IMR Press / FBS / Volume 14 / Issue 1 / DOI: 10.31083/j.fbs1401008
Open Access Review
Aldehyde Dehydrogenase Enzyme Functions in Acute Leukemia Stem Cells
Show Less
1 Department of Computer Science, Eastern Connecticut State University, Willimantic, CT 06226, USA
2 Cancer Immunology & Immunotherapy Center, Saint Savas Cancer Hospital, 11522 Athens, Greece
3 First Department of Pediatrics, National and Kapodistrian University of Athens, 11527 Goudi-Athens, Greece
*Correspondence: sblachop@med.uoa.gr (Spiros Vlahopoulos)
Academic Editor: Bela Ozsvari
Front. Biosci. (Schol Ed) 2022, 14(1), 8; https://doi.org/10.31083/j.fbs1401008
Submitted: 28 December 2021 | Revised: 23 February 2022 | Accepted: 25 February 2022 | Published: 8 March 2022
Copyright: © 2022 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

The enzymes that belong to the aldehyde dehydrogenase family are expressed in a variety of cells; yet activity of their main members characterizes stem cells, both normal and malignant. Several members of this family perform critical functions in stem cells, in general, and a few have been shown to have key roles in malignant tumors and their recurrence. In particular, ALDH1A1, which localizes to the cytosol and the nucleus, is an enzyme critical in cancer stem cells. In acute myeloid leukemia (AML), ALDH1A1 protects leukemia-initiating cells from a number of antineoplastic agents, and proves vital for the establishment of human AML xenografts in mice. ALDH2, which is located in mitochondria, has a major role in alcohol metabolism by clearing ethanol-derived acetaldehyde. Haematopoietic stem cells require ALDH2 for protection against acetaldehyde, which can cause damage to DNA, leading to insertions, deletions, chromosomal rearrangements, and translocations. Mutations compromise stem cell function, and thereby threaten blood homeostasis. We review here the potential of targeting the enzymatic activity of aldehyde dehydrogenases in acute leukemia.

Keywords
aldehyde dehydrogenase
drug resistance
immunosuppression
leukemia
myeloid
acute
neoplastic stem cells
gene expression
biomarker
1. Introduction

Aldehyde dehydrogenases are enzymes that oxidize aldehydes to carboxylic acids [1]. Aldehydes are generated during cellular metabolism, and when accumulated they react and form adducts with a variety of biomolecules, causing damage to proteins, DNA, lipids, etc. [2, 3].

In the organism, ethanol is metabolized to acetaldehyde that interferes with systems that protect cells from oxidant stress. ALDH proteins oxidize acetaldehyde to acetate, protecting cells and tissues from oxidative damage and helping maintain tissue function. In particular, ALDH proteins protect both normal as well as malignant stem cells (cancer stem cells) from reactive aldehydes and certain cytotoxic drugs such as cyclophosphamide [4, 5]. Increases in endogenous formaldehyde deplete blood progenitor cells [6, 7, 8]. According to one line of evidence, hematopoietic stem cells require the enzymatic activity of mitochondrial ALDH2 to protect them from the accumulation of acetaldehyde, which causes damage to DNA [3]. Protection of mice or humans from lethal formaldehyde accumulation has been attributed to ALDH2 in combination with cytoplasmic alcohol dehydrogenase 5 (ADH5) [9]. Among other effects, lack of those enzymatic activities (ALDH2 and ADH5) impairs also the differentiation and proliferation capacity of hematopoietic stem cells [10].

Recently, RNA from the gene that encodes ALDH1A1 was shown to associate with disease outcome in acute myeloid leukemia: lower expression of ALDH1A1 distinguished the group of patients with cytogenetics and molecular genetic markers that give them a favorable prognosis [11, 12].

2. ALDH in Cancer Cells

Not only normal hematopoietic cells, but malignant cells also gain protection when expressing ALDH2: mesenchymal cells from bone marrow stroma secrete transforming growth factor-β1 (TGF-β1), to induce ALDH2 activity in acute myeloid leukemia (AML) cells via non-canonical TGF-β signaling. Consequently, inhibition of ALDH2 sensitizes AML cells to chemotherapy [13]. The neoplasia that has been most extensively studied in respect to ALDH expression is breast cancer. There, expression of two main isoforms has been linked with malignant progression and ensuing metastasis: ALDH1A3 and ALDH1A1 [14].

When a DNA plasmid was used to express ALDH proteins in breast cancer cell line SUM159, three family members were detected by immunofluerescence in the cytoplasm and in the nucleus (ALDH1A1, ALDH3A1, ALDH7A1), five in the cytosol (ALDH1A2, ALDH1A3, ALDH1L1, ALDH8A1, ALDH9A1), seven in mitochondria (ALDH1B1, ALDH1L2, ALDH2, ALDH4A1, ALDH5A1, ALDH6A1, ALDH18A1), one in peroxisomes (ALDH3A2), one on the cell membrane (ALDH3B1), one on the cell membrane and in the cytosol (ALDH16A1) and one gave peri-nuclear staining (ALDH3B2) [15]. From all those, at least ALDH1A1 [16] (oral, ovarian, and breast cancer), ALDH1A3 [17] (pancreatic cancer), ALDH2 [18] (liver), ALDH3A1 [19] (melanoma, lung), have been associated with cancer stem cells, and specific phases of malignant progression, including tumor initiation, drug resistance, and metastasis [16]. Isoforms ALDH1L1, and ALDH1L2 in contrast, function in folate metabolism with the former acting as tumor suppressor, while the latter as oncogene [20].

3. ALDH Activity Identifies Progenitor Cell Fractions in Tissues

ALDH activity is mainly measured as fluorescence intensity of a product. Specifically, live cells expressing high ALDH activity are usually identified by the ALDEFLUOR kit and sorted by fluorescence activated cell sorting (FACS) [21]. The principle of this assay is that an uncharged substrate of ALDH is taken up by living cells, where ALDH converts it to a brightly fluorescent, negatively charged product that is trapped in cells that have an intact cellular membrane. The brightly fluorescent ALDH-expressing cells (ALDHbr) are detected in the green fluorescence channel (520–540 nm) of a standard flow cytometer, or can be purified by use of a fluorescence activated cell sorter [22, 23].

ALDHbr cells isolated from the human bone marrow, exhibited a higher colony-forming capacity when compared to a cells with low ALDH activity (Capoccia, et al. [24]). These cells could function as progenitors for epithelial, endothelial, and mesenchymal lineages [25]. Furthermore, when assayed in a mouse model of myocardial infarction, ALDHbr cells isolated from the human umbilical cord blood had the capacity to augment angiogenesis in the ischemic heart [26]. Lung-resident ALDHbr progenitor cells have shown a higher proliferative and colony-forming potential than cell populations with low ALDH activity, and the capacity to prevent bleomycin-induced pulmonary fibrosis [27]. In general, a variety of tissues harbor ALDHbr cells, which bestow them with a significant regenerative potential. Examples are pancreatic tissue [28], adipose tissue that harbors cells with both adipogenic and osteogenic potential [29], foreskin that harbors cells with multilineage potential [30], skeletal muscle [31], cardiac atrial appendage [32], maternal decidua basalis [33], and trachea [34].

4. ALDH Activity in Acute Leukemia

Several studies have linked increased ALDH activity with acute leukemia stem cells, although the overwhelming majority concerns AML, and some studies have used samples from both acute lymphoblastic leukemia (ALL) and AML.

In the study model of the zebrafish for engrafted human acute leukemia cells, higher ALDH activity, less differentiated cells and a broader and random migration pattern were related with worse clinical outcome after induction chemotherapy for patients [35].

In ALL, and especially in pediatric ALL, ALDH positivity is a marker of immaturity and stemness [36]. From the bone marrow of ALL patients, two cell populations could be separated, normal haematopoietic progenitor cells (ALDH(+)SSC(lo)CD45(hi)Neu5,9Ac2 -GPs(lo)CD34(+)CD38(-)CD90(+)CD117(+)CD133(+)) that differentiated into morphologically discrete, lineage-specific colonies, being essential for autologous HSC transplantation, while leukemic stem cells had the markers (ALDH(+)SSC(lo)CD45(lo)Neu5,9Ac2 -GPs(hi)CD34(+)CD38(+)CD90(-)CD117(-)CD133(-)) [37]. In B-ALL, ALDH+ cells formed 5–7 fold more colonies than ALDHneg cells in methylcellulose [38]. In T-ALL, transcription factor TAL1 activated expression of ALDH1A2, which supported glycolysis and the TCA cycle, NADH production, and ATP production, and decreased the levels of reactive oxygen species in vitro and in vivo; overexpression of ALDH1A2 accelerated tumor onset and increased tumor penetrance in a zebrafish T-ALL model [39].

In AML, high aldehyde dehydrogenase activity at diagnosis predicts relapse in patients with t(8;21) AML [40]. In general, evidence has been mounting that the protein family of the aldehyde dehydrogenases plays a role in the development of AML; yet the interpretation of their exact biological effect on AML progression remains debatable [41]. Higher ALDH activity was shown in normal hematopoietic stem cells, when compared to AML stem cells in a study of 32 patients and five bone marrow donors [42]. In a larger sample size (n = 104), a minority of patients (24 of 104) had numerous ALDH-positive leukemic stem cells that could not be separated from normal hematopoietic stem cells, were drug-resistant, and had high efficiency of xenograft formation in mice [43]. Nevertheless, cases with increased fractions of ALDH-positive AML cells were shown to derive from immature hematopoietic progenitors, suggesting an explanation for the poor prognosis and therapy resistance of this subgroup, which was attributed to the transmission of stem cell properties [44, 45]. It can be therefore hypothesized that ALDH activity is important for AML cells, yet the AML cells that retain ALDH activity are not abundant.

4.1 ALDH Family Members with a Potential Role in Leukemia

Although most, if not all ALDH proteins have been linked to phenomena relevant in malignancy, five members stand out as potential regulators of malignant stem cell activities, which can be implicated in key functions of leukemia stem cells. These are ALDH1A1, ALDH1A3, ALDH2, ALDH3A1, and ALDH3A2.

4.1.1 ALDH1A1

ALDH family member ALDH1A1, encoded in chromosome 9q21, is located in the cytosol and the nucleus, and catalyzes the second step of the main oxidative pathway of alcohol metabolism.

In general, proteins that belong to the subfamily ALDH1 convert the vitamin A oxidation derivative retinaldehyde, into retinoic acid, which activates retinoic acid receptors (RARs) α, β, and γ. In turn, RARs function as regulators of gene expression that controls differentiation of diverse cell types, and primes tissues for repair, regrowth and regeneration after inflammation or injury, further activating genes of the HOX family; the resulting HOX proteins then induce cellular programs of positional memory during embryogenesis and regeneration [46]. ALDH1A1 may also catalyze the oxidation of toxic lipid aldehydes such as 4-hydroxynonenal and malonaldehyde, and is the enzyme responsible for production of retinoid acid during the late stages of dorsal eye development [47]. In neoplasia, in addition to its known enzymatic activity, ALDH1A1 supports tumor growth via glutathione/dihydrolipoic acid-dependent NAD + reduction [48]. ALDH1A1 overexpression has also been associated with sorafenib resistance in AML and other malignancies [49, 50, 51, 52]. ALDH1A1 low expression was recently shown to characterize the favourable risk group in AML [11].

Furthermore, when ALDH1A1 is compared with the recently published strong AML predictor gene CALCRL in the TARGET AML cohort [53], it is found that ALDH1A1 is arguably a better predictor of the favourable risk group (Fig. 1) (For details see Supplementary Material).

Fig. 1.

Comparison of the AML risk group distribution of RNA expression from the genes ALDH1A1 and CALCRL. Association of ALDH1A1 (left) and CALCRL (right) RNA expression (y-axis) with risk groups (x-axis) in the TARGET AML cohort. p-values calculated by one-way ANOVA.

In physiological function, ALDH1A1 is elevated in hematopoietic stem cells, is a predominant ALDH isoform in mammalian tissues, and the key ALDH isozyme linked to both normal and malignant stem cells [54]. ALDH1A1 in hematopoietic stem cells metabolizes reactive aldehydes and reactive oxygen species, and has a potential role in protection of neurons from 3,4-dihydroxyphenylacetaldehyde [55]. Consequently, decreased expression of ALDH1A1 has been linked to the development of Parkinsons’ disease [56]. Furthermore, the ALDH1A1 protein is a critical enzyme for maintaining clarity in human, rat, and mouse lenses, and corneal integrity; ALDH1A1-null mice grow to adulthood, but develop cataracts later in life (by six to nine months of age) [57, 58, 59, 60]. Hematopoietic cells from ALDH1A1-deficient mice exhibit increased sensitivity to liver metabolites of cyclophosphamide; however, ALDH1A1 deficiency did not affect the basal function of stem cells from the hematopoietic and nervous system [61].

4.1.2 ALDH1A3

ALDH1A3, encoded in human chromosome 15 (15q26.3), plays a crucial role in the synthesis of retinoic acid in the cytosol, and it is essential for development of the ventral part of the eye [62, 63] and cardiomyogenesis [32]. In pancreatic cancer, ALDH1A3 was proposed to accelerate metastasis by augmenting glucose metabolism via increased expression of hexokinase 2 [17]. In other cancer types ALDH1A3 was shown to regulate tumorigenicity and metastasis in vivo models, with the net effect depending on the state of the retinoic acid signaling apparatus [14].

4.1.3 ALDH2

ALDH2, encoded in chromosomal locus 12q24.12, has a low Km for acetaldehydes, making mitochondria a subcellular location of efficient acetaldehyde metabolism [64, 65].

Analysis of intermediate-risk cytogenetic acute myeloid leukemia patients by RT-PCR has shown that ALDH2 overexpression correlated to shorter overall survival [66]. It has been postulated that ALDH2 expression reflects the adverse prognostic impact of age in AML, while it has been implicated in chemoresistance [67]. AML cells that express low levels of either ALDH1A1 or ALDH2 were sensitive to blocking of the UBE2T/FANCL-mediated ubiquitination pathway for DNA repair; this vulnerability could be ameliorated by expression of recombinant ALDH2, but not ALDH1A1. ALDH2 alone could suppress formaldehyde accumulation, and appeared highly expressed in AML samples carrying TP53 and KMT2A alterations; in contrast, AML samples with low ALDH2 expression were mostly carrying mutations in NPM1, CEBPA, and the RUNX1–RUNX1T1 (AML1–ETO) translocation [68]. Enzymatic activity of ALDH1A generates 9-cis retinoic acid, which is a ligand for retinoic X receptor (RXR), a transcription factor that activates ALDH2 gene expression [69, 70]. ALDH1A enzymes also generate all-trans-retinoic acid, which stimulates ALDH2 gene expression via retinoic acid receptors (RARs) [71]. High enzymatic activity of ALDH1A family member proteins can therefore be expected to lead to ALDH2 gene expression.

4.1.4 ALDH3A1

ALDH3A1, encoded in chromosomal locus 17p11.2, is expressed in human cornea and is the main ALDH detected in saliva under physiological conditions; it has a poor affinity for short-chain aliphatic aldehydes, while having a high substrate specificity for medium-chain (6 carbons and more) saturated and unsaturated aldehydes, including 4-hydroxy-2-nonenal, which are generated by the peroxidation of cellular lipids [72, 73, 74]. In hepatocellular carcinoma cells, overexpression of ALDH3A1 induced resistance to arsenic trioxide [75].

Also the enzyme ALDH3A2 oxidizes long-chain aliphatic aldehydes to prevent cellular oxidative damage. ALDH3A2 inhibition was synthetically lethal with glutathione peroxidase-4 (GPX4) inhibition; synergistically they triggered ferroptosis in AML cells, which could not be induced by GPX4 inhibition alone [76].

5. AML Features and the Relevance of Leukemia-initiating Cells

AML is the most common form of acute leukemia in adults, commonly accompanied by a poor prognosis, with less than 25% of patients surviving five years after diagnosis [77, 78, 79]. AML is a clonal hematopoetic disease that is also associated with a rather high death rate in pediatric (>30%) patients as well. AML is classified according to patient history into primary, de novo AML, which arises in the absence of an identified exposure or prodromal stem cell disorder and non-denovo AML; non-denovo includes secondary AML, representing transformation of an antecedent diagnosis of myelodysplastic syndrome (MDS) or myeloproliferative neoplasm, therapy-related AML developing as a late complication in patients with prior exposure to leukemogenic therapies [80]. From a biological and therapeutic standpoint, relapsed AML is included in non-de novo AML [81].

AML is genetically heterogeneous, yet it involves a rather small number of genetic alterations, with a marked similarity of the molecular spectrum between adult and pediatric patients [82, 83]. The World Health Organization defines specific acute myeloid leukemia (AML) disease entities by cytogenetic and molecular genetic subgroups: recurring, balanced cytogenetic abnormalities are recognized in AML, with few exceptions [84].

The clinical course of AML is largely determined by the presence of specific chromosomal aberrations in the cell nucleus, classifying patient cases into low risk, intermediate risk, and high risk. Absence of chromosomal aberrations classifies cases into the intermediate risk group, while cases with detected aberrations are classified into any one of the three categories [85]. In 2017, mutational screening for genes RUNX1, ASXL1, and TP53 was added to the prognostic criteria [86].

Within the heterogeneous population of leukemic cells, researchers identified rarer cells that were rather quiescent, and therefore resistant to antineoplastic treatments that target proliferating cells; these cells could give rise to successful mouse xenografts, their frequency increases after disease relapse and their phenotype has been attributed to epigenetic changes that involve DNA methylation and modification of histones on chromatin, which modulate gene expression, leading to changes in the cellular metabolism, and to resistance to agents that induce stem cell differentiation [87, 88].

AML can show durable remission by transplantation of healthy (normal, non-malignant) stem cells; yet substantial work is needed to focus on the effects of therapeutic agents on the interactions between key molecular factors that form the leukemia microenvironment, especially to distinguish between healthy and leukemic stem cells [89]. Research needs to target the leukemia-initiating cells (“leukemia stem cells”) selectively, as LSC appear to cause the most severe conditions of AML [90]. A “stemness signature” has been related to risk prediction in 908 patients of diverse AML subtypes, reinforcing this notion [91]. It was then found that approximately 25% of all acute myeloid leukemias expressed low or undetectable levels of ALDH1A1 and that this ALDH1A1-subset of leukemias correlates with good prognosis cytogenetics. ALDH1A1-cell lines as well as primary leukemia cells were found to be sensitive to treatment with compounds that directly and indirectly generate toxic ALDH substrates including 4-hydroxynonenal and the clinically relevant compounds arsenic trioxide and 4-hydroperoxycyclophosphamide. In contrast, normal hematopoietic stem cells were relatively resistant to these compounds. Using a murine xenotransplant model to emulate a clinical treatment strategy, established ALDH1A1- leukemias were also sensitive to in vivo treatment with cyclophosphamide combined with arsenic trioxide [92]. Furthermore, it was found that approximately 25% of all acute myeloid leukemias expressed low or undetectable levels of ALDH1A1 and that this ALDH1A1- subset of leukemias correlates with good prognosis cytogenetics. ALDH1A1- cell lines as well as primary leukemia cells were found to be sensitive to treatment with compounds that directly and indirectly generate toxic ALDH substrates including 4-hydroxynonenal and the clinically relevant compounds arsenic trioxide and 4-hydroperoxycyclophosphamide. In contrast, normal hematopoietic stem cells were relatively resistant to these compounds. Using a murine xenotransplant model to emulate a clinical treatment strategy, established ALDH1A1-leukemias were also sensitive to in vivo treatment with cyclophosphamide combined with arsenic trioxide [92]. Subsequently, it was found that low expression of that gene could be statistically associated with the low risk group across multiple cohorts, strongly associating low ALDH1A1 expression with the favorable, low-risk AML group [12].

A hematopoietic stem cell when dividing asymmetrically, it can generate myeloid progenitor cells, which give rise to myeloblasts. Aberrantly developed myeloid progenitor cells or myeloblasts give rise to acute myeloid leukemia [93]. In patients, AML cells with a primitive stem cell phenotype (CD34+/CD38- and high aldehyde dehydrogenase activity) cause significantly lower complete remission rates, as well as poorer event-free and overall survival [94]. One key difference between myelodysplastic syndromes and AML, were balanced cytogenetic rearrangements (p < 0.0001), which could thus be associated with initiation of leukaemia [95].

Cytogenetic rearrangements may involve the gene mixed lineage leukaemia (MLL), in which the N-terminal portion of MLL is fused to the C-terminus of the translocation partner; these rearrangements give rise to altered proteins that control epigenetic modifications of nuclear chromatin and regulation of gene expression thus linking cytogenetic rearrangements with molecular changes that show the potential to initiate leukemia [96].

5.1 Maintenance of Defining Features of AML Stem Cells: the Role of Insensitivity to Retinoic Acid

The bone marrow microenvironment of leukemia-initiating cells has a pivotal role in the biology of the disease [97]. Deficient signaling from the vitamin A derivative, all-trans retinoic acid in the hematopoietic microenvironment in mice was shown to cause a myeloproliferative syndrome with significant reduction in the numbers of B lymphocyte subsets and erythrocytes in the bone marrow [98]. In patients with acute promyelocytic leukemia, high blast cell counts and failure to respond to differentiation treatment was associated with low all-trans retinoic acid plasma concentrations, and a high rate of clearance [99]. For the other types of AML, and despite its importance in controlling myeloid differentiation and apoptotic pathways, all-trans retinoic acid has yet to prove itself as a useful agent in the armamentarium of AML therapeutics. The explanation probably lies in the fact that retinoic acid receptor is expressed, but of limited functionality in AML blast cells [100]. However there are both positive and negative associations of CD34+ immature blast cells with sensitivity to all-trans retinoic acid [101, 102]. Presence of internal tandem duplications of the FLT3 gene in AML stem cells, reinforces the leukemia-initiating capacity of those cells, which is experimentally demonstrated by successful engraftment into mice [103].

5.2 Pathways Signaling to NFκB Function in AML Stem Cells

A key transcription factor implicated in ALDH1 regulation is nuclear factor kappa B (NFκB) a factor found constitutively active in malignant myeloblast cells of AML [104, 105, 106]. Cells normally activate NFκB in response to inflammatory stimuli and disruptions of tissue function [107]. NFκB has an important role in the control of AML stem and progenitor cells, especially in the regulation of interactions between AML cells and their microenvironment [88, 108, 109, 110]. In poor-prognosis AML, ALDH1A1 RNA is expressed far above the level that is required for the myeloid lineage. Transcription factors such as TLX1/HOX11 and NFκB, both associated with a severe course of AML [111, 112] drive ALDH1A1 overexpression, promoting myelopoiesis [113].

To activate ALDH1A1 expression indirectly, NFκB induces expression of micro RNA223-3p, which inhibits expression of ARID1A. ARID1A loss, in turn allows histone acetylation of the ALDH1A1 gene promoter [114]; [115]. Having an established capacity to guide epigenetic changes that guide leukemic stem cell programs, NFκB has been associated with cancer recurrence, and with AML relapse [109, 116, 117, 118].

In AML, constitutive NFκB DNA-binding activity is frequently mediated by a Ras/PI3-K/PKB-dependent pathway [119]. Also overexpression of FLT3 induces NFκB-dependent transcriptional activity in cultured cells [120]. The constitutive activation of FLT3-ITD also induces NFκB activity [121]. And conversely, FLT3 inhibition or knockdown reduces constitutive NFκB activation in high-risk myelodysplastic syndrome and AML [122]. NFκB is regulated by changes in cellular proteostasis; its native inhibitor, the IkB protein can be degraded by cellular proteolytic systems such as the proteasome or the lysosome, which are mutually regulated through the induction of transcription factors [123, 124]. Inhibition of NFκB by the sesquiterpene lactone parthenolide was shown to suppress P-glycoprotein expression and adriamycin resistance in cultured AML stem cells [125, 126]. In fact, induction of cell stress combined with NFκB inhibition is a promising approach in AML treatment, due to the connection between proteostatic pathways activated by cell stress and those inducing NFκB activity, as was recently shown in a clinical trial of relapsed patients with acute promyelocytic leukemia [127, 128].

One agent that is used in treatment of AML, is the nucleoside cytosine arabinoside [129]. In fact, cytosine arabinoside (cytarabine, ara-C), in high concentrations (>50 uM) inhibits NFκB in some cell types to a certain extent; yet this inhibition is evidently not enough, since after development of drug resistance, ara-C does not suffice as monotherapy for AML [130, 131, 132].

On the one hand, mutated, constitutively active FLT3, activates signaling molecules RAS, and PI3K/mTOR which activate NFκB [133]. On the other hand, AML can develop resistance against all inhibitors of FLT3 at some point [134]. Consequently, combination of NFκB and FLT3 inhibitors has been previously suggested by in vitro and xenograft studies of AML cells from patients [135, 136]. An additional reason to combine inhibitors is the fact that NFκB causes epigenetic changes in nuclear chromatin, which cannot be reversed by inhibition of NFκB itself in the affected cells [137]. Therefore, in spite of its central role in regulation of inflammatory gene expression, NFκB can be difficult to target under certain conditions, due to positive feedback loops that are formed by some of the products of its target genes (Fig. 2), Downstream target genes of NFκB-guided transcription, therefore, such as ALDH1A1, are attractive aims for AML treatment.

Fig. 2.

Overview of the effects of AML stem cell signaling. Due to the extensive phenotypic changes in the cell, upstream signals are rapidly converted to multi-pronged cascades, which become largely self-sustained. Specifically, the induction of expression of diverse downstream genes triggers both developmental changes and alters sensitivity to differentiation signals and to apoptotic stimuli, leading to cells that are substantially unresponsive to physiological homeostatic regulators and at the same time resist pharmacological intervention and escape immune surveillance.

6. Inhibition of ALDH as a Potential Therapeutic Approach

ALDH1A1, ALDH1A3, ALDH2 and ALDH3A1 have structural and functional differences in substrate binding site, cofactor dissociation, enzyme kinetics, and rate-limiting steps, which facilitates the design of selective inhibitors and enzymatic activity tracers; in particular, ALDH1A1 substrate binding pocket has a wider access tunnel than ALDH2 or ALDH3A1, allowing ALDH1A1 to accommodate larger and more rigid ligands than other ALDH isoforms [138, 139].

ALDH inhibition is not yet established as a therapeutic method in malignant disease; however clinical trials testing ALDH inhibitors in cancer are ongoing. Several trials use disulfiram, an FDA-approved drug to treat alcohol use disorder, to treat central nervous system neoplasms and other types of malignancy [140] (U.S. National Library of Medicine resource platform ClinicalTrials.gov registration numbers NCT01777919, NCT02770378, NCT01907165, NCT03363659, NCT02678975, NCT03151772, NCT03034135, NCT02715609; for an overview see https://clinicaltrials.gov/ct2/results?term=disulfiram&cond=Cancer). Disulfiram inhibits multiple functions of malignant cells, both directly, as well as through its metabolites, especially in the presence of copper that accumulates and is essential for tumor cells [140].

Disulfiram inhibits human lens ALDH1A1 at IC50 values of the micromolar range [59]. In breast cancer cells, ALDH1A1 enzymatic activity facilitated breast tumor growth, and acidified the cytosol to promote phosphorylation of TAK1, activate NFκB signaling, and to increase the secretion of granulocyte macrophage colony-stimulating factor (GM-CSF), which led to myeloid-derived suppressor cell (MDSC) expansion and immunosuppression; disulfiram and chemotherapeutic agent gemcitabine cooperatively inhibited breast tumor growth and tumorigenesis by purging ALDH+ cancer stem cells, and by activating T cell immunity [141].

In fact, in spite of the presence of ALDH activity in hematopoietic stem cells, disulfiram with copper could overcome bortezomib and cytarabine resistance in ALDHbr LSCs via inducing apoptosis and proteasome inhibition [142, 143]. Furthemore, another ALDH inhibitor, namely diethylaminobenzaldehyde (DEAB), could induce the expansion of normal human hematopoietic stem cells [144]. DEAB-inhibition of ALDH delayed hematopoietic differentiation and expanded multipotent myeloid cells that accelerated vascular regeneration following intramuscular transplantation into immunodeficient mice with hind-limb ischemia [145]. Yet a further ALDH inhibitor, namely dimethyl ampal thiolester (DIMATE) eradicated leukemia stem cells while sparing normal progenitors, both in vitro as well as in mouse xenografts of human AML cells [146].

To date, several inhibitors for ALDH are under development, with the aim a selective effect on specific proteins, especially ALDH1A1. Indeed, novel selective inhibitors for ALDH1A1 were developed recently [147, 148, 149, 150]. This field of research shows marked progress due to the fact that ALDH1A1 is implicated in pathological manifestations of inflammatory and metabolic syndromes in addition to cancer [151, 152].

Also compounds derived from natural substances exhibit inhibitory activity against ALDH. Notably, the terpenoid citral (3,7-dimethyl-2,6-octadien-1-al), a component of essential oils obtained from several plants, which is used as a food additive and fragrance, inhibits human lens ALDH1A1 at low IC50 values of the micromolar range [59]. Spice and herb extracts have shown a wealth in modulatory activities of ALDH family members ALDH1A1, ALDH1A2, ALDH1A3 and ALDH2; even though it may be possible that extracts contain selective modulators, certain extracts manifested general effects, with sage and thyme extracts indicating the potential for complete suppression of the enzymatic activity of all four proteins tested [153].

7. Conclusions

Although initially the principal research finding was that cancer stem cell properties are linked to ALDH1A1, it later became apparent that other family members can be implicated in cancer stem cell function. A wider range of ALDH are linked to cancer stem cell activities than what was expected, and it is likely that ALDH members have the capacity to replace one another under specific conditions, which depends strongly on the host tissue and the metabolic conditions encountered by the malignant cells. This condition is expected to hold in leukemia too, with a particular role of the transitions between bone marrow niche and circulation taking effect in leukemia subtypes that show corresponding transitions between ALDH activity.

Abbreviations

AKT/PKB, RAC-alpha serine/threonine-protein kinase; ALDH, Aldehyde dehydrogenase; AML, Acute myeloid leukemia; ara-c, cytosine arabinoside; CEBPA, CCAAT enhancer binding protein alpha; del, deletion; ELN, European LeukemiaNet; EVI1, ecotropic viral integration site 1; FLT3, Fms-like tyrosine kinase-3; FLT3 ITD, internal tandem duplication of FLT3; inv, inversion; MTOR, kinase, mechanistic/mammalian target of rapamycin; NFκB, nuclear factor kappa B; NOD/SCID, nonobese diabetic/severe combined immunodeficiency mice; NPM1, nucleophosmin 1; OGG1, 8-oxoguanine glycosylase; PD-L1, Programmed death-ligand 1; PI3K, phosphatidylinositol 3-kinase; t, translocation; TLX1/HOX11, T cell leukemia/homeobox 1.

Author contributions

GMD, IFV, SV—conception and design of the research study. GMD—data analysis. GMD, IFV, SV—manuscript writing. GMD, IFV, SV—manuscript editing. GMD, IV, SV—approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Acknowledgment

Authors wish to thank the AMLCG study group for help in this project.

Funding

This work was supported, in part, by a grant from the American Association of University Professors and Connecticut State University Board of Regents (YRDA01).

Conflict of interest

The authors declare no conflict of interest.

References
[1]
Koppaka V, Thompson DC, Chen Y, Ellermann M, Nicolaou KC, Juvonen RO, et al. Aldehyde dehydrogenase inhibitors: a comprehensive review of the pharmacology, mechanism of action, substrate specificity, and clinical application. Pharmacological Reviews. 2012; 64: 520–539.
[2]
Sapkota M, Wyatt TA. Alcohol, Aldehydes, Adducts and Airways. Biomolecules. 2015; 5: 2987–3008.
[3]
Garaycoechea JI, Crossan GP, Langevin F, Mulderrig L, Louzada S, Yang F, et al. Alcohol and endogenous aldehydes damage chromosomes and mutate stem cells. Nature. 2018; 553: 171–177.
[4]
Kastan MB, Schlaffer E, Russo JE, Colvin OM, Civin CI, Hilton J. Direct demonstration of elevated aldehyde dehydrogenase in human hematopoietic progenitor cells. Blood. 1990; 75: 1947–1950.
[5]
Bidan N, Bailleul‐Dubois J, Duval J, Winter M, Denoulet M, Hannebicque K, et al. Transcriptomic Analysis of Breast Cancer Stem Cells and Development of a pALDH1A1:mNeptune Reporter System for Live Tracking. Proteomics. 2019; 19: 1800454.
[6]
Pontel LB, Rosado IV, Burgos-Barragan G, Garaycoechea JI, Yu R, Arends MJ, et al. Endogenous Formaldehyde is a Hematopoietic Stem Cell Genotoxin and Metabolic Carcinogen. Molecular Cell. 2015; 60: 177–188.
[7]
Mu A, Hira A, Niwa A, Osawa M, Yoshida K, Mori M, et al. Analysis of disease model iPSCs derived from patients with a novel Fanconi anemia–like IBMFS ADH5/ALDH2 deficiency. Blood. 2021; 137: 2021–2032.
[8]
Jung M, Smogorzewska A. Endogenous formaldehyde destroys blood stem cells. Blood. 2021; 137: 1988–1990.
[9]
Dingler FA, Wang M, Mu A, Millington CL, Oberbeck N, Watcham S, et al. Two Aldehyde Clearance Systems are Essential to Prevent Lethal Formaldehyde Accumulation in Mice and Humans. Molecular Cell. 2020; 80: 996–1012.e9.
[10]
Oka Y, Hamada M, Nakazawa Y, Muramatsu H, Okuno Y, Higasa K, et al. Digenic mutations in ALDH2 and ADH5 impair formaldehyde clearance and cause a multisystem disorder, AMeD syndrome. Science Advances.2020; 6: eabd7197.
[11]
Dancik GM, Voutsas IF, Vlahopoulos S. Lower RNA expression of ALDH1A1 distinguishes the favorable risk group in acute myeloid leukemia. 2022. (in press)
[12]
Dancik GM, Voutsas IF, Vlahopoulos S. Lower RNA expression of ALDH1A1 distinguishes the favorable risk group in acute myeloid leukemia. Molecular Biology Reports. 2022. (in press)
[13]
Yuan B, El Dana F, Ly S, Yan Y, Ruvolo V, Shpall EJ, et al. Bone marrow stromal cells induce an ALDH+ stem cell-like phenotype and enhance therapy resistance in AML through a TGF-β-p38-ALDH2 pathway. PLoS ONE. 2020; 15: e0242809.
[14]
Rodriguez-Torres M, Allan AL. Aldehyde dehydrogenase as a marker and functional mediator of metastasis in solid tumors. Clinical & Experimental Metastasis. 2016; 33: 97–113.
[15]
Zhou L, Sheng D, Wang D, Ma W, Deng Q, Deng L, et al. Identification of cancer-type specific expression patterns for active aldehyde dehydrogenase (ALDH) isoforms in ALDEFLUOR assay. Cell Biology and Toxicology. 2019; 35: 161–177.
[16]
Mohan A, Raj RR, Mohan G, K PP, Thomas Maliekal T. Reporters of Cancer Stem Cells as a Tool for Drug Discovery. Frontiers in Oncology. 2021; 11: 669250.
[17]
Nie S, Qian X, Shi M, Li H, Peng C, Ding X, et al. ALDH1A3 Accelerates Pancreatic Cancer Metastasis by Promoting Glucose Metabolism. Frontiers in Oncology. 2020; 10: 915.
[18]
Chen L, Wu M, Ji C, Yuan M, Liu C, Yin Q. Silencing transcription factor FOXM1 represses proliferation, migration, and invasion while inducing apoptosis of liver cancer stem cells by regulating the expression of ALDH2. IUBMB Life. 2020; 72: 285–295.
[19]
Terzuoli E, Bellan C, Aversa S, Ciccone V, Morbidelli L, Giachetti A, et al. ALDH3A1 Overexpression in Melanoma and Lung Tumors Drives Cancer Stem Cell Expansion, Impairing Immune Surveillance through Enhanced PD-L1 Output. Cancers. 2019; 11: E1963.
[20]
Krupenko SA, Krupenko NI. ALDH1L1 and ALDH1L2 Folate Regulatory Enzymes in Cancer. Advances in Experimental Medicine and Biology. 2018; 1032: 127–143.
[21]
Alison MR, Guppy NJ, Lim SM, Nicholson LJ. Finding cancer stem cells: are aldehyde dehydrogenases fit for purpose? The Journal of Pathology. 2010; 222: 335–344.
[22]
Storms RW, Trujillo AP, Springer JB, Shah L, Colvin OM, Ludeman SM, et al. Isolation of primitive human hematopoietic progenitors on the basis of aldehyde dehydrogenase activity. Proceedings of the National Academy of Sciences of the United States of America. 1999; 96: 9118–9123.
[23]
Storms RW, Green PD, Safford KM, Niedzwiecki D, Cogle CR, Colvin OM, et al. Distinct hematopoietic progenitor compartments are delineated by the expression of aldehyde dehydrogenase and CD34. Blood. 2005; 106: 95–102.
[24]
Capoccia BJ, Robson DL, Levac KD, Maxwell DJ, Hohm SA, Neelamkavil MJ, et al. Revascularization of ischemic limbs after transplantation of human bone marrow cells with high aldehyde dehydrogenase activity. Blood. 2009; 113: 5340–5351.
[25]
Gentry T, Foster S, Winstead L, Deibert E, Fiordalisi M, Balber A. Simultaneous isolation of human BM hematopoietic, endothelial and mesenchymal progenitor cells by flow sorting based on aldehyde dehydrogenase activity: implications for cell therapy. Cytotherapy. 2007; 9: 259–274.
[26]
Sondergaard CS, Hess DA, Maxwell DJ, Weinheimer C, Rosová I, Creer MH, et al. Human cord blood progenitors with high aldehyde dehydrogenase activity improve vascular density in a model of acute myocardial infarction. Journal of Translational Medicine. 2010; 8: 24.
[27]
Takahashi H, Nakashima T, Masuda T, Namba M, Sakamoto S, Yamaguchi K, et al. Antifibrotic effect of lung-resident progenitor cells with high aldehyde dehydrogenase activity. Stem Cell Research & Therapy. 2021; 12: 471.
[28]
Loomans CJM, Williams Giuliani N, Balak J, Ringnalda F, van Gurp L, Huch M, et al. Expansion of Adult Human Pancreatic Tissue Yields Organoids Harboring Progenitor Cells with Endocrine Differentiation Potential. Stem Cell Reports. 2018; 10: 712–724.
[29]
Itoh H, Nishikawa S, Haraguchi T, Arikawa Y, Eto S, Hiyama M, et al. Aldehyde dehydrogenase activity helps identify a subpopulation of murine adipose-derived stem cells with enhanced adipogenic and osteogenic differentiation potential. World Journal of Stem Cells. 2017; 9: 179–186.
[30]
Najar M, Crompot E, van Grunsven LA, Dollé L, Lagneaux L. Foreskin-derived mesenchymal stromal cells with aldehyde dehydrogenase activity: isolation and gene profiling. BMC Cell Biology. 2018; 19: 4.
[31]
Etienne J, Joanne P, Catelain C, Riveron S, Bayer AC, Lafable J, et al. Aldehyde dehydrogenases contribute to skeletal muscle homeostasis in healthy, aging, and Duchenne muscular dystrophy patients. Journal of Cachexia, Sarcopenia and Muscle. 2020; 11: 1047–1069.
[32]
Puttini S, Plaisance I, Barile L, Cervio E, Milano G, Marcato P, et al. ALDH1a3 is the Key Isoform that Contributes to Aldehyde Dehydrogenase Activity and Affects in Vitro Proliferation in Cardiac Atrial Appendage Progenitor Cells. Frontiers in Cardiovascular Medicine. 2018; 5: 90.
[33]
Kusuma GD, Abumaree MH, Pertile MD, Perkins AV, Brennecke SP, Kalionis B. Mesenchymal Stem/Stromal Cells Derived from a Reproductive Tissue Niche under Oxidative Stress have High Aldehyde Dehydrogenase Activity. Stem Cell Reviews and Reports. 2016; 12: 285–297.
[34]
Ghosh M, Helm KM, Smith RW, Giordanengo MS, Li B, Shen H, et al. A single cell functions as a tissue-specific stem cell and the in vitro niche-forming cell. American Journal of Respiratory Cell and Molecular Biology. 2011; 45: 459–469.
[35]
Gacha-Garay MJ, Niño-Joya AF, Bolaños NI, Abenoza L, Quintero G, Ibarra H, et al. Pilot Study of an Integrative New Tool for Studying Clinical Outcome Discrimination in Acute Leukemia. Frontiers in Oncology. 2019; 9: 245.
[36]
Ahlers J, Witte K, Schwarze CP, Lang P, Handgretinger R, Ebinger M. Therapy response correlates with ALDH activity in ALDH low-positive childhood acute lymphoblastic leukemias. Pediatric Hematology and Oncology. 2014; 31: 303–310.
[37]
Chowdhury S, Chandra S, Mandal C. 9-O-acetylated sialic acids differentiating normal haematopoietic precursors from leukemic stem cells with high aldehyde dehydrogenase activity in children with acute lymphoblastic leukaemia. Glycoconjugate Journal. 2014; 31: 523–535.
[38]
Lin TL, Wang QH, Brown P, Peacock C, Merchant AA, Brennan S, et al. Self-renewal of acute lymphocytic leukemia cells is limited by the Hedgehog pathway inhibitors cyclopamine and IPI-926. PLoS ONE. 2010; 5: e15262.
[39]
Zhang C, Amanda S, Wang C, King Tan T, Zulfaqar Ali M, Zhong Leong W, et al. Oncorequisite role of an aldehyde dehydrogenase in the pathogenesis of T-cell acute lymphoblastic leukemia. Haematologica. 2021; 106: 1545–1558.
[40]
Yang L, Chen W, Dao F, Zhang Y, Wang Y, Chang Y, et al. High aldehyde dehydrogenase activity at diagnosis predicts relapse in patients with t(8;21) acute myeloid leukemia. Cancer Medicine. 2019; 8: 5459–5467.
[41]
Gasparetto M, Smith CA. ALDHs in normal and malignant hematopoietic cells: Potential new avenues for treatment of AML and other blood cancers. Chemico-Biological Interactions. 2017; 276: 46–51.
[42]
Schuurhuis GJ, Meel MH, Wouters F, Min LA, Terwijn M, de Jonge NA, et al. Normal hematopoietic stem cells within the AML bone marrow have a distinct and higher ALDH activity level than co-existing leukemic stem cells. PLoS ONE. 2013; 8: e78897.
[43]
Hoang VT, Buss EC, Wang W, Hoffmann I, Raffel S, Zepeda-Moreno A, et al. The rarity of ALDH(+) cells is the key to separation of normal versus leukemia stem cells by ALDH activity in AML patients. International Journal of Cancer. 2015; 137: 525–536.
[44]
Ran D, Schubert M, Pietsch L, Taubert I, Wuchter P, Eckstein V, et al. Aldehyde dehydrogenase activity among primary leukemia cells is associated with stem cell features and correlates with adverse clinical outcomes. Experimental Hematology. 2009; 37: 1423–1434.
[45]
Blume R, Rempel E, Manta L, Saeed BR, Wang W, Raffel S, et al. The molecular signature of AML with increased ALDH activity suggests a stem cell origin. Leukemia & Lymphoma. 2018; 59: 2201–2210.
[46]
Gudas LJ. Emerging roles for retinoids in regeneration and differentiation in normal and disease states. Biochimica Et Biophysica Acta. 2012; 1821: 213–221.
[47]
Fan X, Molotkov A, Manabe S, Donmoyer CM, Deltour L, Foglio MH, et al. Targeted disruption of Aldh1a1 (Raldh1) provides evidence for a complex mechanism of retinoic acid synthesis in the developing retina. Molecular and Cellular Biology. 2003; 23: 4637–4648.
[48]
Wang B, Chen X, Wang Z, Xiong W, Xu T, Zhao X, et al. Aldehyde dehydrogenase 1a1 increases NADH levels and promotes tumor growth via glutathione/dihydrolipoic acid-dependent NAD+ reduction. Oncotarget. 2017; 8: 67043–67055.
[49]
Burchert A. Maintenance therapy for FLT3-ITD-mutated acute myeloid leukemia. Haematologica. 2021; 106: 664–670.
[50]
Man CH, Fung TK, Ho C, Han HHC, Chow HCH, Ma ACH, et al. Sorafenib treatment of FLT3-ITD(+) acute myeloid leukemia: favorable initial outcome and mechanisms of subsequent nonresponsiveness associated with the emergence of a D835 mutation. Blood. 2012; 119: 5133–5143.
[51]
Elcheva IA, Wood T, Chiarolanzio K, Chim B, Wong M, Singh V, et al. RNA-binding protein IGF2BP1 maintains leukemia stem cell properties by regulating HOXB4, MYB, and ALDH1A1. Leukemia. 2020; 34: 1354–1363.
[52]
Allison SE, Chen Y, Petrovic N, Zhang J, Bourget K, Mackenzie PI, et al. Activation of ALDH1A1 in MDA-MB-468 breast cancer cells that over-express CYP2J2 protects against paclitaxel-dependent cell death mediated by reactive oxygen species. Biochemical Pharmacology. 2017; 143: 79–89.
[53]
Angenendt L, Wöste M, Mikesch J, Arteaga MF, Angenendt A, Sandmann S, et al. Calcitonin receptor-like (CALCRL) is a marker of stemness and an independent predictor of outcome in pediatric AML. Blood Advances. 2021; 5: 4413–4421.
[54]
Tomita H, Tanaka K, Tanaka T, Hara A. Aldehyde dehydrogenase 1A1 in stem cells and cancer. Oncotarget. 2016; 7: 11018–11032.
[55]
Anderson DW, Schray RC, Duester G, Schneider JS. Functional significance of aldehyde dehydrogenase ALDH1a1 to the nigrostriatal dopamine system. Brain Research. 2011; 1408: 81–87.
[56]
Fan HH, Guo Q, Zheng J, Lian YZ, Huang SS, Sun Y, et al. ALDH1A1 Genetic Variations May Modulate Risk of Parkinson’s Disease in Han Chinese Population. Frontiers in Neuroscience. 2021; 15: 620929.
[57]
Choudhary S, Xiao T, Vergara LA, Srivastava S, Nees D, Piatigorsky J, et al. Role of aldehyde dehydrogenase isozymes in the defense of rat lens and human lens epithelial cells against oxidative stress. Investigative Ophthalmology & Visual Science. 2005; 46: 259–267.
[58]
Lassen N, Bateman JB, Estey T, Kuszak JR, Nees DW, Piatigorsky J, et al. Multiple and Additive Functions of ALDH3A1 and ALDH1A1. Journal of Biological Chemistry. 2007; 282: 25668–25676.
[59]
Xiao T, Shoeb M, Siddiqui MS, Zhang M, Ramana KV, Srivastava SK, et al. Molecular cloning and oxidative modification of human lens ALDH1A1: implication in impaired detoxification of lipid aldehydes. Journal of Toxicology and Environmental Health, Part A. 2009; 72: 577–584.
[60]
Yu O, Biswas S, Ma G, Zhao P, Li B, Li J. Canonical NF-κB signaling maintains corneal epithelial integrity and prevents corneal aging via retinoic acid. ELife. 2021; 10: e67315.
[61]
Levi BP, Yilmaz OH, Duester G, Morrison SJ. Aldehyde dehydrogenase 1a1 is dispensable for stem cell function in the mouse hematopoietic and nervous systems. Blood. 2009; 113: 1670–1680.
[62]
Vasiliou V, Nebert DW. Analysis and update of the human aldehyde dehydrogenase (ALDH) gene family. Human Genomics. 2005; 2: 138–143.
[63]
Fares-Taie L, Gerber S, Chassaing N, Clayton-Smith J, Hanein S, Silva E, et al. ALDH1A3 mutations cause recessive anophthalmia and microphthalmia. American Journal of Human Genetics. 2013; 92: 265–270.
[64]
Hsu LC, Yoshida A, Mohandas T. Chromosomal assignment of the genes for human aldehyde dehydrogenase-1 and aldehyde dehydrogenase-2. American Journal of Human Genetics. 1986; 38: 641–648.
[65]
Glatt H, Rost K, Frank H, Seidel A, Kollock R. Detoxification of promutagenic aldehydes derived from methylpyrenes by human aldehyde dehydrogenases ALDH2 and ALDH3a1. Archives of Biochemistry and Biophysics. 2008; 477: 196–205.
[66]
Torrebadell M, Díaz-Beyá M, Kalko SG, Pratcorona M, Nomdedeu J, Navarro A, et al. A 4-gene expression prognostic signature might guide post-remission therapy in patients with intermediate-risk cytogenetic acute myeloid leukemia. Leukemia & Lymphoma. 2018; 59: 2394–2404.
[67]
Hernandez-Valladares M, Aasebø E, Berven F, Selheim F, Bruserud Ø. Biological characteristics of aging in human acute myeloid leukemia cells: the possible importance of aldehyde dehydrogenase, the cytoskeleton and altered transcriptional regulation. Aging. 2020; 12: 24734–24777.
[68]
Yang Z, Wu XS, Wei Y, Polyanskaya SA, Iyer SV, Jung M, et al. Transcriptional Silencing of ALDH2 Confers a Dependency on Fanconi Anemia Proteins in Acute Myeloid Leukemia. Cancer Discovery. 2021; 11: 2300–2315.
[69]
Paterson EK, Ho H, Kapadia R, Ganesan AK. 9-cisretinoic acid is the ALDH1a1 product that stimulates melanogenesis. Experimental Dermatology. 2013; 22: 202–209.
[70]
Pinaire J, Hasanadka R, Fang M, Chou WY, Stewart MJ, Kruijer W, et al. The retinoid X receptor response element in the human aldehyde dehydrogenase 2 promoter is antagonized by the chicken ovalbumin upstream promoter family of orphan receptors. Archives of Biochemistry and Biophysics. 2000; 380: 192–200.
[71]
Pinaire J, Chou W, Morton M, You M, Zeng Y, Cho WK, et al. Identification of a retinoid receptor response element in the human aldehyde dehydrogenase-2 promoter. Alcoholism, Clinical and Experimental Research. 2003; 27: 1860–1866.
[72]
Vasiliou V, Bairoch A, Tipton KF, Nebert DW. Eukaryotic aldehyde dehydrogenase (ALDH) genes: human polymorphisms, and recommended nomenclature based on divergent evolution and chromosomal mapping. Pharmacogenetics. 1999; 9: 421–434.
[73]
Pappa A, Estey T, Manzer R, Brown D, Vasiliou V. Human aldehyde dehydrogenase 3A1 (ALDH3A1): biochemical characterization and immunohistochemical localization in the cornea. The Biochemical Journal. 2003; 376: 615–623.
[74]
Alam MF, Laskar AA, Maryam L, Younus H. Activation of Human Salivary Aldehyde Dehydrogenase by Sulforaphane: Mechanism and Significance. PLoS ONE. 2016; 11: e0168463.
[75]
Zhang X, Yang X, Sun C, Hu B, Sun Y, Huang X, et al. Promyelocytic leukemia protein induces arsenic trioxide resistance through regulation of aldehyde dehydrogenase 3 family member a1 in hepatocellular carcinoma. Cancer Letters. 2015; 366: 112–122.
[76]
Yusuf RZ, Saez B, Sharda A, van Gastel N, Yu VWC, Baryawno N, et al. Aldehyde dehydrogenase 3a2 protects AML cells from oxidative death and the synthetic lethality of ferroptosis inducers. Blood. 2020; 136: 1303–1316.
[77]
Pollyea DA, Kohrt HE, Medeiros BC. Acute myeloid leukaemia in the elderly: a review. British Journal of Haematology. 2011; 152: 524–542.
[78]
Rodrigues ACBDC, Costa RGA, Silva SLR, Dias IRSB, Dias RB, Bezerra DP. Cell signaling pathways as molecular targets to eliminate AML stem cells. Critical Reviews in Oncology/Hematology. 2021; 160: 103277.
[79]
Shallis RM, Wang R, Davidoff A, Ma X, Zeidan AM. Epidemiology of acute myeloid leukemia: Recent progress and enduring challenges. Blood Reviews. 2019; 36: 70–87.
[80]
Lindsley RC, Mar BG, Mazzola E, Grauman PV, Shareef S, Allen SL, et al. Acute myeloid leukemia ontogeny is defined by distinct somatic mutations. Blood. 2015; 125: 1367–1376.
[81]
Bachas C, Schuurhuis GJ, Hollink IHIM, Kwidama ZJ, Goemans BF, Zwaan CM, et al. High-frequency type i/II mutational shifts between diagnosis and relapse are associated with outcome in pediatric AML: implications for personalized medicine. Blood. 2010; 116: 2752–2758.
[82]
Marjanovic I, Kostic J, Stanic B, Pejanovic N, Lucic B, Karan-Djurasevic T, et al. Parallel targeted next generation sequencing of childhood and adult acute myeloid leukemia patients reveals uniform genomic profile of the disease. Tumor Biology. 2016; 37: 13391–13401.
[83]
Zjablovskaja P, Florian MC. Acute Myeloid Leukemia: Aging and Epigenetics. Cancers. 2019; 12: 103.
[84]
Arber DA, Orazi A, Hasserjian R, Thiele J, Borowitz MJ, Le Beau MM, et al. The 2016 revision to the World Health Organization classification of myeloid neoplasms and acute leukemia. Blood. 2016; 127: 2391–2405.
[85]
Grimwade D, Walker H, Oliver F, Wheatley K, Harrison C, Harrison G, et al. The Importance of Diagnostic Cytogenetics on Outcome in AML: Analysis of 1,612 Patients Entered into the MRC AML 10 Trial. Blood. 1998; 92: 2322–2333.
[86]
Döhner H, Estey E, Grimwade D, Amadori S, Appelbaum FR, Büchner T, et al. Diagnosis and management of AML in adults: 2017 ELN recommendations from an international expert panel. Blood. 2017; 129: 424–447.
[87]
Thomas D, Majeti R. Biology and relevance of human acute myeloid leukemia stem cells. Blood. 2017; 129: 1577–1585.
[88]
Reinisch A, Chan SM, Thomas D, Majeti R. Biology and Clinical Relevance of Acute Myeloid Leukemia Stem Cells. Seminars in Hematology. 2015; 52: 150–164.
[89]
Mendez LM, Posey RR, Pandolfi PP. The Interplay Between the Genetic and Immune Landscapes of AML: Mechanisms and Implications for Risk Stratification and Therapy. Frontiers in Oncology. 2019; 9: 1162.
[90]
Sendker S, Reinhardt D, Niktoreh N. Redirecting the Immune Microenvironment in Acute Myeloid Leukemia. Cancers. 2021; 13: 1423.
[91]
Ng SWK, Mitchell A, Kennedy JA, Chen WC, McLeod J, Ibrahimova N, et al. A 17-gene stemness score for rapid determination of risk in acute leukaemia. Nature. 2016; 540: 433–437.
[92]
Gasparetto M, Pei S, Minhajuddin M, Khan N, Pollyea DA, Myers JR, et al. Targeted therapy for a subset of acute myeloid leukemias that lack expression of aldehyde dehydrogenase 1A1. Haematologica. 2017; 102: 1054–1065.
[93]
Basilico S, Wang X, Kennedy A, Tzelepis K, Giotopoulos G, Kinston SJ, et al. Dissecting the early steps of MLL induced leukaemogenic transformation using a mouse model of AML. Nature Communications. 2020; 11: 1407.
[94]
Karantanos T, Jones RJ. Acute Myeloid Leukemia Stem Cell Heterogeneity and its Clinical Relevance. Stem Cells Heterogeneity in Cancer. 2019; 1139: 153–169.
[95]
Larson RA, Le Beau MM. Therapy-related myeloid leukaemia: a model for leukemogenesis in humans. Chemico-Biological Interactions. 2005; 153–154: 187–195.
[96]
Zhang Y, Chen A, Yan XM, Huang G. Disordered epigenetic regulation in MLL-related leukemia. International Journal of Hematology. 2012; 96, 428–437.
[97]
Kokkaliaris KD, Scadden DT. Cell interactions in the bone marrow microenvironment affecting myeloid malignancies. Blood Advances. 2020; 4: 3795–3803.
[98]
Walkley CR, Olsen GH, Dworkin S, Fabb SA, Swann J, McArthur GA, et al. A microenvironment-induced myeloproliferative syndrome caused by retinoic acid receptor gamma deficiency. Cell. 2007; 129: 1097–1110.
[99]
Lefebvre P, Thomas G, Gourmel B, Agadir A, Castaigne S, Dreux C, et al. Pharmacokinetics of oral all-trans retinoic acid in patients with acute promyelocytic leukemia. Leukemia. 1991; 5: 1054–1058.
[100]
Johnson DE, Redner RL. An ATRActive future for differentiation therapy in AML. Blood Reviews. 2015; 29: 263–268.
[101]
Bradbury DA, Aldington S, Zhu YM, Russell NH. Down-regulation of bcl-2 in AML blasts by all-trans retinoic acid and its relationship to CD34 antigen expression. British Journal of Haematology. 1996; 94: 671–675.
[102]
Lehmann S, Bengtzen S, Broberg U, Paul C. Effects of retinoids on cell toxicity and apoptosis in leukemic blast cells from patients with non-M3 AML. Leukemia Research. 2000; 24: 19–25.
[103]
Levis M, Murphy KM, Pham R, Kim K, Stine A, Li L, et al. Internal tandem duplications of the FLT3 gene are present in leukemia stem cells. Blood. 2005; 106: 673–680.
[104]
Kallifatidis G, Labsch S, Rausch V, Mattern J, Gladkich J, Moldenhauer G, et al. Sulforaphane increases drug-mediated cytotoxicity toward cancer stem-like cells of pancreas and prostate. Molecular Therapy. 2011; 19: 188–195.
[105]
Liu L, Salnikov AV, Bauer N, Aleksandrowicz E, Labsch S, Nwaeburu C, et al. Triptolide reverses hypoxia-induced epithelial-mesenchymal transition and stem-like features in pancreatic cancer by NF-κB downregulation. International Journal of Cancer. 2014; 134: 2489–2503.
[106]
Rausch V, Liu L, Kallifatidis G, Baumann B, Mattern J, Gladkich J, et al. Synergistic activity of sorafenib and sulforaphane abolishes pancreatic cancer stem cell characteristics. Cancer Research. 2010; 70: 5004–5013.
[107]
Lambrou GI, Hatziagapiou K, Vlahopoulos S. Inflammation and tissue homeostasis: the NF-κB system in physiology and malignant progression. Molecular Biology Reports. 2020; 47: 4047–4063.
[108]
Wang Z, Chen J, Wang M, Zhang L, Yu L. One Stone, Two Birds: The Roles of Tim-3 in Acute Myeloid Leukemia. Front Immunol. 2021; 12: 618710.
[109]
Nasri F, Sadeghi F, Behranvand N, Samei A, Bolouri MR, Azari T, et al. Oridonin Could Inhibit Inflammation and T-cell Immunoglobulin and Mucin-3/Galectin-9 (TIM-3/Gal-9) Autocrine Loop in the Acute Myeloid Leukemia Cell Line (U937) as Compared to Doxorubicin. Iranian Journal of Allergy, Asthma and Immunology. 2020; 19: 602–611.
[110]
Kikushige Y, Miyamoto T, Yuda J, Jabbarzadeh-Tabrizi S, Shima T, Takayanagi S, et al. 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.
[111]
Li Z, Huang H, Li Y, Jiang X, Chen P, Arnovitz S, et al. Up-regulation of a HOXA-PBX3 homeobox-gene signature following down-regulation of miR-181 is associated with adverse prognosis in patients with cytogenetically abnormal AML. Blood. 2012; 119: 2314–2324.
[112]
Ignatz-Hoover JJ, Wang V, Mackowski NM, Roe AJ, Ghansah IK, Ueda M, et al. Aberrant GSK3β nuclear localization promotes AML growth and drug resistance. Blood Advances. 2018; 2: 2890–2903.
[113]
Rice KL, Izon DJ, Ford J, Boodhoo A, Kees UR, Greene WK. Overexpression of stem cell associated ALDH1A1, a target of the leukemogenic transcription factor TLX1/HOX11, inhibits lymphopoiesis and promotes myelopoiesis in murine hematopoietic progenitors. Leukemia Research. 2008; 32: 873–883.
[114]
Yang F, Xu Y, Liu C, Ma C, Zou S, Xu X, et al. NF-κB/miR-223-3p/ARID1a axis is involved in Helicobacter pylori CagA-induced gastric carcinogenesis and progression. Cell Death & Disease. 2018; 9: 12.
[115]
Yoshino J, Akiyama Y, Shimada S, Ogura T, Ogawa K, Ono H, et al. Loss of ARID1a induces a stemness gene ALDH1A1 expression with histone acetylation in the malignant subtype of cholangiocarcinoma. Carcinogenesis. 2020; 21: 734–742.
[116]
Kuo H, Wang Z, Lee D, Iwasaki M, Duque-Afonso J, Wong SK, et al. Epigenetic Roles of MLL Oncoproteins are Dependent on NF-κB. Cancer Cell. 2013; 24: 423–437.
[117]
Prajoko YW, Aryandono T. The Effect of P-Glycoprotein (P-gp), Nuclear Factor-Kappa B (Nf-κb), and Aldehyde Dehydrogenase-1 (ALDH-1) Expression on Metastases, Recurrence and Survival in Advanced Breast Cancer Patients. Asian Pacific Journal of Cancer Prevention. 2019; 20: 1511–1518.
[118]
Lück SC, Russ AC, Du J, Gaidzik V, Schlenk RF, Pollack JR, et al. KIT mutations confer a distinct gene expression signature in core binding factor leukaemia. British Journal of Haematology. 2010; 148: 925–937.
[119]
Birkenkamp KU, Geugien M, Schepers H, Westra J, Lemmink HH, Vellenga E. Constitutive NF-kappaB DNA-binding activity in AML is frequently mediated by a Ras/PI3-K/PKB-dependent pathway. Leukemia. 2004; 18: 103–112.
[120]
Takahashi S, Harigae H, Ishii KK, Inomata M, Fujiwara T, Yokoyama H, et al. Over-expression of Flt3 induces NF-kappaB pathway and increases the expression of IL-6. Leukemia Research. 2005; 29: 893–899.
[121]
Imbert V, Peyron JF. NF-κB in Hematological Malignancies. Biomedicines. 2017; 5: 27.
[122]
Grosjean-Raillard J, Adès L, Boehrer S, Tailler M, Fabre C, Braun T, et al. Flt3 receptor inhibition reduces constitutive NFkappaB activation in high-risk myelodysplastic syndrome and acute myeloid leukemia. Apoptosis. 2008; 13: 1148–1161.
[123]
Zhou J, Ching YQ, Chng W. Aberrant nuclear factor-kappa B activity in acute myeloid leukemia: from molecular pathogenesis to therapeutic target. Oncotarget. 2015; 6: 5490–5500.
[124]
Varisli L, Cen O, Vlahopoulos S. Dissecting pharmacological effects of chloroquine in cancer treatment: interference with inflammatory signaling pathways. Immunology. 2020; 159: 257–278.
[125]
Guzman ML, Rossi RM, Karnischky L, Li X, Peterson DR, Howard DS, et al. The sesquiterpene lactone parthenolide induces apoptosis of human acute myelogenous leukemia stem and progenitor cells. Blood. 2005; 105: 4163–4169.
[126]
Yi J, Wang L, Wang X, Sun J, Yin X, Hou J, et al. Suppression of Aberrant Activation of NF-κB Pathway in Drug-resistant Leukemia Stem Cells Contributes to Parthenolide-potentiated Reversal of Drug Resistance in Leukemia. Journal of Cancer. 2021; 12: 5519–5529.
[127]
Kulkarni U, Mathews V. Evolving Chemotherapy Free Regimens for Acute Promyelocytic Leukemia. Frontiers in Oncology. 2021; 11: 621566.
[128]
Kulkarni U, Ganesan S, Alex AA, Palani H, David S, Balasundaram N, et al. A phase II study evaluating the role of bortezomib in the management of relapsed acute promyelocytic leukemia treated upfront with arsenic trioxide. Cancer Medicine. 2020; 9: 2603–2610.
[129]
Takahashi S. Current Understandings of Myeloid Differentiation Inducers in Leukemia Therapy. Acta Haematologica. 2021; 144: 380–388.
[130]
Sreenivasan Y, Sarkar A, Manna SK. Mechanism of cytosine arabinoside-mediated apoptosis: role of Rel a (p65) dephosphorylation. Oncogene. 2003; 22: 4356–4369.
[131]
Murphy T, Yee KWL. Cytarabine and daunorubicin for the treatment of acute myeloid leukemia. Expert Opinion on Pharmacotherapy. 2017; 18: 1765–1780.
[132]
Löwenberg B. Sense and nonsense of high-dose cytarabine for acute myeloid leukemia. Blood. 2013; 121: 26–28.
[133]
Zhong Y, Qiu R, Sun S, Zhao C, Fan T, Chen M, et al. Small-Molecule Fms-like Tyrosine Kinase 3 Inhibitors: an Attractive and Efficient Method for the Treatment of Acute Myeloid Leukemia. Journal of Medicinal Chemistry. 2020; 63: 12403–12428.
[134]
Mosquera Orgueira A, Bao Pérez L, Mosquera Torre A, Peleteiro Raíndo A, Cid López M, Díaz Arias J, et al. FLT3 inhibitors in the treatment of acute myeloid leukemia: current status and future perspectives. Minerva Medica. 2020; 111: 427–442.
[135]
Griessinger E, Frelin C, Cuburu N, Imbert V, Dageville C, Hummelsberger M, et al. Preclinical targeting of NF-kappaB and FLT3 pathways in AML cells. Leukemia. 2008; 22: 1466–1469.
[136]
Wang C, Lu J, Wang Y, Bai S, Wang Y, Wang L, et al. Combined effects of FLT3 and NF-κB selective inhibitors on acute myeloid leukemia in vivo. Journal of Biochemical and Molecular Toxicology. 2012; 26: 35–43.
[137]
Vlahopoulos SA. Aberrant control of NF-κB in cancer permits transcriptional and phenotypic plasticity, to curtail dependence on host tissue: molecular mode. Cancer Biology & Medicine. 2017; 14: 254–270.
[138]
Pereira R, Gendron T, Sanghera C, Greenwood HE, Newcombe J, McCormick PN, et al. Mapping Aldehyde Dehydrogenase 1A1 Activity using an [18F]Substrate-Based Approach. Chemistry. 2019; 25: 2345–2351.
[139]
Ibrahim AIM, Ikhmais B, Batlle E, AbuHarb WK, Jha V, Jaradat KT, et al. Design, Synthesis, Biological Evaluation and In Silico Study of Benzyloxybenzaldehyde Derivatives as Selective ALDH1A3 Inhibitors. Molecules. 2021; 26: 5770.
[140]
Meraz-Torres F, Plöger S, Garbe C, Niessner H, Sinnberg T. Disulfiram as a Therapeutic Agent for Metastatic Malignant Melanoma-Old Myth or New Logos? Cancers. 2020; 12: E3538.
[141]
Liu C, Qiang J, Deng Q, Xia J, Deng L, Zhou L, et al. ALDH1A1 Activity in Tumor-Initiating Cells Remodels Myeloid-Derived Suppressor Cells to Promote Breast Cancer Progression. Cancer Research. 2021; 81: 5919–5934.
[142]
Yang X, Yao R, Wang H. Update of ALDH as a Potential Biomarker and Therapeutic Target for AML. BioMed Research International. 2018; 2018: 9192104.
[143]
Bista R, Lee DW, Pepper OB, Azorsa DO, Arceci RJ, Aleem E. Disulfiram overcomes bortezomib and cytarabine resistance in down-syndrome-associated acute myeloid leukemia cells. Journal of Experimental & Clinical Cancer Research. 2017; 36: 22.
[144]
Chute JP, Muramoto GG, Whitesides J, Colvin M, Safi R, Chao NJ, et al. Inhibition of aldehyde dehydrogenase and retinoid signaling induces the expansion of human hematopoietic stem cells. Proceedings of the National Academy of Sciences of the United States of America. 2006; 103: 11707–11712.
[145]
Cooper TT, Sherman SE, Kuljanin M, Bell GI, Lajoie GA, Hess DA. Inhibition of Aldehyde Dehydrogenase-Activity Expands Multipotent Myeloid Progenitor Cells with Vascular Regenerative Function. Stem Cells. 2018; 36: 723–736.
[146]
Venton G, Pérez-Alea M, Baier C, Fournet G, Quash G, Labiad Y, et al. Aldehyde dehydrogenases inhibition eradicates leukemia stem cells while sparing normal progenitors. Blood Cancer Journal. 2016; 6: e469.
[147]
Yang S, Martinez NJ, Yasgar A, Danchik C, Johansson C, Wang Y, et al. Discovery of Orally Bioavailable, Quinoline-Based Aldehyde Dehydrogenase 1A1 (ALDH1A1) Inhibitors with Potent Cellular Activity. Journal of Medicinal Chemistry. 2018; 61: 4883–4903.
[148]
Yasgar A, Titus SA, Wang Y, Danchik C, Yang S, Vasiliou V, et al. A High-Content Assay Enables the Automated Screening and Identification of Small Molecules with Specific ALDH1a1-Inhibitory Activity. PLoS ONE. 2017; 12: e0170937.
[149]
Ma Z, Jiang L, Li G, Liang D, Li L, Liu L, et al. Design, synthesis of 1,3-dimethylpyrimidine-2,4-diones as potent and selective aldehyde dehydrogenase 1A1 (ALDH1A1) inhibitors with glucose consumption improving activity. Bioorganic Chemistry. 2020; 101: 103971.
[150]
Li B, Yang K, Liang D, Jiang C, Ma Z. Discovery and development of selective aldehyde dehydrogenase 1A1 (ALDH1A1) inhibitors. European Journal of Medicinal Chemistry. 2021; 209: 112940.
[151]
Petrosino JM, Disilvestro D, Ziouzenkova O. Aldehyde dehydrogenase 1A1: friend or foe to female metabolism? Nutrients. 2014; 6: 950–973.
[152]
Zhong G, Kirkwood J, Won KJ, Tjota N, Jeong H, Isoherranen N. Characterization of Vitamin A Metabolome in Human Livers With and Without Nonalcoholic Fatty Liver Disease. Journal of Pharmacology and Experimental Therapeutics. 2019; 370: 92–103.
[153]
Bui TBC, Nosaki S, Kokawa M, Xu Y, Kitamura Y, Tanokura M, et al. Evaluation of spice and herb as phyto-derived selective modulators of human retinaldehyde dehydrogenases using a simple in vitro method. Bioscience Reports. 2021; 41: BSR20210491.
Share
Back to top