto reductase 1C3
Br-iPMbromo-isophosphoramide mustard

References

  1. Short, N.J.; Rytting, M.E.; Cortes, J.E. Acute myeloid leukaemia. Lancet 2018, 392, 593–606. [Google Scholar] [CrossRef]
  2. Vosberg, S.; Greif, P.A. Clonal evolution of acute myeloid leukemia from diagnosis to relapse. Genes Chromosom. Cancer 2019, 58, 839–849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Morita, K.; Jahn, K.; Hu, T.; Tanaka, T.; Sasaki, Y.; Kuipers, J.; Loghavi, S.; Wang, S.A.; Yan, Y.; Furudate, K.; et al. Clonal evolution of acute myeloid leukemia revealed by high-throughput single-cell genomics. Nat. Commun. 2020, 11, 5327. [Google Scholar] [CrossRef] [PubMed]
  4. Döhner, H.; Estey, D.; Grimwade, D.; Amadori, S.; Appelbaum, F.R.; Büchner, T.; Dombret, H.; Ebert, B.L.; Fenaux, P.; Larson, R.A.; et al. 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]
  5. Shafer, D.; Grant, S. Update on rational targeted therapy in AML. Blood Rev. 2016, 30, 275–283. [Google Scholar] [CrossRef] [Green Version]
  6. Kantarjian, H.; Kadia, T.; DiNardo, C.; Daver, N.; Borthakur, G.; Jabbour, E.; Garcia-Manero, G.; Konopleva, M.; Ravandi, F. Acute myeloid leukemia: Current progress and future directions. Blood Cancer J. 2021, 11, 41. [Google Scholar] [CrossRef]
  7. Löwenberg, B.; Ossenkoppele, G.J.; Van Putten, W.; Schouten, H.C.; Graux, C.; Ferrant, A.; Sonneveld, P.; Maertens, J.; Jongen-Lavrencic, M.; von Lilienfeld-Toal, M.; et al. High-Dose Daunorubicin in Older Patients with Acute Myeloid Leukemia. N. Engl. J. Med. 2009, 361, 1235–1248. [Google Scholar] [CrossRef] [PubMed]
  8. Fernandez, H.F.; Sun, Z.; Yao, X.; Litzow, M.R.; Luger, S.M.; Paietta, E.M.; Racevskis, J.; Dewald, G.W.; Ketterling, R.P.; Bennett, J.M.; et al. Anthracycline dose intensification in acute myeloid leukemia. N. Engl. J. Med. 2009, 361, 1249–1259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Ding, L.; Ley, T.J.; Larson, D.E.; Miller, C.A.; Koboldt, D.C.; Welch, J.S.; Ritchey, J.K.; Young, M.A.; Lamprecht, T.; McLellan, M.D.; et al. Clonal evolution in relapsed acute myeloid leukaemia revealed by whole-genome sequencing. Nature 2012, 481, 506–510. [Google Scholar] [CrossRef] [PubMed]
  10. Levin, M.; Stark, M.; Ofran, Y.; Assaraf, Y.G. Deciphering molecular mechanisms underlying chemoresistance in relapsed AML patients: Towards precision medicine overcoming drug resistance. Cancer Cell Int. 2021, 21, 53. [Google Scholar] [CrossRef]
  11. Calvi, L.M.; Adams, G.B.; Weibrecht, K.W.; Weber, J.M.; Olson, D.P.; Knight, M.C.; Martin, R.P.; Schipani, E.; Divieti, P.; Bringhurst, F.R.; et al. Osteoblastic cells regulate the haematopoietic stem cell niche. Nature 2003, 425, 841–846. [Google Scholar] [CrossRef] [PubMed]
  12. Kiel, M.J.; Yilmaz, O.H.; Iwashita, T.; Yilmaz, O.H.; Terhorst, C.; Morrison, S.J. SLAM family receptors distinguish hematopoietic stem and progenitor cells and reveal endothelial niches for stem cells. Cell 2005, 121, 1109–1121. [Google Scholar] [CrossRef] [Green Version]
  13. Ding, L.; Saunders, T.L.; Enikolopov, G.; Morrison, S.J. Endothelial and perivascular cells maintain haematopoietic stem cells. Nature 2012, 481, 457–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhang, J.; Niu, C.; Ye, L.; Huang, H.; He, X.; Tong, W.; Ross, J.; Haug, J.; Johnson, T.; Feng, J.Q.; et al. Identification of the haematopoietic stem cell niche and control of the niche size. Nature 2003, 425, 836–841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lévesque, J.-P.; Helwani, F.M.; Winkler, I.G. The endosteal ‘osteoblastic’ niche and its role in hematopoietic stem cell homing and mobilization. Leukemia 2010, 24, 1979–1992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sugimura, R. The significance and application of vascular niche in the development and maintenance of hematopoietic stem cells. Int. J. Hematol. 2018, 107, 642–645. [Google Scholar] [CrossRef] [Green Version]
  17. Eliasson, P.; Jönsson, J.-I. The hematopoietic stem cell niche: Low in oxygen but a nice place to be. J. Cell. Physiol. 2010, 222, 17–22. [Google Scholar] [CrossRef]
  18. El-Showk, S. Low-oxygen landscape mapped in bone marrow. Nat. Middle East 2014, 508, 269–273. [Google Scholar] [CrossRef]
  19. Spencer, J.A.; Ferraro, F.; Roussakis, E.; Klein, A.; Wu, J.; Runnels, J.M.; Zaher, W.; Mortensen, L.J.; Alt, C.; Turcotte, R.; et al. Direct measurement of local oxygen concentration in the bone marrow of live animals. Nature 2014, 508, 269–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Nombela-Arrieta, C.; Silberstein, L.E. The science behind the hypoxic niche of hematopoietic stem and progenitors. Hematol. Am. Soc. Hematol. Educ. Progr. 2014, 2014, 542–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Najafi, M.; Farhood, B.; Mortezaee, K.; Kharazinejad, E.; Majidpoor, J.; Ahadi, R. Hypoxia in solid tumors: A key promoter of cancer stem cell (CSC) resistance. J. Cancer Res. Clin. Oncol. 2020, 146, 19–31. [Google Scholar] [CrossRef]
  22. Kubota, Y.; Takubo, K.; Suda, T. Bone marrow long label-retaining cells reside in the sinusoidal hypoxic niche. Biochem. Biophys. Res. Commun. 2008, 366, 335–339. [Google Scholar] [CrossRef] [PubMed]
  23. Lo Celso, C.; Fleming, H.E.; Wu, J.W.; Zhao, C.X.; Miake-Lye, S.; Fujisaki, J.; Côté, D.; Rowe, D.W.; Lin, C.P.; Scadden, D.T. Live-animal tracking of individual haematopoietic stem/progenitor cells in their niche. Nature 2009, 457, 92–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Parmar, K.; Mauch, P.; Vergilio, J.-A.; Sackstein, R.; Down, J.D. Distribution of hematopoietic stem cells in the bone marrow according to regional hypoxia. Proc. Natl. Acad. Sci. USA 2007, 104, 5431–5436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zheng, J.; Umikawa, M.; Zhang, S.; Huynh, H.; Silvany, R.; Chen, B.P.C.; Chen, L.; Zhang, C.C. Ex vivo expanded hematopoietic stem cells overcome the MHC barrier in allogeneic transplantation. Cell Stem Cell 2011, 9, 119–130. [Google Scholar] [CrossRef] [Green Version]
  26. Shima, H.; Takubo, K.; Iwasaki, H.; Yoshihara, H.; Gomei, Y.; Hosokawa, K.; Arai, F.; Takahashi, T.; Suda, T. Reconstitution activity of hypoxic cultured human cord blood CD34-positive cells in NOG mice. Biochem. Biophys. Res. Commun. 2009, 378, 467–472. [Google Scholar] [CrossRef]
  27. Shima, H.; Takubo, K.; Tago, N.; Iwasaki, H.; Arai, F.; Takahashi, T.; Suda, T. Acquisition of G0 state by CD34-positive cord blood cells after bone marrow transplantation. Exp. Hematol. 2010, 38, 1231–1240. [Google Scholar] [CrossRef]
  28. Moirangthem, R.D.; Singh, S.; Adsul, A.; Jalnapurkar, S.; Limaye, L.; Kale, V.P. Hypoxic Niche-Mediated Regeneration of Hematopoiesis in the Engraftment Window Is Dominantly Affected by Oxygen Tension in the Milieu. Stem Cells Dev. 2015, 24, 2423–2436. [Google Scholar] [CrossRef] [Green Version]
  29. Bapat, A.; Schippel, N.; Shi, X.; Jasbi, P.; Gu, H.; Kala, M.; Sertil, A.; Sharma, S. Hypoxia promotes erythroid differentiation through the development of progenitors and proerythroblasts. Exp. Hematol. 2021, 97, 32–46.e35. [Google Scholar] [CrossRef]
  30. Mostafa, S.S.; Miller, W.M.; Papoutsakis, E.T. Oxygen tension influences the differentiation, maturation and apoptosis of human megakaryocytes. Br. J. Haematol. 2000, 111, 879–889. [Google Scholar]
  31. Chen, S.; Su, Y.; Wang, J. ROS-mediated platelet generation: A microenvironment-dependent manner for megakaryocyte proliferation, differentiation, and maturation. Cell Death Dis. 2013, 4, e722. [Google Scholar] [CrossRef] [Green Version]
  32. Taneja, R.; Rameshwar, P.; Upperman, J.; Wang, M.T.; Livingston, D.H. Effects of hypoxia on granulocytic-monocytic progenitors in rats. Role of bone marrow stroma. Am. J. Hematol. 2000, 64, 20–25. [Google Scholar] [CrossRef]
  33. Zhou, H.-S.; Carter, B.Z.; Andreeff, M. Bone marrow niche-mediated survival of leukemia stem cells in acute myeloid leukemia: Yin and Yang. Cancer Biol. Med. 2016, 13, 248–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wood, S.M.; Gleadle, J.M.; Pugh, C.W.; Hankinson, O.; Ratcliffe, P.J. The role of the aryl hydrocarbon receptor nuclear translocator (ARNT) in hypoxic induction of gene expression. Studies in ARNT-deficient cells. J. Biol. Chem. 1996, 271, 15117–15123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ke, Q.; Costa, M. Hypoxia-inducible factor-1 (HIF-1). Mol. Pharmacol. 2006, 70, 1469–1480. [Google Scholar] [CrossRef] [PubMed]
  36. Lee, J.-W.; Bae, S.-H.; Jeong, J.-W.; Kim, S.-H.; Kim, K.-W. Hypoxia-inducible factor (HIF-1)alpha: Its protein stability and biological functions. Exp. Mol. Med. 2004, 36, 1–12. [Google Scholar] [CrossRef]
  37. Mazure, N.M.; Pouysségur, J. Hypoxia-induced autophagy: Cell death or cell survival? Curr. Opin. Cell Biol. 2010, 22, 177–180. [Google Scholar] [CrossRef]
  38. Winkler, I.G.; Barbier, V.; Wadley, R.; Zannettino, A.C.W.; Williams, S.; Lévesque, J. Positioning of bone marrow hematopoietic and stromal cells relative to blood flow in vivo: Serially reconstituting hematopoietic stem cells reside in distinct nonperfused niches. Blood 2010, 116, 375–385. [Google Scholar] [CrossRef]
  39. Acar, M.; Kocherlakota, K.S.; Murphy, M.M.; Peyer, J.G.; Oguro, H.; Inra, C.N.; Jaiyeola, C.; Zhao, Z.; Luby-Phelps, K.; Morrison, S.J. Deep imaging of bone marrow shows non-dividing stem cells are mainly perisinusoidal. Nature 2015, 526, 126–130. [Google Scholar] [CrossRef] [Green Version]
  40. Kunisaki, Y.; Bruns, I.; Scheiermann, C.; Ahmed, J.; Pinho, S.; Zhang, D.; Mizoguchi, T.; Wei, Q.; Lucas, D.; Ito, K.; et al. Arteriolar niches maintain haematopoietic stem cell quiescence. Nature 2013, 502, 637–643. [Google Scholar] [CrossRef] [Green Version]
  41. Xie, Y.; Yin, T.; Wiegraebe, W.; He, X.C.; Miller, D.; Stark, D.; Perko, K.; Alexander, R.; Schwartz, J.; Grindley, J.C.; et al. Detection of functional haematopoietic stem cell niche using real-time imaging. Nature 2009, 457, 97–101. [Google Scholar] [CrossRef] [PubMed]
  42. Guezguez, B.; Campbell, C.J.V.; Boyd, A.L.; Karanu, F.; Casado, F.L.; Di Cresce, C.; Collins, T.J.; Shapovalova, Z.; Xenocostas, A.; Bhatia, M. Regional Localization within the Bone Marrow Influences the Functional Capacity of Human HSCs. Cell Stem Cell 2013, 13, 175–189. [Google Scholar] [CrossRef] [Green Version]
  43. Fujisaki, J.; Wu, J.; Carlson, A.L.; Silberstein, L.; Putheti, P.; Larocca, R.; Gao, W.; Saito, T.I.; Lo Celso, C.; Tsuyuzaki, H.; et al. In vivo imaging of Treg cells providing immune privilege to the haematopoietic stem-cell niche. Nature 2011, 474, 216–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Shehade, H.; Acolty, V.; Moser, M.; Oldenhove, G. Cutting Edge: Hypoxia-Inducible Factor 1 Negatively Regulates Th1 Function. J. Immunol. 2015, 195, 1372–1376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Simsek, T.; Kocabas, F.; Zheng, J.; DeBerardinis, R.J.; Mahmoud, A.I.; Olson, E.N.; Schneider, J.W.; Zhang, C.C.; Sadek, H.A. The distinct metabolic profile of hematopoietic stem cells reflects their location in a hypoxic niche. Cell Stem Cell 2010, 7, 380–390. [Google Scholar] [CrossRef] [Green Version]
  46. Fatrai, S.; Wierenga, A.T.J.; Daenen, S.M.G.J.; Vellenga, E.; Schuringa, J.J. Identification of HIF2α as an important STAT5 target gene in human hematopoietic stem cells. Blood 2011, 117, 3320–3330. [Google Scholar] [CrossRef]
  47. Takubo, K.; Goda, N.; Yamada, W.; Iriuchishima, H.; Ikeda, E.; Kubota, Y.; Shima, H.; Johnson, R.; Hirao, A.; Suematsu, M.; et al. Regulation of the HIF-1alpha level is essential for hematopoietic stem cells. Cell Stem Cell 2010, 7, 391–402. [Google Scholar] [CrossRef] [Green Version]
  48. Forristal, C.E.; Nowlan, B.; Jacobsen, R.; Barbier, V.; Walkinshaw, G.; Walkley, C.; Winkler, I.; Levesque, J.P. HIF-1α is required for hematopoietic stem cell mobilization and 4-prolyl hydroxylase inhibitors enhance mobilization by stabilizing HIF-1α. Leukemia 2015, 29, 1366–1378. [Google Scholar] [CrossRef] [Green Version]
  49. Nombela-Arrieta, C.; Pivarnik, G.; Winkel, B.; Canty, K.J.; Harley, B.; Mahoney, J.E.; Park, S.-Y.; Lu, J.; Protopopov, A.; Silberstein, L.E. Quantitative imaging of haematopoietic stem and progenitor cell localization and hypoxic status in the bone marrow microenvironment. Nat. Cell Biol. 2013, 15, 533–543. [Google Scholar] [CrossRef]
  50. Eliasson, P.; Rehn, M.; Hammar, P.; Larsson, P.; Sirenko, O.; Flippin, L.A.; Cammenga, J.; Jönsson, J.-I. Hypoxia mediates low cell-cycle activity and increases the proportion of long-term–reconstituting hematopoietic stem cells during in vitro culture. Exp. Hematol. 2010, 38, 301–310.e2. [Google Scholar] [CrossRef] [Green Version]
  51. Forristal, C.E.; Winkler, I.G.; Nowlan, B.; Barbier, V.; Walkinshaw, G.; Levesque, J.-P. Pharmacologic stabilization of HIF-1α increases hematopoietic stem cell quiescence in vivo and accelerates blood recovery after severe irradiation. Blood 2013, 121, 759–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Krock, B.L.; Eisinger-Mathason, T.S.; Giannoukos, D.N.; Shay, J.E.; Gohil, M.; Lee, D.S.; Nakazawa, M.S.; Sesen, J.; Skuli, N.; Simon, M.C. The aryl hydrocarbon receptor nuclear translocator is an essential regulator of murine hematopoietic stem cell viability. Blood 2015, 125, 3263–3272. [Google Scholar] [CrossRef] [Green Version]
  53. Zhang, H.; Wong, C.C.L.; Wei, H.; Gilkes, D.M.; Korangath, P.; Chaturvedi, P.; Schito, L.; Chen, J.; Krishnamachary, B.; Winnard, P.T.; et al. HIF-1-dependent expression of angiopoietin-like 4 and L1CAM mediates vascular metastasis of hypoxic breast cancer cells to the lungs. Oncogene 2012, 31, 1757–1770. [Google Scholar] [CrossRef]
  54. Kocabas, F.; Xie, L.; Xie, J.; Yu, Z.; DeBerardinis, R.J.; Kimura, W.; Thet, S.; Elshamy, A.F.; Abouellail, H.; Muralidhar, S.; et al. Hypoxic metabolism in human hematopoietic stem cells. Cell Biosci. 2015, 5, 39. [Google Scholar] [CrossRef] [Green Version]
  55. Miharada, K.; Karlsson, G.; Rehn, M.; Rörby, E.; Siva, K.; Cammenga, J.; Karlsson, S. Cripto Regulates Hematopoietic Stem Cells as a Hypoxic-Niche-Related Factor through Cell Surface Receptor GRP78. Cell Stem Cell 2011, 9, 330–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Suda, T.; Takubo, K.; Semenza, G.L. Metabolic Regulation of Hematopoietic Stem Cells in the Hypoxic Niche. Cell Stem Cell 2011, 9, 298–310. [Google Scholar] [CrossRef] [Green Version]
  57. Takubo, K.; Nagamatsu, G.; Kobayashi, C.I.; Nakamura-Ishizu, A.; Kobayashi, H.; Ikeda, E.; Goda, N.; Rahimi, Y.; Johnson, R.S.; Soga, T.; et al. Regulation of Glycolysis by Pdk Functions as a Metabolic Checkpoint for Cell Cycle Quiescence in Hematopoietic Stem Cells. Cell Stem Cell 2013, 12, 49–61. [Google Scholar] [CrossRef] [Green Version]
  58. Kocabas, F.; Zheng, J.; Thet, S.; Copeland, N.G.; Jenkins, N.A.; DeBerardinis, R.J.; Zhang, C.; Sadek, H.A. Meis1 regulates the metabolic phenotype and oxidant defense of hematopoietic stem cells. Blood 2012, 120, 4963–4972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Guitart, A.V.; Subramani, C.; Armesilla-Diaz, A.; Smith, G.; Sepulveda, C.; Gezer, D.; Vukovic, M.; Dunn, K.; Pollard, P.; Holyoake, T.L.; et al. Hif-2α is not essential for cell-autonomous hematopoietic stem cell maintenance. Blood 2013, 122, 1741–1745. [Google Scholar] [CrossRef]
  60. Rouault-Pierre, K.; Lopez-Onieva, L.; Foster, K.; Anjos-Afonso, F.; Lamrissi-Garcia, I.; Serrano-Sanchez, M.; Mitter, R.; Ivanovic, Z.; De Verneuil, H.; Gribben, J.; et al. HIF-2α Protects Human Hematopoietic Stem/Progenitors and Acute Myeloid Leukemic Cells from Apoptosis Induced by Endoplasmic Reticulum Stress. Cell Stem Cell 2013, 13, 549–563. [Google Scholar] [CrossRef] [Green Version]
  61. Arai, F.; Hirao, A.; Ohmura, M.; Sato, H.; Matsuoka, S.; Takubo, K.; Ito, K.; Koh, G.Y.; Suda, T. Tie2/angiopoietin-1 signaling regulates hematopoietic stem cell quiescence in the bone marrow niche. Cell 2004, 118, 149–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Nilsson, S.K.; Johnston, H.M.; Whitty, G.A.; Williams, B.; Webb, R.J.; Denhardt, D.T.; Bertoncello, I.; Bendall, L.J.; Simmons, P.J.; Haylock, D.N. Osteopontin, a key component of the hematopoietic stem cell niche and regulator of primitive hematopoietic progenitor cells. Blood 2005, 106, 1232–1239. [Google Scholar] [CrossRef] [PubMed]
  63. Naveiras, O.; Daley, G.Q. Stem cells and their niche: A matter of fate. Cell. Mol. Life Sci. 2006, 63, 760–766. [Google Scholar] [CrossRef] [PubMed]
  64. Colmone, A.; Amorim, M.; Pontier, A.L.; Wang, S.; Jablonski, E.; Sipkins, D.A. Leukemic cells create bone marrow niches that disrupt the behavior of normal hematopoietic progenitor cells. Science 2008, 322, 1861–1865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Boyd, A.L.; Campbell, C.J.V.; Hopkins, C.I.; Fiebig-Comyn, A.; Russell, J.; Ulemek, J.; Foley, R.; Leber, B.; Xenocostas, A.; Collins, T.J.; et al. Niche displacement of human leukemic stem cells uniquely allows their competitive replacement with healthy HSPCs. J. Exp. Med. 2014, 211, 1925–1935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Deeb, G.; Vaughan, M.M.; McInnis, I.; Ford, L.; Ann Sait, S.N.J.; Starostik, P.; Wetzler, M.; Mashtare, T.; Wang, E.S. Hypoxia-inducible factor-1α protein expression is associated with poor survival in normal karyotype adult acute myeloid leukemia. Leuk. Res. 2011, 35, 579–584. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, Y.; Liu, Y.; Malek, S.N.; Zheng, P.; Liu, Y. Targeting HIF1α eliminates cancer stem cells in hematological malignancies. Cell Stem Cell 2011, 8, 399–411. [Google Scholar] [CrossRef] [Green Version]
  68. Vukovic, M.; Guitart, A.V.; Sepulveda, C.; Villacreces, A.; O’Duibhir, E.; Panagopoulou, T.I.; Ivens, A.; Menendez-Gonzalez, J.; Iglesias, J.M.; Allen, L.; et al. Hif-1α and Hif-2α synergize to suppress AML development but are dispensable for disease maintenance. J. Exp. Med. 2015, 212, 2223–2234. [Google Scholar] [CrossRef]
  69. Herst, P.M.; Howman, R.A.; Neeson, P.J.; Berridge, M.V.; Ritchie, D.S. The level of glycolytic metabolism in acute myeloid leukemia blasts at diagnosis is prognostic for clinical outcome. J. Leukoc. Biol. 2011, 89, 51–55. [Google Scholar] [CrossRef]
  70. Lodi, A.; Tiziani, S.; Khanim, F.L.; Drayson, M.T.; Günther, U.L.; Bunce, C.M.; Viant, M.R. Hypoxia triggers major metabolic changes in AML cells without altering indomethacin-induced TCA cycle deregulation. ACS Chem. Biol. 2011, 6, 169–175. [Google Scholar] [CrossRef]
  71. Martinez, M.R.; Dias, T.B.; Natov, P.S.; Zachara, N.E. Stress-induced O-GlcNAcylation: An adaptive process of injured cells. Biochem. Soc. Trans. 2017, 45, 237–249. [Google Scholar] [CrossRef] [PubMed]
  72. Bruno, S.; Pazzaglia, M.; Cerchione, C.; Soverini, S.; Cavo, M.; Montanaro, L.; Simonetti, G.; Martinelli, G. Abstract 2651: Deep hypoxia and the genomic background cooperate to shape the metabolic profile of acute myeloid leukemia cells. In Proceedings of the AACR Annual Meeting 2019, Atlanta, GA, USA, 29 March–3 April 2019. [Google Scholar] [CrossRef]
  73. Goto, M.; Miwa, H.; Suganuma, K.; Tsunekawa-Imai, N.; Shikami, M.; Mizutani, M.; Mizuno, S.; Hanamura, I.; Nitta, M. Adaptation of leukemia cells to hypoxic condition through switching the energy metabolism or avoiding the oxidative stress. BMC Cancer 2014, 14, 76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Cunningham, I.; Kohno, B. 18 FDG-PET/CT: 21st century approach to leukemic tumors in 124 cases. Am. J. Hematol. 2016, 91, 379–384. [Google Scholar] [CrossRef]
  75. Chen, W.-L.; Wang, J.; Zhao, A.; Xu, X.; Wang, Y.; Chen, T.; Li, J.; Mi, J.; Zhu, Y.; Liu, Y.; et al. A distinct glucose metabolism signature of acute myeloid leukemia with prognostic value. Blood 2014, 124, 1645–1654. [Google Scholar] [CrossRef] [PubMed]
  76. Ward, P.S.; Patel, J.; Wise, D.R.; Abdel-Wahab, O.; Bennett, B.D.; Coller, H.A.; Cross, J.R.; Fantin, V.R.; Hedvat, C.V.; Perl, A.E.; et al. The Common Feature of Leukemia-Associated IDH1 and IDH2 Mutations Is a Neomorphic Enzyme Activity Converting α-Ketoglutarate to 2-Hydroxyglutarate. Cancer Cell 2010, 17, 225–234. [Google Scholar] [CrossRef] [Green Version]
  77. Bledea, R.; Vasudevaraja, V.; Patel, S.; Stafford, J.; Serrano, J.; Esposito, G.; Tredwin, L.M.; Goodman, N.; Kloetgen, A.; Golfinos, J.G.; et al. Functional and topographic effects on DNA methylation in IDH1/2 mutant cancers. Sci. Rep. 2019, 9, 16830. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, Y.-P.; Li, J.-T.; Qu, J.; Yin, M.; Lei, Q.-Y. Metabolite sensing and signaling in cancer. J. Biol. Chem. 2020, 295, 11938–11946. [Google Scholar] [CrossRef] [PubMed]
  79. Waitkus, M.S.; Diplas, B.H.; Yan, H. Biological Role and Therapeutic Potential of IDH Mutations in Cancer. Cancer Cell 2018, 34, 186–195. [Google Scholar] [CrossRef] [Green Version]
  80. Medeiros, B.C.; Fathi, A.T.; DiNardo, C.D.; Pollyea, D.A.; Chan, S.M.; Swords, R. Isocitrate dehydrogenase mutations in myeloid malignancies. Leukemia 2017, 31, 272–281. [Google Scholar] [CrossRef] [PubMed]
  81. Intlekofer, A.M.; Dematteo, R.G.; Venneti, S.; Finley, L.W.S.; Lu, C.; Judkins, A.R.; Rustenburg, A.S.; Grinaway, P.B.; Chodera, J.D.; Cross, J.R.; et al. Hypoxia Induces Production of L-2-Hydroxyglutarate. Cell Metab. 2015, 22, 304–311. [Google Scholar] [CrossRef] [Green Version]
  82. Laukka, T.; Mariani, C.J.; Ihantola, T.; Cao, J.Z.; Hokkanen, J.; Kaelin, W.G.; Godley, L.A.; Koivunen, P. Fumarate and Succinate Regulate Expression of Hypoxia-inducible Genes via TET Enzymes. J. Biol. Chem. 2016, 291, 4256–4265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Hewitson, K.S.; Liénard, B.M.R.; McDonough, M.A.; Clifton, I.J.; Butler, D.; Soares, A.S.; Oldham, N.J.; McNeill, L.A.; Schofield, C.J. Structural and Mechanistic Studies on the Inhibition of the Hypoxia-inducible Transcription Factor Hydroxylases by Tricarboxylic Acid Cycle Intermediates. J. Biol. Chem. 2007, 282, 3293–3301. [Google Scholar] [CrossRef] [Green Version]
  84. Selak, M.A.; Armour, S.M.; MacKenzie, E.D.; Boulahbel, H.; Watson, D.G.; Mansfield, K.D.; Pan, Y.; Simon, M.C.; Thompson, C.B.; Gottlieb, E. Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-α prolyl hydroxylase. Cancer Cell 2005, 7, 77–85. [Google Scholar] [CrossRef] [Green Version]
  85. Lee, D.C.; Sohn, H.A.; Park, Z.; Oh, S.; Kang, Y.K.; Lee, K.; Kang, M.; Jang, Y.J.; Yang, S.; Hong, Y.K.; et al. A Lactate-Induced Response to Hypoxia. Cell 2015, 161, 595–609. [Google Scholar] [CrossRef] [Green Version]
  86. Park, K.C.; Lee, D.C.; Yeom, Y.I. NDRG3-mediated lactate signaling in hypoxia. BMB Rep. 2015, 48, 301–302. [Google Scholar] [CrossRef] [Green Version]
  87. He, W.; Miao, F.J.; Lin, D.C.; Schwandner, R.T.; Wang, Z.; Gao, J.; Chen, J.; Tian, H.; Ling, L. Citric acid cycle intermediates as ligands for orphan G-protein-coupled receptors. Nature 2004, 429, 188–193. [Google Scholar] [CrossRef]
  88. Hakak, Y.; Lehmann-Bruinsma, K.; Phillips, S.; Le, T.; Liaw, C.; Connolly, D.T.; Behan, D.P. The role of the GPR91 ligand succinate in hematopoiesis. J. Leukoc. Biol. 2009, 85, 837–843. [Google Scholar] [CrossRef]
  89. Fu, X.; Chin, R.M.; Vergnes, L.; Hwang, H.; Deng, G.; Xing, Y.; Pai, M.Y.; Li, S.; Ta, L.; Fazlollahi, F.; et al. 2-Hydroxyglutarate Inhibits ATP Synthase and mTOR Signaling. Cell Metab. 2015, 22, 508–515. [Google Scholar] [CrossRef] [Green Version]
  90. Padró, T.; Ruiz, S.; Bieker, R.; Bürger, H.; Steins, M.; Kienast, J.; Büchner, T.; Berdel, W.E.; Mesters, R.M. Increased angiogenesis in the bone marrow of patients with acute myeloid leukemia. Blood 2000, 95, 2637–2644. [Google Scholar] [CrossRef] [PubMed]
  91. Hussong, J.W.; Rodgers, G.M.; Shami, P.J. Evidence of increased angiogenesis in patients with acute myeloid leukemia. Blood 2000, 95, 309–313. [Google Scholar] [CrossRef] [PubMed]
  92. Dias, S.; Hattori, K.; Zhu, Z.; Heissig, B.; Choy, M.; Lane, W.; Wu, Y.; Chadburn, A.; Hyjek, E.; Gill, M.; et al. Autocrine stimulation of VEGFR-2 activates human leukemic cell growth and migration. J. Clin. Investig. 2000, 106, 511–521. [Google Scholar] [CrossRef]
  93. Fiegl, M.; Samudio, I.; Clise-Dwyer, K.; Burks, J.K.; Mnjoyan, Z.; Andreeff, M. CXCR4 expression and biologic activity in acute myeloid leukemia are dependent on oxygen partial pressure. Blood 2009, 113, 1504–1512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Sison, E.A.R.; McIntyre, E.; Magoon, D.; Brown, P. Dynamic Chemotherapy-Induced Upregulation of CXCR4 Expression: A Mechanism of Therapeutic Resistance in Pediatric AML. Mol. Cancer Res. 2013, 11, 1004–1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Konoplev, S.; Rassidakis, G.Z.; Estey, E.; Kantarjian, H.; Liakou, C.I.; Huang, X.; Xiao, L.; Andreeff, M.; Konopleva, M.; Medeiros, L.J. Overexpression of CXCR4 predicts adverse overall and event-free survival in patients with unmutated FLT3 acute myeloid leukemia with normal karyotype. Cancer 2007, 109, 1152–1156. [Google Scholar] [CrossRef] [PubMed]
  96. Rombouts, E.J.C.; Pavic, B.; Löwenberg, B.; Ploemacher, R.E. Relation between CXCR-4 expression, FLT3 mutations, and unfavorable prognosis of adult acute myeloid leukemia. Blood 2004, 104, 550–557. [Google Scholar] [CrossRef]
  97. Abdul-Aziz, A.M.; Shafat, M.S.; Sun, Y.; Marlein, C.R.; Piddock, R.E.; Robinson, S.D.; Edwards, D.R.; Zhou, Z.; Collins, A.; Bowles, K.M.; et al. HIF1α drives chemokine factor pro-tumoral signaling pathways in acute myeloid leukemia. Oncogene 2018, 37, 2676–2686. [Google Scholar] [CrossRef]
  98. Abdul-Aziz, A.M.; Shafat, M.S.; Sun, Y.; Marlein, C.R.; Piddock, R.E.; Robinson, S.D.; Edwards, D.R.; Zhou, Z.; Collins, A.; Bowles, K.M.; et al. MIF-Induced Stromal PKCβ/IL8 Is Essential in Human Acute Myeloid Leukemia. Cancer Res. 2017, 77, 303–311. [Google Scholar] [CrossRef] [Green Version]
  99. Kuett, A.; Rieger, C.; Perathoner, D.; Herold, T.; Wagner, M.; Sironi, S.; Sotlar, K.; Horny, H.; Deniffel, C.; Drolle, H.; et al. IL-8 as mediator in the microenvironment-leukaemia network in acute myeloid leukaemia. Sci. Rep. 2015, 5, 18411. [Google Scholar] [CrossRef]
  100. Sironi, S.; Wagner, M.; Kuett, A.; Drolle, H.; Polzer, H.; Spiekermann, K.; Rieger, C.; Fiegl, M. Microenvironmental hypoxia regulates FLT3 expression and biology in AML. Sci. Rep. 2015, 5, 17550. [Google Scholar] [CrossRef]
  101. Dumas, P.-Y.; Naudin, C.; Martin-Lannerée, S.; Izac, B.; Casetti, L.; Mansier, O.; Rousseau, B.; Artus, A.; Dufossée, M.; Giese, A.; et al. Hematopoietic niche drives FLT3-ITD acute myeloid leukemia resistance to quizartinib via STAT5-and hypoxia-dependent upregulation of AXL. Haematologica 2019, 104, 2017–2027. [Google Scholar] [CrossRef] [Green Version]
  102. Elhoseiny, S.M. Hypoxia-Inducible Factor 1 Alpha (HIF-1α) and Its Prognostic Value in Acute Myeloid Leukemia. Hematol. Transfus. Int. J. 2017, 4, 19–25. [Google Scholar] [CrossRef]
  103. Geva, M.; Shouval, R.; Fein, J.A.; Danylesko, I.; Shem-Tov, N.; Yerushalmi, R.; Shimoni, A.; Nagler, A. Lactate Dehydrogenase Is a Key Prognostic Factor in Acute Myeloid Leukemia and Lymphoma Patients Undergoing Allogeneic Hematopoietic Stem Cell Transplantation. Blood 2019, 134, 3304. [Google Scholar] [CrossRef]
  104. Drolle, H.; Wagner, M.; Vasold, J.; Kütt, A.; Deniffel, C.; Sotlar, K.; Sironi, S.; Herold, T.; Rieger, C.; Fiegl, M. Hypoxia regulates proliferation of acute myeloid leukemia and sensitivity against chemotherapy. Leuk. Res. 2015, 39, 779–785. [Google Scholar] [CrossRef]
  105. Matsunaga, T.; Imataki, O.; Torii, E.; Kameda, T.; Shide, K.; Shimoda, H.; Kamiunten, A.; Sekine, M.; Taniguchi, Y.; Yamamoto, S.; et al. Elevated HIF-1α expression of acute myelogenous leukemia stem cells in the endosteal hypoxic zone may be a cause of minimal residual disease in bone marrow after chemotherapy. Leuk. Res. 2012, 36, e122–e124. [Google Scholar] [CrossRef] [PubMed]
  106. Chua, Y.L.; Dufour, E.; Dassa, E.P.; Rustin, P.; Jacobs, H.T.; Taylor, C.T.; Hagen, T. Stabilization of Hypoxia-inducible Factor-1α Protein in Hypoxia Occurs Independently of Mitochondrial Reactive Oxygen Species Production. J. Biol. Chem. 2010, 285, 31277–31284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Ishikawa, F.; Yoshida, S.; Saito, Y.; Hijikata, A.; Kitamura, H.; Tanaka, S.; Nakamura, R.; Tanaka, T.; Tomiyama, H.; Saito, N.; et al. Chemotherapy-resistant human AML stem cells home to and engraft within the bone-marrow endosteal region. Nat. Biotechnol. 2007, 25, 1315–1321. [Google Scholar] [CrossRef]
  108. Coltella, N.; Percio, S.; Valsecchi, R.; Cuttano, R.; Guarnerio, J.; Ponzoni, M.; Pandolfi, P.P.; Melillo, G.; Pattini, L.; Bernardi, R. HIF factors cooperate with PML-RARα to promote acute promyelocytic leukemia progression and relapse. EMBO Mol. Med. 2014, 6, 640–650. [Google Scholar] [CrossRef]
  109. Abraham, M.; Klein, S.; Bulvik, B.; Wald, H.; Weiss, I.D.; Olam, D.; Weiss, L.; Beider, K.; Eizenberg, O.; Wald, O.; et al. The CXCR4 inhibitor BL-8040 induces the apoptosis of AML blasts by downregulating ERK, BCL-2, MCL-1 and cyclin-D1 via altered miR-15a/16-1 expression. Leukemia 2017, 31, 2336–2346. [Google Scholar] [CrossRef]
  110. Borthakur, G.; Ofran, Y.; Tallman, M.S.; Foran, J.; Uy, G.L.; DiPersio, J.F.; Showel, M.M.; Shimoni, A.; Nagler, A.; Rowe, J.M.; et al. BL-8040 CXCR4 antagonist is safe and demonstrates antileukemic activity in combination with cytarabine for the treatment of relapsed/refractory acute myelogenous leukemia: An open-label safety and efficacy phase 2a study. Cancer 2021, 127, 1246–1259. [Google Scholar] [CrossRef]
  111. Molina, J.R.; Sun, Y.; Protopopova, M.; Gera, S.; Bandi, M.; Bristow, C.; McAfoos, T.; Morlacchi, P.; Ackroyd, J.; Agip, A.A.; et al. An inhibitor of oxidative phosphorylation exploits cancer vulnerability. Nat. Med. 2018, 24, 1036–1046. [Google Scholar] [CrossRef] [Green Version]
  112. Ullah, M.S.; Davies, A.J.; Halestrap, A.P. The Plasma Membrane Lactate Transporter MCT4, but Not MCT1, Is Up-regulated by Hypoxia through a HIF-1α-dependent Mechanism. J. Biol. Chem. 2006, 281, 9030–9037. [Google Scholar] [CrossRef] [Green Version]
  113. Saulle, E.; Spinello, I.; Quaranta, M.T.; Pasquini, L.; Pelosi, E.; Iorio, E.; Castelli, G.; Chirico, M.; Pisanu, M.E.; Ottone, T.; et al. Targeting Lactate Metabolism by Inhibiting MCT1 or MCT4 Impairs Leukemic Cell Proliferation, Induces Two Different Related Death-Pathways and Increases Chemotherapeutic Sensitivity of Acute Myeloid Leukemia Cells. Front. Oncol. 2021, 10, 3394. [Google Scholar] [CrossRef]
  114. Singleton, R.S.; Guise, C.P.; Ferry, D.M.; Pullen, S.M.; Dorie, M.J.; Brown, J.M.; Patterson, A.V.; Wilson, W.R. DNA Cross-Links in Human Tumor Cells Exposed to the Prodrug PR-104A: Relationships to Hypoxia, Bioreductive Metabolism, and Cytotoxicity. Cancer Res. 2009, 69, 3884–3891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Patterson, A.V.; Ferry, D.M.; Edmunds, S.J.; Gu, Y.; Singleton, R.S.; Patel, K.; Pullen, S.M.; Hicks, K.O.; Syddall, S.P.; Atwell, G.J.; et al. Mechanism of Action and Preclinical Antitumor Activity of the Novel Hypoxia-Activated DNA Cross-Linking Agent PR-104. Clin. Cancer Res. 2007, 13, 3922–3932. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Guise, C.P.; Abbattista, M.R.; Singleton, R.S.; Holford, S.D.; Connolly, J.; Dachs, G.U.; Fox, S.B.; Pollock, R.; Harvey, J.; Guilford, P.; et al. The Bioreductive Prodrug PR-104A Is Activated under Aerobic Conditions by Human Aldo-Keto Reductase 1C3. Cancer Res. 2010, 70, 1573–1584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Jamieson, S.M.F.; Gu, Y.; Manesh, D.M.; El-Hoss, J.; Jing, D.; MacKenzie, K.L.; Guise, C.P.; Foehrenbacher, A.; Pullen, S.M.; Benito, J.; et al. A novel fluorometric assay for aldo-keto reductase 1C3 predicts metabolic activation of the nitrogen mustard prodrug PR-104A in human leukaemia cells. Biochem. Pharmacol. 2014, 88, 36–45. [Google Scholar] [CrossRef] [PubMed]
  118. Konopleva, M.; Thall, P.F.; Yi, C.A.; Borthakur, G.; Coveler, A.; Bueso-Ramos, C.; Benito, J.; Konoplev, S.; Gu, Y.; Ravandi, F.; et al. Phase I/II study of the hypoxia-activated prodrug PR104 in refractory/relapsed acute myeloid leukemia and acute lymphoblastic leukemia. Haematologica 2015, 100, 927–934. [Google Scholar] [CrossRef] [Green Version]
  119. Portwood, S.; Lal, D.; Hsu, Y.; Vargas, R.; Johnson, M.K.; Wetzler, M.; Hart, C.P.; Wang, E.S. Activity of the Hypoxia-Activated Prodrug, TH-302, in Preclinical Human Acute Myeloid Leukemia Models. Clin. Cancer Res. 2013, 19, 6506–6519. [Google Scholar] [CrossRef] [Green Version]
  120. Benito, J.; Ramirez, M.S.; Millward, N.Z.; Velez, J.; Harutyunyan, K.G.; Lu, H.; Shi, Y.; Matre, P.; Jacamo, R.; Ma, H.; et al. Hypoxia-Activated Prodrug TH-302 Targets Hypoxic Bone Marrow Niches in Preclinical Leukemia Models. Clin. Cancer Res. 2016, 22, 1687–1698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Badar, T.; Handisides, D.R.; Benito, J.M.; Richie, M.A.; Borthakur, G.; Jabbour, E.; Harutyunyan, K.; Konoplev, S.; Faderl, S.; Kroll, S.; et al. Phase I study of evofosfamide, an investigational hypoxia-activated prodrug, in patients with advanced leukemia. Am. J. Hematol. 2016, 91, 800–805. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Regulation of HIF-1α protein under normoxia and hypoxia. Abbreviations: PHD: prolyl hydroxylase; VHL: Von Hippel-Lindau-associated E3-ubiquitin ligase; HIF: hypoxia-inducible factor; HRE, hypoxia-responsive element.
Figure 1. Regulation of HIF-1α protein under normoxia and hypoxia. Abbreviations: PHD: prolyl hydroxylase; VHL: Von Hippel-Lindau-associated E3-ubiquitin ligase; HIF: hypoxia-inducible factor; HRE, hypoxia-responsive element.
Ijms 22 06857 g001
Figure 2. Representation of bone marrow endosteal and vascular niches with the principal cellular components and the oxygen gradient that controls HSC fate. HSC: hematopoietic stem cell.
Figure 2. Representation of bone marrow endosteal and vascular niches with the principal cellular components and the oxygen gradient that controls HSC fate. HSC: hematopoietic stem cell.
Ijms 22 06857 g002
Figure 3. Major metabolic alterations induced by hypoxia in AML. The exposure to hypoxia induces the up-regulation (green) of glycolysis associated with increased pyruvate, lactate, and alanine levels, and the down-regulation (red) of TCA and OXPHOS responsible for TCA intermediate accumulation. Moreover, depending on cell type, hypoxic conditions can increase some detoxifying metabolites, including glutamate, reduced and oxidized glutathione (GSSG/GSH), as well as fatty acid synthesis. Abbreviations: 2-HG: 2-hydroxyglutarate; GSSG: oxidized glutathione; GSH: reduced glutathione; TCA: tricarboxylic acid; OXPHOS: oxidative phosphorylation.
Figure 3. Major metabolic alterations induced by hypoxia in AML. The exposure to hypoxia induces the up-regulation (green) of glycolysis associated with increased pyruvate, lactate, and alanine levels, and the down-regulation (red) of TCA and OXPHOS responsible for TCA intermediate accumulation. Moreover, depending on cell type, hypoxic conditions can increase some detoxifying metabolites, including glutamate, reduced and oxidized glutathione (GSSG/GSH), as well as fatty acid synthesis. Abbreviations: 2-HG: 2-hydroxyglutarate; GSSG: oxidized glutathione; GSH: reduced glutathione; TCA: tricarboxylic acid; OXPHOS: oxidative phosphorylation.
Ijms 22 06857 g003
Figure 4. Hypoxia-altered metabolites control key cellular processes of relevance for leukemic progressions, such as epigenetic modification and signaling. Abbreviations: 2-HG: 2-hydroxyglutarate; αKG: α ketoglutarate; Lac: lactate; Suc: succinate; NDGR3: N-Myc downstream-regulated 3; ERK1/2: extracellular-signal-regulated kinase 1/2; GPR91: G-protein-coupled receptor 91; mTOR: mammalian target of rapamycin.
Figure 4. Hypoxia-altered metabolites control key cellular processes of relevance for leukemic progressions, such as epigenetic modification and signaling. Abbreviations: 2-HG: 2-hydroxyglutarate; αKG: α ketoglutarate; Lac: lactate; Suc: succinate; NDGR3: N-Myc downstream-regulated 3; ERK1/2: extracellular-signal-regulated kinase 1/2; GPR91: G-protein-coupled receptor 91; mTOR: mammalian target of rapamycin.
Ijms 22 06857 g004
Figure 5. Hallmarks of hypoxia in AML.
Figure 5. Hallmarks of hypoxia in AML.
Ijms 22 06857 g005
Table 1. List of novel compounds developed to target LSCs in the hypoxic microenvironment. Abbreviation: na: not available.
Table 1. List of novel compounds developed to target LSCs in the hypoxic microenvironment. Abbreviation: na: not available.
CompoundMechanism of ActionClinical Trial IDReferences Preclinical Studies
BL8040CXCR4 inhibitionNCT01838395
NCT03154827
[99]
IMC-1C11VEGFR-2 inhibitionna[82]
IACS-010759Complex I (OXPHOS) inhibitionNCT02882321[101]
AR-C155858/SyrosingopineMCT1 and MCT4
inhibition
na[103]
PR104DNA cross-linksNCT01037556[104,105,108]
TH-302DNA cross-linksNCT01149915[109,110,111]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bruno, S.; Mancini, M.; De Santis, S.; Monaldi, C.; Cavo, M.; Soverini, S. The Role of Hypoxic Bone Marrow Microenvironment in Acute Myeloid Leukemia and Future Therapeutic Opportunities. Int. J. Mol. Sci. 2021, 22, 6857. https://doi.org/10.3390/ijms22136857

AMA Style

Bruno S, Mancini M, De Santis S, Monaldi C, Cavo M, Soverini S. The Role of Hypoxic Bone Marrow Microenvironment in Acute Myeloid Leukemia and Future Therapeutic Opportunities. International Journal of Molecular Sciences. 2021; 22(13):6857. https://doi.org/10.3390/ijms22136857

Chicago/Turabian Style

Bruno, Samantha, Manuela Mancini, Sara De Santis, Cecilia Monaldi, Michele Cavo, and Simona Soverini. 2021. "The Role of Hypoxic Bone Marrow Microenvironment in Acute Myeloid Leukemia and Future Therapeutic Opportunities" International Journal of Molecular Sciences 22, no. 13: 6857. https://doi.org/10.3390/ijms22136857

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