The retinoic acid receptors (RARA, RARB, and RARG) are ligand-regulated nuclear receptors that act as transcriptional switches. These master genes drew significant interest in the 1990s because of their key roles in embryogenesis and involvement in a rare malignancy, acute promyelocytic leukemia (APL), in which the RARA (and very rarely, RARG or RARB) genes are rearranged, underscoring the central role of deregulated retinoid signaling in leukemogenesis. Several recent provocative observations have revived interest in the roles of retinoids in non-APL acute myeloid leukemia (AML), as well as in normal hematopoietic differentiation. We review the role of retinoids in hematopoiesis, as well as in the treatment of non-APL AMLs. From this perspective, broader uses of retinoids in the management of hematopoietic tumors are discussed.

Nuclear receptors are transcription factors with activities that can be regulated by a variety of small molecules. These master genes raise fascinating biological questions, notably with respect to their roles in developmental programs or adaptive responses. They also open major therapeutic avenues in multiple pathologies in which they are directly or indirectly implicated. One subgroup, retinoic acid receptors (RARs), is particularly relevant to the field of hematopoiesis, as 2 of the 3 receptors, RARA and RARG, have been implicated as key regulators of normal and transformed blood cells. Moreover, acute promyelocytic leukemia (APL) is driven by fusion genes involving RARs, primarily RARA.1  Remarkably, APL is clinically exquisitely sensitive to all-trans-retinoic acid (ATRA), the natural ligand of the RARs, exemplifying one of the best examples of targeted therapy, and (at least when combined with arsenic trioxide), the only rapidly and definitively curative one. The physiopathology of APL and its basis for response to therapy has been reviewed elsewhere.2,3  The APL model highlights 2 important ideas discussed here in the context of both normal hematopoiesis and non-APL acute myeloid leukemias (AMLs): (1) RARs and ATRA signaling are important for stem cell self-renewal and hematopoietic differentiation; and (2) ATRA therapy can have profound impacts on the outcomes of hematopoietic malignancies.

The cloning of the RARs gave a molecular basis to ATRA signaling and prompted substantial efforts to understand its complex developmental and physiological impacts.4  In brief, ATRA signaling relies on 3 receptors: RARA, RARB, and RARG, as well as their retinoid X receptor (RXR) heterodimeric partners. Each receptor is expressed as several spliced isoform variants. ATRA is synthetized locally, through very complex mechanisms involving capture of retinol from the circulation, oxidation into retinal and retinoic acid, transport and sequestration, and, finally, degradation.5-7  Most, if not all, of the enzymes involved in retinoid metabolism are transcriptionally regulated by ATRA, ensuring, through local feed-forward and feed-back controls, the fine spatiotemporal tuning of ATRA abundance. Initially, simple models in which ligand-bound RARs complexed to DNA response elements to promote transcriptional activation of target genes prevailed.8  Yet, over the years, our knowledge of the molecular mechanisms involved has become more and more complex (Figure 1). RARA seems to exert primarily ATRA-reversible basal repressive functions, whereas RARG is generally a potent ligand-dependent transcriptional activator.9  Thus, different RARs may have antagonistic effects on shared targets. ATRA exposure can yield target transcriptional activation and initiate proteasome-dependent receptor degradation.10-12  Ligand-activated repression is well established for some targets, such as FGF8, which play a critical role in developmental pathways.5,7  Receptor activity can be regulated by posttranslational modifications, with functional consequences for embryonic stem cell (ESC) differentiation.13  Physiologically, basal repressive functions play a key role in development (especially in head formation in xenopus and mouse skeletal development),14  so that receptor loss or extinction can be quite different from vitamin A deprivation. Use of synthetic ligands can promote a variety of interactions between specific RARs and coactivator or corepressor proteins, broadening the diversity of molecular and cellular responses. Retinoic acid signaling can also be modulated by RXRs and their poorly characterized natural ligands, which seem to be important in hematopoietic cells.15  However, RXR signaling controls a much broader network, including the vitamin D3 receptor and other type 2 nuclear receptors, which also play critical roles in the pathophysiology of hematopoiesis and AML.16  This multiplicity of cross-connected effects stresses the importance of “clean” gene ablation experiments (or even ablation of response elements within target genes) in a physiological in vivo setting, rather than use of agonist or antagonist receptor ligands or dominant negative receptor expression in cell lines, to explore physiological questions.7 

Figure 1.

Overall effects of ATRA on RAR-mediated signaling. The data stress the diversity of mechanisms (basal repression, ligand-induced repression or activation, RAR degradation, cross talk with other key pathways) as well as important differences in ATRA levels implicated in these effects. In the hematopoietic lineage, RARA seems primarily involved in granulocytic differentiation, whereas RARG plays important roles in the biology of HSC and theirs niches.

Figure 1.

Overall effects of ATRA on RAR-mediated signaling. The data stress the diversity of mechanisms (basal repression, ligand-induced repression or activation, RAR degradation, cross talk with other key pathways) as well as important differences in ATRA levels implicated in these effects. In the hematopoietic lineage, RARA seems primarily involved in granulocytic differentiation, whereas RARG plays important roles in the biology of HSC and theirs niches.

Close modal

ATRA and RARA have also been implicated in imperfectly understood “nongenomic effects,” wherein RARA and ATRA effects are mediated outside of the nucleus, through modulation of protein kinase activity.17  Conversely, several drugs that target kinases modulate the effects of ATRA in AML cells (Table 1).18-20  Synergistic target gene activation and subsequent differentiation were reported with myelomonocytic growth factors (granulocyte and granulocyte-macrophage stimulating factor [G/GM-CSF]) and were thought to rely on activation of the mitogen-activated protein kinase (MAPK) pathway.21  Interestingly, upon administration of ATRA, MAPK can rapidly phosphorylate RARs and several associated proteins (corepressors, coactivators) such that this synergy may ultimately converge on ATRA target genes.20,22  In addition, activation of other kinases (including cyclin-dependent ones) targets not only RARA, but also coactivators and repressors.23  Finally, the notion that all ATRA-triggered effects necessarily occur through interactions with RARs has been challenged. Indeed, ATRA was shown to potently inhibit the peptidyl-prolyl cis-trans isomerase PIN-1 enzyme, which posttranslationally modifies a large number of proteins to promote growth.24-26  Importantly, PIN-1 is often activated in AMLs, such that its targeting by ATRA can impair growth.27  Collectively, these examples point to the versatility and complexity of the interplay between RARs, RXRs, and their ligands for the control of multiple biological processes. Although direct linking of a specific molecular mechanism to a given cellular phenotype has often been challenging, terminal granulocytic differentiation is thought to be RARA inhibited and ATRA/RARA activated,28  neuronal differentiation of ESCs requires a specific RARG isoform,13  and nongenomic RARA functions have recently been implicated in T-cell activation.29 

Table 1.

Some synergistic effects of anticancer drugs combined with ATRA in non-APL AML

Type of drugDrug targetTested modelsCombined effect with ATRAReference
DNA hypomethylating agents DNMT High-risk AML patient Complete clinical remissions with ATRA+decitabine 63, 101 
Histone modifiers Histone demethylase (LSD-1/KDM-1A) AML cell lines Enhanced AML cell differentiation 71, 73-75 
Histone acetyltransferase(GCN-5) AML cell lines Enhanced cell differentiation 124 
SUMOylation inhibitors SUMOylation enzymes (E1, E2) AML cell lines
Mouse xenografts 
Increased ATRA-induced differentiation 81 
Kinase inhibitors FLT3-ITD AML cell lines Increased apoptosis of leukemia cells and survival in mice 19 
HER-2 Mouse xenografts
AML cell lines 
Promotion of AML cell differentiation 18 
 
ATRA metabolism CYP26 AML cell lines Inhibited stromal CYP26 favoring ATRA-induced differentiation 125 
Type of drugDrug targetTested modelsCombined effect with ATRAReference
DNA hypomethylating agents DNMT High-risk AML patient Complete clinical remissions with ATRA+decitabine 63, 101 
Histone modifiers Histone demethylase (LSD-1/KDM-1A) AML cell lines Enhanced AML cell differentiation 71, 73-75 
Histone acetyltransferase(GCN-5) AML cell lines Enhanced cell differentiation 124 
SUMOylation inhibitors SUMOylation enzymes (E1, E2) AML cell lines
Mouse xenografts 
Increased ATRA-induced differentiation 81 
Kinase inhibitors FLT3-ITD AML cell lines Increased apoptosis of leukemia cells and survival in mice 19 
HER-2 Mouse xenografts
AML cell lines 
Promotion of AML cell differentiation 18 
 
ATRA metabolism CYP26 AML cell lines Inhibited stromal CYP26 favoring ATRA-induced differentiation 125 

An elegant study has demonstrated that RAR signaling is essential for the emergence of mouse hematopoietic stem cells (HSCs) from the hemogenic endothelium, through activation of WNT signaling.30  After completion of embryonic development, vitamin A, the precursor to ATRA, promotes HSC dormancy in adult mice,31  which is a likely explanation for the ability of ATRA to enhance the long-term repopulating activity of HSCs32  and protect them from exhaustion after stress-induced proliferative activation by interferons (IFNs). Conversely, deprivation of vitamin A induces a progressive loss of HSCs, impairs efficient restoration of HSC dormancy after inflammatory stress and promotes myeloid expansion.33  These ATRA/HSC connections illustrate one of the broad roles of retinoid signaling on multiple types of stem cells in vivo.7  Some master regulators of hematopoietic differentiation are under direct or indirect transcriptional control of retinoids, including KIT, CEBPA, WNT, and HOX genes.31  The HSC niche or, more broadly, the HSC microenvironment, appears to protect HSCs from pharmacological doses of ATRA,34  at least in part, through the expression of ATRA-degrading cytochromes.35  Downstream of HSCs, retinoids favor the emergence, growth, and differentiation of normal early committed myeloid progenitors with granulocytic potential, at the expense of erythroid and mastocytic ones36,37  (Figure 2). Overall, RA signaling is complex, depending on the stage of differentiation and drug concentrations, with low physiological levels and pharmacological doses occasionally conveying different signals.

Figure 2.

Overview of hematopoietic cell differentiation and impact of retinoid signaling. Green denotes activation; red denotes inhibition. RARA++, overexpressed RARA; RARAMut, mutated RARA; TAM, tumor-associated macrophage.

Figure 2.

Overview of hematopoietic cell differentiation and impact of retinoid signaling. Green denotes activation; red denotes inhibition. RARA++, overexpressed RARA; RARAMut, mutated RARA; TAM, tumor-associated macrophage.

Close modal

Molecularly, the contributions of individual RARs to hematopoiesis were first explored in RAR-deficient mice, demonstrating that basal granulocytic differentiation was accelerated not only by RARA excision, but also by pharmacological doses of ATRA solely in the presence of RARA.28  Thus, under physiological conditions, endogenous RARA slows down granulopoiesis. RARA is expressed from 2 alternative promoters, yielding isoforms with different N termini. These are differentially expressed during hematopoietic differentiation, with RARA-2 being massively induced upon myeloid differentiation.38  Therapeutically, acceleration of terminal differentiation by pharmacological concentrations of retinoids or RARA-specific agonists has been proposed in settings where functional granulocytes are limiting.39,40 

RARA is a low-efficacy receptor for ATRA (50% effective concentration [EC50] 240 nM), whereas RARG very avidly binds its ligand and activates target genes (EC50 6 nM) and accordingly plays a key role in embryonic development.41,42  Importantly, RARG appears to be a physiological regulator of HSC self-renewal.43  Remarkably, this effect also involves the microenvironment, as RARG ablation within this niche suffices to induce a myeloproliferative syndrome, even when RARG expression is retained in reintroduced wild-type HSCs.44  More studies are needed to decipher the respective roles of HSCs and the niche in RARG effects. RARG and retinoids facilitate conversion of fibroblasts to induced pluripotent stem cells, again underscoring their tight relationships with stem cell differentiation.45  RARG expression and a specific polymorphism have also been linked to doxorubicin-induced cardiotoxicity.46,47 

In contrast to RARA loss, dominant negative versions of RARA dramatically alter hematopoietic cell fate in several ways (reviewed by Collins36 ). First, these mutants may immortalize very early progenitors that are subsequently capable of multilineage differentiation (eg, toward B cells). Second, when expressed in a multipotent cell line, dominant negative RARA shifts spontaneous differentiation toward mastocytes, while abolishing the basal myelomonocytic fate. Moreover, when these progenitors are forced toward granulocytic lineage differentiation by exposure to GM-CSF, they undergo a promyelocytic differentiation block, which can be reversed by ATRA. In these conditions of overexpression, the dominant negative RARA mutant may perturb normal signaling of multiple other nuclear receptors (eg, by titrating their common RXR partners). Mysteriously, mere overexpression of normal RARA suffices to induce myeloid progenitor immortalization, as opposed to RARB or RARG.48  Altogether, through a variety of complementary mechanisms, retinoids and their cognate receptors exert multiple effects at various stages of hematopoietic differentiation: on self-renewal of very early progenitors (primarily RARG), mastocyte lineage commitment, and myeloid differentiation (primarily RARA; Figure 2).

Historically, one of the first models of ATRA-induced differentiation was the AML cell line HL60, which differentiates into apparently normal granulocytes. Why these cells lacking the PML-RARA fusion are so exquisitely ATRA sensitive remains unresolved. Whereas ATRA-resistant HL60 cells harbor RARA mutations, demonstrating that RARA is the actual effector of ATRA activity in that setting, the key downstream targets remain largely unknown. Multiple studies have since demonstrated that, similar to normal progenitors, growth and differentiation of some primary AML blasts are, at least in part, ATRA- or RARA-agonist sensitive.49,50  Similar findings were made with ligands specific for RXRs alone, or in the presence of cAMP which allows transcriptional activation from RAR/RXR heterodimers.51,52 

Most, if not all, of these initial studies used morphological differentiation as an end point. For decades, the field considered that terminal differentiation necessarily entailed loss of clonogenic activity. Yet, recent studies in APL or AML models have demonstrated that cancer cells profoundly engaged in the differentiation pathway can nevertheless revert to immature cells and regain clonogenic features.53-56  These provocative observations cast some caution on conclusions drawn from many previous findings. Thus, rather than terminal granulocytic differentiation, ATRA-induced loss of clonogenic activity or clear-cut prolongation of survival in vivo are the end points that are more likely to adequately reflect durable therapeutic efficacy in patients with AML. Although loss of leukemia-initiating cell activity may be envisioned as a specific form of differentiation, it is mechanistically distinct and may be uncoupled from terminal granulocytic maturation.55  While there are examples where retinoids modestly prolong survival of AML models or patient-derived xenografts,50  AML remission, not to mention eradication, is never achieved by ATRA alone, in sharp contrast to APL.

The natural vitamin A derivative used in most studies is ATRA, which activates all 3 RARs isoforms (although with different efficiencies; Figure 1). Receptor selectivity of drugs provides an interesting angle to approach their respective role. Intriguingly, a study found opposing effects of ATRA and a RARA-specific agonist for clonogenic growth of AML/ETO-transformed progenitors, suggestive of antagonist functions of RARs α and γ isoforms.57  The results of these experiments suggest that, in this model, ligand activation of RARA signaling efficiently promotes myeloid differentiation and loss of self-renewal, whereas ATRA, through RARG activation, may enhance self-renewal44  and/or decrease retinoid levels by activating cytochromes implicated in their catabolism.34,35  However, the respective roles of RARA and RARG in AML terminal differentiation and clonogenic activity remain open areas of investigation. Recent studies have explored tamibarotene, a RARA/RARB-specific agonist with a better pharmacodynamic profile than ATRA (in particular, a longer half-life).58  Tamibarotene, also known as SY-1425, can induce differentiation and initiate clinical responses in some clinically ATRA-resistant patients with APL, demonstrating its biological activity.59  When used in maintenance therapy after ATRA/chemotherapy APL treatment, it is slightly more efficient than ATRA in preventing late relapses,60  particularly in patients with FLT3 mutations, which antagonize ATRA signaling.19,61  Tamibarotene can also efficiently promote differentiation of some AML cells ex vivo and in vivo, and this effect has been linked to existence of an RARA superenhancer and monocytic differentiation.50 

Studies in ESCs have revealed that differentiation responses are intimately regulated by the cell cycle,62  possibly explaining the synergistic effects of ATRA with many drugs favoring S-phase arrest of AML cells. The modest in vivo antileukemia effects of ATRA alone (despite its ability to initiate differentiation in AML) stresses the importance of combinations with other drugs. In that respect, a broad repertoire of drugs enhances the effects of ATRA to block AML proliferation or promote differentiation (Table 1), with sometimes additive survival effects in xenograft models or even in patients. Whether this represents additive effects of parallel pathways or rather convergence of distinct pathways onto RAR signaling is currently not known. Through induction of cell cycle arrest, ATRA-induced Pin-1 inhibition may synergize with “classic” RAR activation when using pharmacological treatments with very high doses of ATRA, as in AML therapy. Importantly, synergies between retinoids and DNA hypomethylating agents have been reported recently in patients with AML.63 

Recent studies have tried to genetically define ATRA-sensitive AMLs50,64  and explore the reciprocal roles of AML oncogenes and ATRA signaling, but an unambiguous picture has not emerged as yet. Many of these correlative studies were performed in cell lines rather than in primary patient cells and were used differentiation as an end point. The first set of studies focused on NPM-1c–associated AMLs, because these initially appeared clinically to be ATRA sensitive.65,66  Similar to its role in APL, ATRA may initiate blast differentiation in these AML cells, along with activation of the TP-53 pathway and degradation of mutant NPM-1c,67,68  a protein required for maintenance of the transformed state.69  However, NPM-1c degradation may not fully recapitulate ATRA effects in theses AML cell lines (A. Bazarbachi, Beirut, Lebanon, written communication, 23 November 2020).

Other important AML oncogenes are the isocitrate dehydrogenase (IDH) enzymes, and some studies have shown correlation of ex vivo ATRA efficacy with their mutation.26,70  Several studies have suggested that reactive oxygen species (ROS) basally induced by IDH mutation facilitate ATRA-induced AML differentiation by inhibiting lysine demethylase 1A (LSD-1/KDM1A) activity. The LSD-1 histone demethylase impedes ATRA response in AML,71  as well as in some normal cells.72  Accordingly, LSD-1 inhibition favors ATRA-induced AML differentiation,71,73  a concept that is currently being investigated in patients.74  Intriguingly, mechanistic studies have demonstrated that LSD-1 inhibitors act to sensitize AML cells to ATRA-induced differentiation by impeding LSD-1 interaction with growth factor independent 1 transcriptional repressor (GFI-1),75  rather than through inactivation of its histone demethylase enzymatic activity. Thus, in AML, as in medulloblastoma,76  LSD-1 and GFI-1 cooperate to enforce transformation, but, at least in AML, LSD-1 has a scaffolding, rather than an enzymatic, function. GFI-1 also appears to be a critical factor in controlling ATRA response and granulocytic differentiation in APL.77  Overall, GFI-1 may be a master prodifferentiation transcription factor downstream of ATRA-induced, LSD-1–repressed signaling, in several different tumors, including AML.78  Pharmacological induction of ROS through iron deprivation was also proposed to promote differentiation through enhanced VitD-3 signaling79,80  or control of global sumoylation, which facilitates ATRA-induced differentiation.81 

Significant evidence, in APL and other AMLs, has suggested that FLT3 activation precludes efficient ATRA signaling and that clinically used FLT3 inhibitors can rescue efficient ATRA-induced differentiation and clinical responses.19,61,82  Other studies have linked expression of EVI1 with ATRA sensitivity.83-86  At the target gene level, EVI-1 can indeed enhance or inhibit activation of various ATRA target genes. However, EVI-1 sensitizes to ATRA-induced differentiation, whereas ATRA enhances the leukemia stem cell (LSC)-promoting activities of EVI-1, so that the overall clinical benefit of ATRA in these malignancies may be complex.64 MN1 is an oncogene in AML that regulates RARA signaling. It has been suggested that high MN1 expression is associated with ATRA responsiveness,87,88  but this remains to be further explored. Apart from direct oncogenic drivers, molecules with deregulated expression in malignancies such as PRAME89,90  were proposed to blunt ATRA response. How this relates to AML biology and resistance to ATRA signaling remains to be established.

Some trials initially suggested that ATRA could benefit AML therapy when added to standard anthracycline/cytarabine–based AML regimens.65,66  Patients with good prognoses who bear the NPM-1c mutation, but not FLT3 pathway activation, seem to benefit the most from ATRA adjunction.65,91  Unfortunately, these results were not confirmed by other groups, and so confusion prevails.92-94  Other clinical studies explored the addition of ATRA and/or valproic acid (a histone deacetylase inhibitor) to chemotherapy and observed a decrease in late relapses in elderly patients receiving ATRA and chemotherapy95,96  (discussed later). In another study, the combination of ATRA, VitD-3 and low-dose maintenance chemotherapy yielded a significant survival advantage in AML.97 

Most often, AML is a disease which occurs after 70 years of age, when many patients are unfit for cardiotoxic anthracycline-based therapies. This has raised considerable interest in nongenotoxic approaches based on “biomodulators,” often a combination of differentiation inducers with histone deacetylase inhibitors and DNA demethylating agents, all of which are predicted to facilitate re-expression of differentiation genes repressed by AML oncogenes. Combinations of ATRA and valproic acid without chemotherapy initiated ex vivo differentiation, but yielded disappointing results in patients.98,99  In contrast, when ATRA was combined with the hypomethylating agent decitabine, unambiguous enhancement of decitabine responses were observed in elderly patients with AML.63  Essentially similar results were very recently reported with tamibarotene and azacytidine in unfit patients with AML, responses being more frequent in RARA-overexpressing AMLs (registered on www.clinicaltrials.gov as #NCT02807558),100  pressing for a direct comparison between the potencies of ATRA and tamibarotene in this setting. Simultaneous or sequential drug administration schemes may also bear some importance in long-term efficacy. Mechanistically, a study suggested that hypomethylating agents and RARA-specific retinoids can synergize to induce apoptosis of AML cell lines.101  Clinically, in the absence of clear physiopathological mechanisms and simple biomarkers, identification of intrinsically ATRA-sensitive AMLs has remained an unmet challenge. Yet, the clinical results of the decitabine/retinoid combinations raise a question as to whether there is any beneficial role for the addition of retinoids (ATRA or tamibarotene) to the current decitabine/venetoclax standard for unfit patients with AML.102 

A striking observation was recently reported in 2 children morphologically and immunophenotypically diagnosed with APL and accordingly treated with a frontline ATRA-chemotherapy APL regimen. Remarkably, in these patients, the RARA gene was mutated at relapse, although extensive molecular studies failed to demonstrate any RAR fusions.103  These somatic RARA mutations acquired during ATRA therapy are known to confer dominant negative effects, similar to those found in rare forms of breast cancers.104  This unheralded, but key, clinical observation implies that heavy exposure to ATRA (both induction and maintenance regimen) and chemotherapy exerted a potent, RARA-dependent, negative selective pressure on the leukemic clone (Figure 3). These definitive observations could further renew interest for ATRA-containing regimens in defined AML subsets, as discussed earlier. Overall, this field of investigation remains active, and exciting discoveries may be expected, particularly if combination therapy for AML with ATRA and DNA demethylating agents yields unambiguous survival advantages and reliable biomarkers for ATRA sensitivity are identified.

Figure 3.

Sequential dual effects of ATRA therapy in patients with AML. (A) Selection of RARA mutations conferring ATRA resistance in ATRA-treated pediatric AML,102  fs, frameshift. (B) Proposed mechanism for ATRA activity and mutant selection in ATRA-treated AML. Treatment-selected dominant-negative RARA mutants (RARAMut) enhance self-renewal and block differentiation, contributing to AML relapses. CHEMO, anthracycline-based chemotherapy; LBD, ligand-binding domain; DBD, DNA-binding domain.

Figure 3.

Sequential dual effects of ATRA therapy in patients with AML. (A) Selection of RARA mutations conferring ATRA resistance in ATRA-treated pediatric AML,102  fs, frameshift. (B) Proposed mechanism for ATRA activity and mutant selection in ATRA-treated AML. Treatment-selected dominant-negative RARA mutants (RARAMut) enhance self-renewal and block differentiation, contributing to AML relapses. CHEMO, anthracycline-based chemotherapy; LBD, ligand-binding domain; DBD, DNA-binding domain.

Close modal

APL was considered for decades to be the premier model for differentiation therapy.54,105  Yet, it is now clear that differentiation is not the main contributor to the clinical efficacy of ATRA or arsenic and that terminal AML differentiation can spontaneously reverse.2,56  When exploring AML sensitivity to retinoids, the clinical expectations were drawn from the APL models and were very likely much too high.54  The dramatic ATRA activity in APL reflects PML/RARA degradation and requires reformation of PML nuclear bodies initially disrupted by expression of the fusion protein.106,107  Nuclear body reformation activates a PML/P53 senescence checkpoint that enforces loss of self-renewal,2  which explains the exquisite therapy sensitivity of PML/RARA–driven APLs2  and, because arsenic targets PML, the ATRA/arsenic resistance of X-RAR–driven variant APLs.1,53,108  Interestingly, in neuroblastoma as in AML, efficiency of retinoids in mouse models was boosted by decitabine.63,101,109  Decitabine may activate senescence or loss of self-renewal by activating expression of transposable elements and IFN secretion,110  a process enhanced by retinoids, and possibly implicating PML as a downstream effector.111  Thus, AML response to the retinoid/decitabine combination may somehow mechanistically resemble APL response, implicating PML, as does IFN/arsenic activity in myeloproliferative neoplasms.112  Paradoxically, it is possible that the concept of differentiation therapy actually better applies to AML than to APL. These may resemble neuroblastoma, where 13-cis-retinoic acid has significant (although modest) clinical utility as maintenance therapy, presumably by promoting differentiation and loss of self-renewal of residual neuroblastoma-derived neuronal precursors.113  This concept of retinoid-based maintenance may actually apply beyond AML and neuroblastoma. The broad effects of ATRA on growth control and LSC status may have clinical usefulness in other hematological malignancies, such as acute lymphocytic leukemia.114 

One recently reexplored axis is the interplay between ATRA and the immune system, notably IFN signaling.115  A RARA/STAT1 pathway was proposed to explain how ATRA enhances IFN target genes.116  Moreover, the ATRA-regulated RIG-1 gene controls basal IFN signaling.117  This RARA/IFN cross talk was explored in a liver cancer model resulting from loss of Tif1a, wherein mere Rara haploinsufficiency was sufficient to abrogate tumor development.118  In that context, Tif1a loss was associated with massive induction of IFN signaling in developing hepatocytes, which, similar to transformation in vivo, was reversed by loss of a single Rara allele.119  This model underscores the key, but mechanistically still poorly understood, connections between RARA and IFN-pathway activation. Accordingly, in the immune system, retinoids modulate dendritic cell differentiation120  and bacterial quorum sensing.121  Some of the ATRA effects in cancers may reflect not only cell-autonomous pathways, but also the interplay between the niche and the immune stroma (Figure 4). In that respect, tumor-derived ATRA has been proposed to orient the microenvironment toward an immune-suppressive phenotype, and RARA antagonists proposed to facilitate the activity of checkpoint inhibitors.122  Similarly, in chronic lymphocytic leukemia (CLL), leukemic cells induce ATRA synthesis in the stromal microenvironment to favor leukemia expansion.123  This emerging role of ATRA in the microenvironment may be important in the context of ATRA/chemotherapy combinations in AMLs.

Figure 4.

Multiple targeting of AML cell biology by retinoic acid.

Figure 4.

Multiple targeting of AML cell biology by retinoic acid.

Close modal

Overall, although many aspects of retinoid physiopathology remain incompletely understood, they have unexpectedly emerged in diverse aspects of medicine. As in APL, clinical breakthroughs often occur by chance and the mechanics are only later understood. The multiple facets of ATRA activity in normal hematopoiesis do not yet allow for a simple prediction of its therapeutic effects in leukemia; moreover, the emerging roles of ATRA in normal and stromal niche interactions further complicate these issues. Therapeutically, these are exciting times: after the APL explosion, followed by a window of boredom and neglect, retinoids may be revived into efficient “à la carte” combinatorial therapies based on sound physiopathological explorations.

The authors thank V. Lallemand-Breitenbach, J. Soulier, H. Dombret, E. Laplantine, A. Bazarbachi, Christopher Bassil, and laboratory members for advice and comments on the manuscript, and F. Maloumian for the artwork.

Work in the authors' laboratory is supported by INSERM, Centre National de la Recherche Scientifique (CNRS), Paris, Sciences, Lettres (PSL) University, the CAMELIA Project from Institut National du Cancer (INCA), Foundation Saint Michel, European Research Council Advanced Grant 785917–PML-THERAPY and the European Research Area Network on Translational Cancer Research (TRANSCAN) DRAMA Project.

Contribution: All authors wrote the manuscript.

Conflict-of-interest disclosure: The authors declare no competing financial interests.

Correspondence: Hugues de Thé, Collège de France, Place Marcelin Berthelot, 75005 Paris, France; e-mail: hugues.dethe@inserm.fr.

1.
Geoffroy
MC
,
de Thé
H
.
Classic and Variants APLs, as Viewed from a Therapy Response
.
Cancers (Basel)
.
2020
;
12
(
4
):
E967
.
2.
de Thé
H
,
Pandolfi
PP
,
Chen
Z
.
Acute Promyelocytic Leukemia: A Paradigm for Oncoprotein-Targeted Cure
.
Cancer Cell
.
2017
;
32
(
5
):
552
-
560
.
3.
dos Santos
GA
,
Kats
L
,
Pandolfi
PP
.
Synergy against PML-RARa: targeting transcription, proteolysis, differentiation, and self-renewal in acute promyelocytic leukemia
.
J Exp Med
.
2013
;
210
(
13
):
2793
-
2802
.
4.
Kastner
P
,
Mark
M
,
Chambon
P
.
Nonsteroid nuclear receptors: what are genetic studies telling us about their role in real life?
Cell
.
1995
;
83
(
6
):
859
-
869
.
5.
Cunningham
TJ
,
Duester
G
.
Mechanisms of retinoic acid signalling and its roles in organ and limb development
.
Nat Rev Mol Cell Biol
.
2015
;
16
(
2
):
110
-
123
.
6.
Rhinn
M
,
Dollé
P
.
Retinoic acid signalling during development
.
Development
.
2012
;
139
(
5
):
843
-
858
.
7.
Ghyselinck
NB
,
Duester
G
.
Retinoic acid signaling pathways
.
Development
.
2019
;
146
(
13
):
dev167502
.
8.
de Thé
H
,
Vivanco-Ruiz
MM
,
Tiollais
P
,
Stunnenberg
H
,
Dejean
A
.
Identification of a retinoic acid responsive element in the retinoic acid receptor beta gene
.
Nature
.
1990
;
343
(
6254
):
177
-
180
.
9.
Kashyap
V
,
Laursen
KB
,
Brenet
F
,
Viale
AJ
,
Scandura
JM
,
Gudas
LJ
.
RARγ is essential for retinoic acid induced chromatin remodeling and transcriptional activation in embryonic stem cells
.
J Cell Sci
.
2013
;
126
(
pt 4
):
999
-
1008
.
10.
Zhu
J
,
Gianni
M
,
Kopf
E
, et al
.
Retinoic acid induces proteasome-dependent degradation of retinoic acid receptor alpha (RARalpha) and oncogenic RARalpha fusion proteins
.
Proc Natl Acad Sci USA
.
1999
;
96
(
26
):
14807
-
14812
.
11.
Koken
MHM
,
Daniel
M-T
,
Gianni
M
, et al
.
Retinoic acid, but not arsenic trioxide, degrades the PLZF/RARalpha fusion protein, without inducing terminal differentiation or apoptosis, in a RA-therapy resistant t(11;17)(q23;q21) APL patient
.
Oncogene
.
1999
;
18
(
4
):
1113
-
1118
.
12.
Giannì
M
,
Bauer
A
,
Garattini
E
,
Chambon
P
,
Rochette-Egly
C
.
Phosphorylation by p38MAPK and recruitment of SUG-1 are required for RA-induced RAR gamma degradation and transactivation
.
EMBO J
.
2002
;
21
(
14
):
3760
-
3769
.
13.
Al Tanoury
Z
,
Gaouar
S
,
Piskunov
A
, et al
.
Phosphorylation of the retinoic acid receptor RARγ2 is crucial for the neuronal differentiation of mouse embryonic stem cells
.
J Cell Sci
.
2014
;
127
(
pt 9
):
2095
-
2105
.
14.
Weston
AD
,
Blumberg
B
,
Underhill
TM
.
Active repression by unliganded retinoid receptors in development: less is sometimes more
.
J Cell Biol
.
2003
;
161
(
2
):
223
-
228
.
15.
Niu
H
,
Fujiwara
H
,
di Martino
O
, et al
.
Endogenous retinoid X receptor ligands in mouse hematopoietic cells
.
Sci Signal
.
2017
;
10
(
503
):
eaan1011
.
16.
Paubelle
E
,
Zylbersztejn
F
,
Maciel
TT
, et al
.
Vitamin D Receptor Controls Cell Stemness in Acute Myeloid Leukemia and in Normal Bone Marrow
.
Cell Rep
.
2020
;
30
(
3
):
739
-
754.e4
.
17.
Piskunov
A
,
Rochette-Egly
C
.
A retinoic acid receptor RARα pool present in membrane lipid rafts forms complexes with G protein αQ to activate p38MAPK
.
Oncogene
.
2012
;
31
(
28
):
3333
-
3345
.
18.
Shao
X
,
Liu
Y
,
Li
Y
, et al
.
The HER2 inhibitor TAK165 Sensitizes Human Acute Myeloid Leukemia Cells to Retinoic Acid-Induced Myeloid Differentiation by activating MEK/ERK mediated RARα/STAT1 axis
.
Sci Rep
.
2016
;
6
(
1
):
24589
.
19.
Ma
HS
,
Greenblatt
SM
,
Shirley
CM
, et al
.
All-trans retinoic acid synergizes with FLT3 inhibition to eliminate FLT3/ITD+ leukemia stem cells in vitro and in vivo
.
Blood
.
2016
;
127
(
23
):
2867
-
2878
.
20.
Gianni
M
,
Peviani
M
,
Bruck
N
, et al
.
p38αMAPK interacts with and inhibits RARα: suppression of the kinase enhances the therapeutic activity of retinoids in acute myeloid leukemia cells
.
Leukemia
.
2012
;
26
(
8
):
1850
-
1861
.
21.
Glasow
A
,
Prodromou
N
,
Xu
K
,
von Lindern
M
,
Zelent
A
.
Retinoids and myelomonocytic growth factors cooperatively activate RARA and induce human myeloid leukemia cell differentiation via MAP kinase pathways
.
Blood
.
2005
;
105
(
1
):
341
-
349
.
22.
Bruck
N
,
Vitoux
D
,
Ferry
C
, et al
.
A coordinated phosphorylation cascade initiated by p38MAPK/MSK1 directs RARalpha to target promoters
.
EMBO J
.
2009
;
28
(
1
):
34
-
47
.
23.
Bour
G
,
Gaillard
E
,
Bruck
N
, et al
.
Cyclin H binding to the RARalpha activation function (AF)-2 domain directs phosphorylation of the AF-1 domain by cyclin-dependent kinase 7
.
Proc Natl Acad Sci USA
.
2005
;
102
(
46
):
16608
-
16613
.
24.
Wei
S
,
Kozono
S
,
Kats
L
, et al
.
Active Pin1 is a key target of all-trans retinoic acid in acute promyelocytic leukemia and breast cancer
.
Nat Med
.
2015
;
21
(
5
):
457
-
466
.
25.
Kozono
S
,
Lin
YM
,
Seo
HS
, et al
.
Arsenic targets Pin1 and cooperates with retinoic acid to inhibit cancer-driving pathways and tumor-initiating cells
.
Nat Commun
.
2018
;
9
(
1
):
3069
.
26.
Mugoni
V
,
Panella
R
,
Cheloni
G
, et al
.
Vulnerabilities in mEND POINT2 AML confer sensitivity to APL-like targeted combination therapy
.
Cell Res
.
2019
;
29
(
6
):
446
-
459
.
27.
Lian
X
,
Lin
YM
,
Kozono
S
, et al
.
Pin1 inhibition exerts potent activity against acute myeloid leukemia through blocking multiple cancer-driving pathways [published correction appears in J Hematol Oncol. 2018;11(1):94]
.
J Hematol Oncol
.
2018
;
11
(
1
):
73
.
28.
Kastner
P
,
Lawrence
HJ
,
Waltzinger
C
,
Ghyselinck
NB
,
Chambon
P
,
Chan
S
.
Positive and negative regulation of granulopoiesis by endogenous RARalpha
.
Blood
.
2001
;
97
(
5
):
1314
-
1320
.
29.
Friesen
LR
,
Gu
B
,
Kim
CH
.
A ligand-independent fast function of RARalpha promotes exit from metabolic quiescence upon T cell activation and controls T cell differentiation
.
Mucosal Immunol
.
2021
;
14
(
1
):
100
-
112
.
30.
Chanda
B
,
Ditadi
A
,
Iscove
NN
,
Keller
G
.
Retinoic acid signaling is essential for embryonic hematopoietic stem cell development
.
Cell
.
2013
;
155
(
1
):
215
-
227
.
31.
Cabezas-Wallscheid
N
,
Buettner
F
,
Sommerkamp
P
, et al
.
Vitamin A-Retinoic Acid Signaling Regulates Hematopoietic Stem Cell Dormancy
.
Cell
.
2017
;
169
(
5
):
807
-
823.e19
.
32.
Purton
LE
,
Bernstein
ID
,
Collins
SJ
.
All-trans retinoic acid enhances the long-term repopulating activity of cultured hematopoietic stem cells
.
Blood
.
2000
;
95
(
2
):
470
-
477
.
33.
Kuwata
T
,
Wang
IM
,
Tamura
T
, et al
.
Vitamin A deficiency in mice causes a systemic expansion of myeloid cells
.
Blood
.
2000
;
95
(
11
):
3349
-
3356
.
34.
Ghiaur
G
,
Yegnasubramanian
S
,
Perkins
B
,
Gucwa
JL
,
Gerber
JM
,
Jones
RJ
.
Regulation of human hematopoietic stem cell self-renewal by the microenvironment’s control of retinoic acid signaling
.
Proc Natl Acad Sci USA
.
2013
;
110
(
40
):
16121
-
16126
.
35.
Hernandez
D
,
Palau
L
,
Norsworthy
K
, et al
.
Overcoming microenvironment-mediated protection from ATRA using CYP26-resistant retinoids
.
Leukemia
.
2020
;
34
(
11
):
3077
-
3081
.
36.
Collins
SJ
.
The role of retinoids and retinoic acid receptors in normal hematopoiesis
.
Leukemia
.
2002
;
16
(
10
):
1896
-
1905
.
37.
Hjertson
M
,
Kivinen
PK
,
Dimberg
L
,
Nilsson
K
,
Harvima
IT
,
Nilsson
G
.
Retinoic acid inhibits in vitro development of mast cells but has no marked effect on mature human skin tryptase- and chymase-positive mast cells
.
J Invest Dermatol
.
2003
;
120
(
2
):
239
-
245
.
38.
Zhu
J
,
Heyworth
CM
,
Glasow
A
, et al
.
Lineage restriction of the RARalpha gene expression in myeloid differentiation
.
Blood
.
2001
;
98
(
8
):
2563
-
2567
.
39.
Li
L
,
Qi
X
,
Sun
W
, et al
.
Am80-GCSF synergizes myeloid expansion and differentiation to generate functional neutrophils that reduce neutropenia-associated infection and mortality
.
EMBO Mol Med
.
2016
;
8
(
11
):
1340
-
1359
.
40.
Ding
W
,
Shimada
H
,
Li
L
, et al
.
Retinoid agonist Am80-enhanced neutrophil bactericidal activity arising from granulopoiesis in vitro and in a neutropenic mouse model
.
Blood
.
2013
;
121
(
6
):
996
-
1007
.
41.
Germain
P
,
Chambon
P
,
Eichele
G
, et al
.
International Union of Pharmacology. LX. Retinoic acid receptors
.
Pharmacol Rev
.
2006
;
58
(
4
):
712
-
725
.
42.
Beard
RL
,
Duong
TT
,
Teng
M
,
Klein
ES
,
Standevan
AM
,
Chandraratna
RA
.
Synthesis and biological activity of retinoic acid receptor-alpha specific amides
.
Bioorg Med Chem Lett
.
2002
;
12
(
21
):
3145
-
3148
.
43.
Purton
LE
,
Dworkin
S
,
Olsen
GH
, et al
.
RARgamma is critical for maintaining a balance between hematopoietic stem cell self-renewal and differentiation
.
J Exp Med
.
2006
;
203
(
5
):
1283
-
1293
.
44.
Walkley
CR
,
Olsen
GH
,
Dworkin
S
, et al
.
A microenvironment-induced myeloproliferative syndrome caused by retinoic acid receptor gamma deficiency
.
Cell
.
2007
;
129
(
6
):
1097
-
1110
.
45.
Wang
W
,
Yang
J
,
Liu
H
, et al
.
Rapid and efficient reprogramming of somatic cells to induced pluripotent stem cells by retinoic acid receptor gamma and liver receptor homolog 1
.
Proc Natl Acad Sci USA
.
2011
;
108
(
45
):
18283
-
18288
.
46.
Christidi
E
,
Huang
H
,
Shafaattalab
S
, et al
.
Variation in RARG increases susceptibility to doxorubicin-induced cardiotoxicity in patient specific induced pluripotent stem cell-derived cardiomyocytes
.
Sci Rep
.
2020
;
10
(
1
):
10363
.
47.
Aminkeng
F
,
Bhavsar
AP
,
Visscher
H
, et al;
Canadian Pharmacogenomics Network for Drug Safety Consortium
.
A coding variant in RARG confers susceptibility to anthracycline-induced cardiotoxicity in childhood cancer
.
Nat Genet
.
2015
;
47
(
9
):
1079
-
1084
.
48.
Du
C
,
Redner
RL
,
Cooke
MP
,
Lavau
C
.
Overexpression of wild-type retinoic acid receptor alpha (RARalpha) recapitulates retinoic acid-sensitive transformation of primary myeloid progenitors by acute promyelocytic leukemia RARalpha-fusion genes
.
Blood
.
1999
;
94
(
2
):
793
-
802
.
49.
van Gils
N
,
Verhagen
HJMP
,
Smit
L
.
Reprogramming acute myeloid leukemia into sensitivity for retinoic-acid-driven differentiation
.
Exp Hematol
.
2017
;
52
:
12
-
23
.
50.
McKeown
MR
,
Corces
MR
,
Eaton
ML
, et al
.
Superenhancer Analysis Defines Novel Epigenomic Subtypes of Non-APL AML, Including an RARα Dependency Targetable by SY-1425, a Potent and Selective RARα Agonist
.
Cancer Discov
.
2017
;
7
(
10
):
1136
-
1153
.
51.
Kamashev
DE
,
Vitoux
D
,
De Thé
H
.
PML–RARA-RXR oligomers mediate retinoid- and rexinoid/cAMP cross-talk in acute promyelocytic leukecell differentiation
.
J Exp Med
.
2004
;
199
(
8
):
1163
-
1174
.
52.
Altucci
L
,
Rossin
A
,
Hirsch
O
, et al
.
Rexinoid-Triggered Differentiation and Tumor-Selective Apoptosis of AML by Protein Kinase A–Mediated Desubordination of Retinoid X Receptor
.
Cancer Res
.
2005
;
65
:
8754
-
8765
.
53.
Nasr
R
,
Guillemin
MC
,
Ferhi
O
, et al
.
Eradication of acute promyelocytic leukemia-initiating cells through PML-RARA degradation [published correction appears in Nat Med. 2009 Jan;15(1):117]
.
Nat Med
.
2008
;
14
(
12
):
1333
-
1342
.
54.
de Thé
H
.
Differentiation therapy revisited
.
Nat Rev Cancer
.
2018
;
18
(
2
):
117
-
127
.
55.
Ablain
J
,
Leiva
M
,
Peres
L
,
Fonsart
J
,
Anthony
E
,
de Thé
H
.
Uncoupling RARA transcriptional activation and degradation clarifies the bases for APL response to therapies
.
J Exp Med
.
2013
;
210
(
4
):
647
-
653
.
56.
McKenzie
MD
,
Ghisi
M
,
Oxley
EP
, et al
.
Interconversion between Tumorigenic and Differentiated States in Acute Myeloid Leukemia
.
Cell Stem Cell
.
2019
;
25
(
2
):
258
-
272.e9
.
57.
Chee
LC
,
Hendy
J
,
Purton
LE
,
McArthur
GA
.
ATRA and the specific RARα agonist, NRX195183, have opposing effects on the clonogenicity of pre-leukemic murine AML1-ETO bone marrow cells
.
Leukemia
.
2013
;
27
(
6
):
1369
-
1380
.
58.
Miwako
I
,
Kagechika
H
.
Tamibarotene
.
Drugs Today (Barc)
.
2007
;
43
(
8
):
563
-
568
.
59.
Sanford
D
,
Lo-Coco
F
,
Sanz
MA
, et al
.
Tamibarotene in patients with acute promyelocytic leukaemia relapsing after treatment with all-trans retinoic acid and arsenic trioxide
.
Br J Haematol
.
2015
;
171
(
4
):
471
-
477
.
60.
Takeshita
A
,
Asou
N
,
Atsuta
Y
, et al;
the Japanese Adult Leukemia Study Group
.
Tamibarotene maintenance improved relapse-free survival of acute promyelocytic leukemia: a final result of prospective, randomized, JALSG-APL204 study
.
Leukemia
.
2019
;
33
(
2
):
358
-
370
.
61.
Esnault
C
,
Rahmé
R
,
Rice
KL
, et al
.
FLT3-ITD impedes retinoic acid, but not arsenic, responses in murine acute promyelocytic leukemias
.
Blood
.
2019
;
133
(
13
):
1495
-
1506
.
62.
Pauklin
S
,
Vallier
L
.
The Cell-Cycle State of Stem Cells Determines Cell Fate Propensity [published correction appears in 2014;156(6):1338]
.
Cell
.
2013
;
155
(
1
):
135
-
147
.
63.
Lübbert
M
,
Grishina
O
,
Schmoor
C
, et al;
DECIDER Study Team
.
Valproate and Retinoic Acid in Combination With Decitabine in Elderly Nonfit Patients With Acute Myeloid Leukemia: Results of a Multicenter, Randomized, 2 × 2, Phase II Trial
.
J Clin Oncol
.
2020
;
38
(
3
):
257
-
270
.
64.
Nguyen
CH
,
Grandits
AM
,
Purton
LE
,
Sill
H
,
Wieser
R
.
All-trans retinoic acid in non-promyelocytic acute myeloid leukemia: driver lesion dependent effects on leukemic stem cells
.
Cell Cycle
.
2020
;
19
(
20
):
2573
-
2588
.
65.
Schlenk
RF
,
Döhner
K
,
Kneba
M
, et al;
German-Austrian AML Study Group (AMLSG)
.
Gene mutations and response to treatment with all-trans retinoic acid in elderly patients with acute myeloid leukemia. Results from the AMLSG Trial AML HD98B
.
Haematologica
.
2009
;
94
(
1
):
54
-
60
.
66.
Schlenk
RF
,
Fröhling
S
,
Hartmann
F
, et al;
AML Study Group Ulm
.
Phase III study of all-trans retinoic acid in previously untreated patients 61 years or older with acute myeloid leukemia
.
Leukemia
.
2004
;
18
(
11
):
1798
-
1803
.
67.
Martelli
MP
,
Gionfriddo
I
,
Mezzasoma
F
, et al
.
Arsenic trioxide and all-trans retinoic acid target NPM1 mutant oncoprotein levels and induce apoptosis in NPM1-mutated AML cells
.
Blood
.
2015
;
125
(
22
):
3455
-
3465
.
68.
El Hajj
H
,
Dassouki
Z
,
Berthier
C
, et al
.
Retinoic acid and arsenic trioxide trigger degradation of mutated NPM1, resulting in apoptosis of AML cells
.
Blood
.
2015
;
125
(
22
):
3447
-
3454
.
69.
Brunetti
L
,
Gundry
MC
,
Sorcini
D
, et al
.
Mutant NPM1 Maintains the Leukemic State through HOX Expression
.
Cancer Cell
.
2018
;
34
(
3
):
499
-
512.e9
.
70.
Boutzen
H
,
Saland
E
,
Larrue
C
, et al
.
Isocitrate dehydrogenase 1 mutations prime the all-trans retinoic acid myeloid differentiation pathway in acute myeloid leukemia
.
J Exp Med
.
2016
;
213
(
4
):
483
-
497
.
71.
Schenk
T
,
Chen
WC
,
Göllner
S
, et al
.
Inhibition of the LSD1 (KDM1A) demethylase reactivates the all-trans-retinoic acid differentiation pathway in acute myeloid leukemia
.
Nat Med
.
2012
;
18
(
4
):
605
-
611
.
72.
Vinckier
NK
,
Patel
NA
,
Geusz
RJ
, et al
.
LSD1-mediated enhancer silencing attenuates retinoic acid signalling during pancreatic endocrine cell development
.
Nat Commun
.
2020
;
11
(
1
):
2082
.
73.
Smitheman
KN
,
Severson
TM
,
Rajapurkar
SR
, et al
.
Lysine specific demethylase 1 inactivation enhances differentiation and promotes cytotoxic response when combined with all-trans retinoic acid in acute myeloid leukemia across subtypes
.
Haematologica
.
2019
;
104
(
6
):
1156
-
1167
.
74.
Wass
M
,
Göllner
S
,
Besenbeck
B
, et al;
Study Alliance Leukemia (SAL)
.
A proof of concept phase I/II pilot trial of LSD1 inhibition by tranylcypromine combined with ATRA in refractory/relapsed AML patients not eligible for intensive therapy
.
Leukemia
.
2021
;
35
:
701
-
711
.S
75.
Ravasio
R
,
Ceccacci
E
,
Nicosia
L
, et al
.
Targeting the scaffolding role of LSD1 (KDM1A) poises acute myeloid leukemia cells for retinoic acid-induced differentiation
.
Sci Adv
.
2020
;
6
(
15
):
eaax2746
.
76.
Lee
C
,
Rudneva
VA
,
Erkek
S
, et al
.
Lsd1 as a therapeutic target in Gfi1-activated medulloblastoma
.
Nat Commun
.
2019
;
10
(
1
):
332
.
77.
Tan
Y
,
Wang
X
,
Song
H
, et al
.
A PML/RARα direct target atlas redefines transcriptional deregulation in acute promyelocytic leukemia [published online ahead of print August 27, 2020]
.
Blood
.
2020
;
blood.2020005698
.
78.
Maiques-Diaz
A
,
Spencer
GJ
,
Lynch
JT
, et al
.
Enhancer Activation by Pharmacologic Displacement of LSD1 from GFI1 Induces Differentiation in Acute Myeloid Leukemia
.
Cell Rep
.
2018
;
22
(
13
):
3641
-
3659
.
79.
Callens
C
,
Coulon
S
,
Naudin
J
, et al
.
Targeting iron homeostasis induces cellular differentiation and synergizes with differentiating agents in acute myeloid leukemia
.
J Exp Med
.
2010
;
207
(
4
):
731
-
750
.
80.
Paubelle
E
,
Zylbersztejn
F
,
Alkhaeir
S
, et al
.
Deferasirox and vitamin D improves overall survival in elderly patients with acute myeloid leukemia after demethylating agents failure
.
PLoS One
.
2013
;
8
(
6
):
e65998
.
81.
Baik
H
,
Boulanger
M
,
Hosseini
M
, et al
.
Targeting the SUMO Pathway Primes All-trans Retinoic Acid-Induced Differentiation of Nonpromyelocytic Acute Myeloid Leukemias
.
Cancer Res
.
2018
;
78
(
10
):
2601
-
2613
.
82.
Masciarelli
S
,
Capuano
E
,
Ottone
T
, et al
.
Retinoic acid synergizes with the unfolded protein response and oxidative stress to induce cell death in FLT3-ITD+ AML
.
Blood Adv
.
2019
;
3
(
24
):
4155
-
4160
.
83.
Bingemann
SC
,
Konrad
TA
,
Wieser
R
.
Zinc finger transcription factor ecotropic viral integration site 1 is induced by all-trans retinoic acid (ATRA) and acts as a dual modulator of the ATRA response
.
FEBS J
.
2009
;
276
(
22
):
6810
-
6822
.
84.
Nguyen
CH
,
Bauer
K
,
Hackl
H
, et al
.
All-trans retinoic acid enhances, and a pan-RAR antagonist counteracts, the stem cell promoting activity of EVI1 in acute myeloid leukemia
.
Cell Death Dis
.
2019
;
10
(
12
):
944
.
85.
Steinmetz
B
,
Hackl
H
,
Slabáková
E
, et al
.
The oncogene EVI1 enhances transcriptional and biological responses of human myeloid cells to all-trans retinoic acid [published correction appears in Cell Cycle. 2015;14(11):1759]
.
Cell Cycle
.
2014
;
13
(
18
):
2931
-
2943
.
86.
Verhagen
HJ
,
Smit
MA
,
Rutten
A
, et al
.
Primary acute myeloid leukemia cells with overexpression of EVI-1 are sensitive to all-trans retinoic acid
.
Blood
.
2016
;
127
(
4
):
458
-
463
.
87.
Heuser
M
,
Argiropoulos
B
,
Kuchenbauer
F
, et al
.
MN1 overexpression induces acute myeloid leukemia in mice and predicts ATRA resistance in patients with AML
.
Blood
.
2007
;
110
(
5
):
1639
-
1647
.
88.
van Wely
KH
,
Meester-Smoor
MA
,
Janssen
MJ
,
Aarnoudse
AJ
,
Grosveld
GC
,
Zwarthoff
EC
.
The MN1-TEL myeloid leukemia-associated fusion protein has a dominant-negative effect on RAR-RXR-mediated transcription
.
Oncogene
.
2007
;
26
(
39
):
5733
-
5740
.
89.
Epping
MT
,
Wang
L
,
Edel
MJ
,
Carlée
L
,
Hernandez
M
,
Bernards
R
.
The human tumor antigen PRAME is a dominant repressor of retinoic acid receptor signaling
.
Cell
.
2005
;
122
(
6
):
835
-
847
.
90.
Oehler
VG
,
Guthrie
KA
,
Cummings
CL
, et al
.
The preferentially expressed antigen in melanoma (PRAME) inhibits myeloid differentiation in normal hematopoietic and leukemic progenitor cells
.
Blood
.
2009
;
114
(
15
):
3299
-
3308
.
91.
Schlenk
RF
,
Lübbert
M
,
Benner
A
, et al;
German-Austrian Acute Myeloid Leukemia Study Group
.
All-trans retinoic acid as adjunct to intensive treatment in younger adult patients with acute myeloid leukemia: results of the randomized AMLSG 07-04 study
.
Ann Hematol
.
2016
;
95
(
12
):
1931
-
1942
.
92.
Burnett
AK
,
Hills
RK
,
Green
C
, et al
.
The impact on outcome of the addition of all-trans retinoic acid to intensive chemotherapy in younger patients with nonacute promyelocytic acute myeloid leukemia: overall results and results in genotypic subgroups defined by mutations in NPM1, FLT3, and CEBPA
.
Blood
.
2010
;
115
(
5
):
948
-
956
.
93.
Küley-Bagheri
Y
,
Kreuzer
KA
,
Monsef
I
,
Lübbert
M
,
Skoetz
N
.
Effects of all-trans retinoic acid (ATRA) in addition to chemotherapy for adults with acute myeloid leukaemia (AML) (non-acute promyelocytic leukaemia (non-APL))
.
Cochrane Database Syst Rev
.
2018
;
8
:
CD011960
.
94.
Nazha
A
,
Bueso-Ramos
C
,
Estey
E
, et al
.
The Addition of All-Trans Retinoic Acid to Chemotherapy May Not Improve the Outcome of Patient with NPM1 Mutated Acute Myeloid Leukemia
.
Front Oncol
.
2013
;
3
:
218
.
95.
Tassara
M
,
Döhner
K
,
Brossart
P
, et al
.
Valproic acid in combination with all-trans retinoic acid and intensive therapy for acute myeloid leukemia in older patients [published correction appears in Blood. 2015 May 7;125(19):3037]
.
Blood
.
2014
;
123
(
26
):
4027
-
4036
.
96.
Di Febo
A
,
Laurenti
L
,
Falcucci
P
, et al
.
All-trans retinoic acid in association with low dose cytosine arabinoside in the treatment of acute myeoid leukemia in elderly patients
.
Am J Ther
.
2007
;
14
(
4
):
351
-
355
.
97.
Ferrero
D
,
Crisà
E
,
Marmont
F
, et al
.
Survival improvement of poor-prognosis AML/MDS patients by maintenance treatment with low-dose chemotherapy and differentiating agents
.
Ann Hematol
.
2014
;
93
(
8
):
1391
-
1400
.
98.
Bellos
F
,
Mahlknecht
U
.
Valproic acid and all-trans retinoic acid: meta-analysis of a palliative treatment regimen in AML and MDS patients
.
Onkologie
.
2008
;
31
(
11
):
629
-
633
.
99.
Raffoux
E
,
Chaibi
P
,
Dombret
H
,
Degos
L
.
Valproic acid and all-trans retinoic acid for the treatment of elderly patients with acute myeloid leukemia
.
Haematologica
.
2005
;
90
(
7
):
986
-
988
.
100.
De Botton
S
,
Cluzeau
T
,
Vigil
CE
, et al
.
SY-1425, a potent and selective RARα agonist, in combination with azacitidine demonstrates a high complete response rate and a rapid onset of response in RARA-positive newly diagnosed unfit acute myeloid leukemia. Abstract 112
.
Blood
.
2020
;
136
(
suppl
):
4
-
5
.
101.
McKeown
MR
,
Johannessen
L
,
Lee
E
,
Fiore
C
,
di Tomaso
E
.
Antitumor synergy with SY-1425, a selective RARα agonist, and hypomethylating agents in retinoic acid receptor pathway activated models of acute myeloid leukemia
.
Haematologica
.
2019
;
104
(
4
):
e138
-
e142
.
102.
DiNardo
CD
,
Jonas
BA
,
Pullarkat
V
, et al
.
Azacitidine and Venetoclax in Previously Untreated Acute Myeloid Leukemia
.
N Engl J Med
.
2020
;
383
(
7
):
617
-
629
.
103.
Zhao
J
,
Liang
JW
,
Xue
HL
, et al
.
The genetics and clinical characteristics of children morphologically diagnosed as acute promyelocytic leukemia
.
Leukemia
.
2019
;
33
(
6
):
1387
-
1399
.
104.
Tan
J
,
Ong
CK
,
Lim
WK
, et al
.
Genomic landscapes of breast fibroepithelial tumors
.
Nat Genet
.
2015
;
47
(
11
):
1341
-
1345
.
105.
Degos
L
,
Dombret
H
,
Chomienne
C
, et al
.
All-trans-retinoic acid as a differentiating agent in the treatment of acute promyelocytic leukemia
.
Blood
.
1995
;
85
(
10
):
2643
-
2653
.
106.
Daniel
M-T
,
Koken
M
,
Romagné
O
, et al
.
PML protein expression in hematopoietic and acute promyelocytic leukemia cells
.
Blood
.
1993
;
82
(
6
):
1858
-
1867
.
107.
Koken
MHM
,
Puvion-Dutilleul
F
,
Guillemin
MC
, et al
.
The t(15;17) translocation alters a nuclear body in a retinoic acid-reversible fashion
.
EMBO J
.
1994
;
13
(
5
):
1073
-
1083
.
108.
Zhu
J
,
Chen
Z
,
Lallemand-Breitenbach
V
,
de Thé
H
.
How acute promyelocytic leukaemia revived arsenic
.
Nat Rev Cancer
.
2002
;
2
(
9
):
705
-
714
.
109.
Wen
QJ
,
Yang
Q
,
Goldenson
B
, et al
.
Targeting megakaryocytic-induced fibrosis in myeloproliferative neoplasms by AURKA inhibition
.
Nat Med
.
2015
;
21
(
12
):
1473
-
1480
.
110.
Licht
JD
.
DNA Methylation Inhibitors in Cancer Therapy: The Immunity Dimension
.
Cell
.
2015
;
162
(
5
):
938
-
939
.
111.
Stadler
M
,
Chelbi-Alix
MK
,
Koken
MHM
, et al
.
Transcriptional induction of the PML growth suppressor gene by interferons is mediated through an ISRE and a GAS element
.
Oncogene
.
1995
;
11
(
12
):
2565
-
2573
.
112.
Dagher
T
,
Maslah
N
,
Edmond
V
, et al
.
JAK2V617F myeloproliferative neoplasm eradication by a novel interferon/arsenic therapy involves PML
.
J Exp Med
.
2021
;
218
(
2
):
e20201268
.
113.
Matthay
KK
,
Reynolds
CP
,
Seeger
RC
, et al
.
Long-term results for children with high-risk neuroblastoma treated on a randomized trial of myeloablative therapy followed by 13-cis-retinoic acid: a children’s oncology group study
.
J Clin Oncol
.
2009
;
27
(
7
):
1007
-
1013
.
114.
Churchman
ML
,
Low
J
,
Qu
C
, et al
.
Efficacy of Retinoids in IKZF1-Mutated BCR-ABL1 Acute Lymphoblastic Leukemia
.
Cancer Cell
.
2015
;
28
(
3
):
343
-
356
.
115.
Erkelens
MN
,
Mebius
RE
.
Retinoic Acid and Immune Homeostasis: A Balancing Act
.
Trends Immunol
.
2017
;
38
(
3
):
168
-
180
.
116.
Dimberg
A
,
Karlberg
I
,
Nilsson
K
,
Oberg
F
.
Ser727/Tyr701-phosphorylated Stat1 is required for the regulation of c-Myc, cyclins, and p27Kip1 associated with ATRA-induced G0/G1 arrest of U-937 cells
.
Blood
.
2003
;
102
(
1
):
254
-
261
.
117.
Soye
KJ
,
Trottier
C
,
Di Lenardo
TZ
, et al
.
In vitro inhibition of mumps virus by retinoids
.
Virol J
.
2013
;
10
(
1
):
337
.
118.
Khetchoumian
K
,
Teletin
M
,
Tisserand
J
, et al
.
Loss of Trim24 (Tif1alpha) gene function confers oncogenic activity to retinoic acid receptor alpha
.
Nat Genet
.
2007
;
39
(
12
):
1500
-
1506
.
119.
Tisserand
J
,
Khetchoumian
K
,
Thibault
C
,
Dembélé
D
,
Chambon
P
,
Losson
R
.
Tripartite motif 24 (Trim24/Tif1α) tumor suppressor protein is a novel negative regulator of interferon (IFN)/signal transducers and activators of transcription (STAT) signaling pathway acting through retinoic acid receptor α (Rarα) inhibition
.
J Biol Chem
.
2011
;
286
(
38
):
33369
-
33379
.
120.
Johnson
JS
,
De Veaux
N
,
Rives
AW
, et al
.
A Comprehensive Map of the Monocyte-Derived Dendritic Cell Transcriptional Network Engaged upon Innate Sensing of HIV
.
Cell Rep
.
2020
;
30
(
3
):
914
-
931.e9
.
121.
Wu
R
,
Li
X
,
Ma
N
, et al
.
Bacterial Quorum Sensing Molecules Promote Allergic Airway Inflammation by Activating the Retinoic Acid Response
.
iScience
.
2020
;
23
(
7
):
101288
.
122.
Devalaraja
S
,
To
TKJ
,
Folkert
IW
, et al
.
Tumor-Derived Retinoic Acid Regulates Intratumoral Monocyte Differentiation to Promote Immune Suppression
.
Cell
.
2020
;
180
(
6
):
1098
-
1114.e16
.
123.
Farinello
D
,
Wozińska
M
,
Lenti
E
, et al
.
A retinoic acid-dependent stroma-leukemia crosstalk promotes chronic lymphocytic leukemia progression
.
Nat Commun
.
2018
;
9
(
1
):
1787
.
124.
Kahl
M
,
Brioli
A
,
Bens
M
, et al
.
The acetyltransferase GCN5 maintains ATRA-resistance in non-APL AML [published correction appears in Leukemia. 2020;34(7):1972]
.
Leukemia
.
2019
;
33
(
11
):
2628
-
2639
.
125.
Su
M
,
Alonso
S
,
Jones
JW
, et al
.
All-Trans Retinoic Acid Activity in Acute Myeloid Leukemia: Role of Cytochrome P450 Enzyme Expression by the Microenvironment
.
PLoS One
.
2015
;
10
(
6
):
e0127790
.
Sign in via your Institution