Next Article in Journal
A Review of the Use of Hyperthermic Intraperitoneal Chemotherapy for Peritoneal Malignancy in Pediatric Patients
Previous Article in Journal
Genomic and Transcriptomic Analyses of Malignant Pleural Mesothelioma (MPM) Samples Reveal Crucial Insights for Preclinical Testing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Beyond Corticoresistance, A Paradoxical Corticosensitivity Induced by Corticosteroid Therapy in Pediatric Acute Lymphoblastic Leukemias

by
Laure Angot
1,
Pascale Schneider
1,2,
Jean-Pierre Vannier
1 and
Souleymane Abdoul-Azize
1,*
1
Normandie University, UNIROUEN, IRIB, Inserm, U1234, 76183 Rouen, France
2
Department of Pediatric Immuno-Hemato-Oncology, Rouen University Hospital, 76038 Rouen, France
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(10), 2812; https://doi.org/10.3390/cancers15102812
Submission received: 21 March 2023 / Revised: 10 May 2023 / Accepted: 16 May 2023 / Published: 18 May 2023
(This article belongs to the Section Pediatric Oncology)

Abstract

:

Simple Summary

Although remarkable progress in the treatment of acute lymphoblastic leukemia (ALL) has been observed, some patients (~20%) relapse. Resistance to therapies is a hallmark of relapses and treatment failure in ALL. Such resistances may involve different cellular mechanisms, including the modulation by therapeutic drugs of cell survival signaling pathways that may lead to therapy-induced resistance. This article will begin by providing an update of mechanisms of resistance that may lead to therapy-induced resistance in ALL. It also provides proof of concept for the therapeutic exploitation of these signaling pathways to improve treatments.

Abstract

Known as a key effector in relapse of acute lymphoblastic leukemia (ALL), resistance to drug-induced apoptosis, is tightly considered one of the main prognostic factors for the disease. ALL cells are constantly developing cellular strategies to survive and resist therapeutic drugs. Glucocorticoids (GCs) are one of the most important agents used in the treatment of ALL due to their ability to induce cell death. The mechanisms of GC resistance of ALL cells are largely unknown and intense research is currently focused on this topic. Such resistance can involve different cellular and molecular mechanisms, including the modulation of signaling pathways involved in the regulation of proliferation, apoptosis, autophagy, metabolism, epigenetic modifications and tumor suppressors. Recently, several studies point to the paradoxical role of GCs in many survival processes that may lead to therapy-induced resistance in ALL cells, which we called “paradoxical corticosensitivity”. In this review, we aim to summarize all findings on cell survival pathways paradoxically activated by GCs with an emphasis on previous and current knowledge on gene expression and signaling pathways.

1. Introduction

Acute lymphoblastic leukemias (ALL) are diseases resulting from transformations and mutations of hematopoietic cells. These mutations occur during cell development driving aberrant proliferation and survival of B and T cells [1]. Pediatric ALL developed in children until the age of 12 and present a peak prevalence between 2 and 5 years old [2]. The exact causes remain unclear, but some genetic, immunologic, viral and environmental factors seem to be implicated. Less than 5% of cases have been associated with inherited predisposing genetic syndromes such as Down’s syndrome, Bloom’s syndrome, ataxia-telangiectasia or Nijmegen breakage syndrome or with ionizing radiation [3,4]. ALL is the most common childhood malignancy with 80% of childhood leukemias and 25% of all childhood cancers [3]. Huge strides have been made over the past 50 years in the understanding of management of childhood ALL resulting in improvement in cure rates from approximately 10% to approximately 90% [3,5,6,7]. It has been shown that 85% of the cases of childhood ALL are of the B–lineage [3,8]. At the same time, T–ALL accounts for 10%–15% of the cases of childhood ALL and the outcome is more severe [9,10].

2. Glucocorticoids in Leukemia Treatment

Glucocorticoids (GCs) are small compounds derived from cortisol which is produced by the adrenal gland. Upon binding with GC, the glucocorticoid receptor, GR, encoded by the NR3C1 gene, dimerizes and migrates to the nucleus where it functions as a transcription factor and regulates the expression of multiple steroid-response genes [11,12]. GCs are involved in many biological processes including metabolism, development, differentiation and neural activity [13]. Because of their plural activity capacity, they became used as therapeutic agents in the treatment of many diseases. For example, GCs are used as anti-inflammatory or anti-proliferative treatments in immunopathies or leukemias, respectively. [13]. Nowadays, ALL patient care is based on chemotherapy protocols, in which GC take an important place (Table 1 and Table 2) [14]. At the beginning, prednisone was used in treatment. Protocols consisted of 4 weeks of prednisone in association with other drugs [14]. In recent years, dexamethasone, a synthetic analog of cortisol, which differs molecularly from prednisone by several atoms, has been increasingly used to treat ALL for its superior efficacy (for review see ref. [14]). For example, dexamethasone has an additional fluorine atom and an additional methyl group. These changes slow the metabolism of dexamethasone and lead to an extension of plasma half-life (200 min vs. 60 min for prednisolone) and biological half-life (36–54 h vs. 24–36 h) [14]. Nevertheless, while dexamethasone has increased efficiency, its cytotoxic effects are much more important than those of prednisone [14] and they should be considered.
Furthermore, the essential role of endogenous GC in normal cell physiology has been demonstrated, in particular via the interaction of the GR with the T cell receptor (TCR) to modulate T cell development [23,24] and homeostasis [25].

3. Glucocorticoids and Leukemic Cell Death: Mechanisms

GCs have a cytotoxic effect by binding to GR in the cytoplasm. Binding of GC to the GR triggers the dissociation of proteins bound to the receptor such as hsp and BAG-1 [26]. This activates the nuclear localization signal (NLS) domains of the receptor. After that, receptors are dimerized and translocate to the nucleus where they can interact with glucocorticoid-response elements, GRE. This interaction leads to the activation, inactivation or modulation of responsive genes and repress mostly the transcriptional activation of the activating protein-1 (AP-1), the nuclear factor (NF-kB) and even the GR itself [13]. Receptors can also stay in monomeric form and repress directly, by interaction, the activity of transcription factors AP-1 and NF-kB. The inhibition of these pathways induced cell-cycle arrest and apoptosis [13,14,27]. The pro-apoptotic response following GC binding to their receptors depends on their capacity to induce transcription of BCL2L11 which encodes the pro-apoptotic BH3-only factor BIM [28].
Two major apoptosis pathways can be distinguished: an intrinsic or mitochondrial pathway characterized by disruption of membrane potential, release of cytochrome c and activation of caspases [29] and pathways induced by membrane death receptors, such as TNFR1 or Fas, which bind their ligand (TNF-α, FasL etc.) and recruit a series of downstream factors, such as caspase 8, which leads to cell death. Importantly, these two pathways can be activated or modulated by GC [29]. In addition, GC might induce apoptosis indirectly by gene deregulation, which drives distress and cellular damage, such as production of oxygen radicals, alteration of metabolic pathways and Ca2+ fluxes [29]. However, many key components that associate Ca2+ signaling in GC sensitivity are not fully understood, as discussed below.

4. Glucocorticoids Stimulate Leukemic Cell Resistance: The Paradox

Despite its strong anti-inflammatory capacity, GC therapy is limited by two major drawbacks. First, GCs are well-known to be associated with side effects, particularly when it is administrated in high doses for long time periods. The toxic effects of the major GCs (prednisone and dexamethasone) are still investigated in ALL patients but are most often reported with dexamethasone causing sides effects such as myopathy, obesity or bone fracture [14].
In the meantime, some patients are refractory to the therapy and become GC-resistant. Yet, the mechanistic basis of GC resistance remain elusive [30]. Resistance can either be inherited, mostly via mutations [30] in the GR, NR3C1 gene [31] or other loss-of-functions mutations in the GR [32,33]. Resistance can also be induced after relapse; it has been shown that relapsed leukemia cells are more resistant to GC [34]. In addition to this secondary resistance observed during relapses, treatment of ALL by GCs is also limited by primary resistance, i.e., from the upfront GC treatment [35,36], and this initial response to GC therapy is a major predictor of response to chemotherapy and long-term B-ALL patient outcomes.
Theoretically, resistance to GC treatment is manifested by the absence of a cell response (i.e., no cell death) or by genetic changes involved in the induction of cell death. We believe that this view is incomplete; indeed, the absence of cell death in response to GC treatment does not mean that the cells do not respond to GC. Here, we rather propose a “paradoxical corticosensitivity” because it turns out that, in addition to the absence of response to GC treatment, this treatment also activates cell survival pathways, contrary to the expected effect. Thus, this underlines a paradox in which GCs stimulate leukemic cell growth and resistance, rather than induce cell death. The activation of one or several pathways could be the cause of GC resistance in ALL. These pathways can act directly on the GR function, or indirectly by interfering with the GC-induced apoptosis signaling (Figure 1). A few pathways that play a role in GC-stimulated tumor ALL cell survival (Table 3) have been highlighted.

4.1. Calcium Signaling

Ca2+ is a versatile secondary messenger that regulates many cellular functions (cell proliferation, invasiveness, angiogenesis, migration and metastasis) by activating or inhibiting a variety of signaling pathways through Ca2+-dependent proteins [53,54,55]. Several reports have shown transient increases in intracellular Ca2+ signaling in multiple models after dexamethasone administration [37,38,56,57], but many key elements of that association are not fully understood, especially in the sensitivity or resistance of target cells to GC. It would be tempting to consider that this Ca2+ signaling, induced after stimulation of GC, could be involved in the destruction of cells treated by GC. However, this view would be reductive with respect to Ca2+ signaling, a very early event that occurs in the first seconds or even minutes after stimulation of cells with GC. As a second messenger, Ca2+ represents a key regulator in survival and cell death. Indeed, we previously reported evidence that chelation of this intracellular Ca2+ mobilization would increase the sensitivity of leukemic cells to GC. This suggests its involvement in a cell survival pathway that is therefore resistant to the devastating effect of GC. Furthermore, we and others have provided evidence that Ca2+ signaling might induce GC resistance in ALL. In fact, GC-induced Ca2+ increases are greater in resistant cells compared to sensitive cells [37,56].An additional pathway by which GC induces its own resistance is through serum GC-inducible kinase-1 (SGK1). In addition to promoting tumor growth, SGK1 signaling contributes to GC resistance [58,59], consistent with its aberrant upregulation found in GC-resistant ALL [60]. Remarkably, SGK1 activates several Ca2+ channels and modulators, including Orai1 and Stim [61,62], the main Ca2+ pathway involved in lymphocyte activation. SGK1 phosphorylates the protein ubiquitin ligase Nedd4-2, then binds with the protein 14-3-3, and blocks its ability to ubiquinate Orai-1 [62], resulting in increased Orai1 activity, which confers survival of tumor cells [63]. The voltage-sensitive K+ channel, Kv1.3, underlies sustained Ca2+ entry via Orai1/STIM1 caused by membrane repolarization [64,65]. Kv1.3 is robustly expressed in ALL [66,67] and upregulated by SGK signaling [68]. Given these observations, a possible hypothesis is that SGK1-mediated activation of Kv1.3 by GC may reflect a mechanism of GC resistance in ALL via a promotion of Ca2+ entry [65].

4.2. IL-7 Dependent Pathway

IL-7 is a critical actor in the ALL microenvironment. In fact, it promotes the survival and differentiation of thymocytes and the IL-7 receptor/ JAK/STAT5 (IL-7R/JAK/STAT5) signaling pathway contributes to ALL pathogenesis [69,70]. IL-7 positively regulates the expression of the anti-apoptotic gene, BCL2, which promotes thymocyte survival. Conversely, GCs are known to downregulate expression of BCL2 [29,71] leading to pro-apoptotic factors. Recently, At the mechanistic level, the relationship between steroid resistance and IL7R-driven cell survival has recently been proposed in ALL patients [40,72]. Thus, it was shown that GCs elicit their own resistance [39] in the presence of IL-7, by promoting upregulation of IL-7 receptor (IL-7R) expression, resulting in increased STAT5 activation, which consequentially enhances the upregulation of the pro-survival gene BCL-2 [39,40,73,74,75] and the PIM1 kinase gene [76,77]. In addition to the downstream activation of the PI3K-AKT and STAT5 pathways, IL7R signaling also results in the activation of MAPK-ERK signaling [72,78,79,80,81,82]. It has been demonstrated that these signaling pathways can individually contribute to steroid resistance in pediatric ALL via different mechanisms. Activation of STAT5 promotes cell survival through a Bcl-2-independent mechanism [69], instead of activating Bcl-2, STAT5 can activate the PI3K/AKT/mTOR intracellular signaling pathway and leads to leukemia progression [69,70]. As recently suggested, upon GC treatment, strong upregulation of BIM, a pro-apoptotic target gene of GR, may counteract GR/STAT5-induced BCL-XL and BCL2 activation downstream of IL-7-induced or pathogenic IL7R signaling, [40,83]. Consistent with this, recent studies show that JAK-STAT overexpression inhibits GC hypersensitivity by increasing Bcl-2 transcription in ALL cells [84].

4.3. PI3K/AKT/mTOR Signaling

The phosphoinositide 3-kinase (PI3K)/Akt pathway plays a huge role in integration of survival signals including cell metabolism, cell cycle progression and proliferation [85]. PI3Ks transduce signals subsequent to growth factors and cytokines; they transform these signals into intracellular messages by generating phosphatidylinositol-3,4,5 trisphosphate (PIP3) from phosphatidylinositol-3,4 bisphosphate (PIP2) at the membrane. PIP3 activates the phosphoinositide-dependent kinase-1 (PDK-1) and the serine–threonine protein kinase (AKT) [86]. Following activation, AKT controls expression of several apoptotic genes. For example, AKT inhibits a pro-apoptotic Bcl-2 family member (BAD)-mediated apoptosis [87,88], prevents XIAP (an anti-apoptotic factor) autoubiquitination and degradation and drives cell survival [87]. AKT activation may result in an increased metabolism which can antagonize the metabolic inhibition induced by GC [89]. mTOR activation by AKT compromises GC-induced apoptosis by increasing the expression of anti-apoptotic MCL1, a Bcl-2 family member [1]. It was demonstrated that FoxO transcription factors can regulate lymphocyte apoptosis and, therefore, play a huge role in GC-induced apoptosis in these cells [48]. BIM is usually induced after GC treatment of cells but an abnormal activation of Akt results in an inhibition of the FoxO3a/Bim pathway leading to downregulation of BIM expression and a defective GC-induced transcription [48]. C-MYC oncogene activation has also been shown to be a signaling pathway associated with GC resistance [90].
In ALL, stimulation of tumor cell viability and proliferation, resulting from constitutive activation of the PI3K pathway, has been shown to counteract drug sensitivity in vitro [81,91]. Interestingly, it has been demonstrated that activation of the GR by GC promotes its interaction with PI3K, resulting in TNF-α activation [92], which, by activating the nuclear factor kappa B (NF-KB), promotes ALL cell development and survival [93]. Disruption of GR/PI3K interaction reduces GC effects, suggesting that some functions regulated by GR might occur through kinase interaction [92]. Furthermore, as a determinant of GC resistance, the AKT pathway induces GR phosphorylation (non-functional form) at S134 to prevent its nuclear translocation and result in blockage of transcriptional regulation of GC targets genes [11]. The fact that AKT signaling inhibits GR nuclear translocation following GC treatment indicates that GCs paradoxically activate AKT to regulate GR phosphorylation and prevent GC-mediated cell death [11]. Indeed, functional studies have shown that AKT inhibition facilitates GC-mediated translocation of GR to the nucleus, leading to an absence of GC resistance in vitro and in vivo [11]. Moreover, by activating AKT, GC leads to FoxO3a inhibition [94], activity which is required for GC-induced cell death [95], resulting in its own resistance [48]. In another way, as described above, upregulation of IL-7R pathway by GC [39], in turn, via PI3K stimulation, leads to the activation of Akt/mTOR axes, involved in ALL cell viability, growth, survival and proliferation [80,96,97].

4.4. MAPK Signaling Pathway

Studies about crosstalk between the GC and MAPK pathways show a direct role for MAPK for GR signaling [25]. ERK, JNK and p38 are three of the major classes of MAPK [49]. ERK and JNK seems to protect cells against GC-dependent apoptosis. On the other hand, p38 promoted GC-mediated apoptosis. In fact, p38 could phosphorylate GR on serine 211, resulting in enhanced transcriptional and apoptotic activity [49]. Regarding the ERK and JNK pathways, a GC-resistant T cell line, CEM-C1-15, shows high ERK/JNK phosphorylation/activity after GC exposure, whose inhibition confers a Dex-sensitive phenotype [49,50], consistent with the hypothesis that ERK/JNK induce a paradoxical protective effect against Dex-dependent apoptosis [50,98]. Moreover, MEK1 and MEK2 are known to phosphorylate ERK and enhance this activity, leading to cell survival [99]. In T-ALL cells, IL-7 induces activation of the MEK-ERK pathway [80]. Thus, GC, by upregulating IL-7 signaling through IL-7R [44], leads to activation of the MEK/ERK pathway, a cell survival pathway [78,82]. Also, Ca2+ signaling is known to phosphorylate ERK [100]. We showed previously that GCs induce their own resistance in B-ALL cells by paradoxically eliciting a Ca2+ signaling-mediated pro-survival process that results in activation of ERK pathway. Thus, this activation counteracts the negative action of GCs in B-ALL [37].

4.5. Lck Signaling Pathway

Lck is a proto-oncogene, which was found to be overexpressed and hyperactivated in both T-ALL [2] and B-ALL [45,101], where it plays an essential role in cell survival, proliferation and activation [102]. Consistent with this finding, a correlation has been observed between higher basal Lck expression and poor clinical response to GCs in ALL, as measured by phosphoproteomic profiling. This analysis identified that newly diagnosed pediatric ALL patients with a poor response to prednisone tended to have higher Lck expression than did patients with a good response to prednisone [2,103]. Moreover, supporting these observations, Lck hyperactivation has recently been reported in early pediatric T-ALL patients with poor response to initial GC therapy [104]. These findings suggest that LCK-driven oncogenic signaling may be aberrantly activated in ALL, and thus, may play a role in GC resistance. Interestingly, Dex treatment induces Lck phosphorylation at Y394, resulting in a profound activation of several downstream mediators (such as Fyn and ZAP70 kinases) in Jurkat T-ALL cells [44,105]. Maximal Lck phosphorylation occurs between 5 and 15 min after Dex exposure, suggesting a transcription-independent mechanism [44]. Similarly, Dex treatment appears to increase the enzymatic activity of LCK by decreasing its phosphorylation in Y505, resulting in phosphorylation of ZAP70 [44,46].
Other signaling pathways such as IL-7R and PAX5 activated by GC represent additional LCK-mediated processes that may lead to therapy-induced resistance. Indeed, LCK activation drives IL-7R signaling, a paradoxical resistance pathway activated by GC [39], through STAT5 activation and its downstream target genes cMYC and CCND2. Furthermore, STAT5 signaling is attenuated by LCK inhibition in PAX5 translocated BCP-ALL patients [101]. PAX5 alterations have been frequently detected in 30% of ALL patients [45,101], and cases with PAX5 have been found to have a specific driver activity signature and LCK hyperactivation [45]. This is supported by the observation that Dex exposure increases PAX5 mRNA expression in ALL cells [50]. This activation of LCK signaling by Dex was surprising and paradoxical, considering the established role of Dex as therapy and inductor of cell death in ALL. These findings were confirmed in patients with a good response to prednisone, whose cells showed that TCR engagement-mediated LCK activation was able to limit drug sensitivity. [2]. Furthermore, improvements in GC sensitivity or reversal of GC resistance has been observed following LCK inhibition, both in cell lines and pediatric primary ALL [2,43,45,46,106].
In addition, others studies show that in chronic lymphocytic leukemia (CLL), GC resistance is mediated by LCK, through the aberrant expression of the lymphocyte cell-specific tyrosine kinase (Lck), [107]. Activation of LCK leads to its translocation to the membrane to phosphorylate TCR [108] and BCR. As a result, the phospholipase C signaling pathway is activated and production of IP3 results in the release of Ca2+ from the endoplasmic reticulum into the cytoplasm through IP3 receptors [109]. Lck can also regulate intracellular Ca2+ signals by directly activating IP3 channels [110]. Following this, calcineurin is activated by IP3-mediated Ca2+ signaling, leading to translocation and nuclear activation of NF-AT and proinflammatory cytokines such as Il-4 [2,110]. Importantly, calcineurin activation protects cells from GC-induced apoptosis [111]. Lck is also hyperactivated in GC-resistant T-ALL patients which lead to an IL-4 overexpression and resistance to GC [2].

4.6. Hedgehog Signaling

In addition to regulating tissue homeostasis and early T-cell development, the Hedgehog (Hh) signaling pathway contributes to cell cycles and stem cell biology maintenance during embryogenesis [52]. Few studies in the literature have focused on the role and physiological function of the Hedgehog pathway in ALL. Nevertheless, Hh mutations and active signaling have been reported in T-ALL patients [112,113]. Interestingly, it has been shown that Hh pathway contributes to the T-ALL cell growth and survival [51,52] through BCL-2 activation [114]. Regarding the role of GC in Hh signaling, GC positively regulates Hh pathway activation in ALL cell line CEM and inhibition of this pathway by a PKA inhibitor contributes to increased GC-induced cell death [51]. This hypothesis was confirmed both in T-ALL cell lines and in PDX samples showing that inhibition of GLI, the key transcription factor of Hh signaling, enhanced GC-induced cytotoxicity [52]. This protective effect of Hh signaling in GC-induced cytotoxicity has been also demonstrated in others models [115].
As reviewed by Martelli et al. [116], therapeutically relevant signaling pathways interacting with Hh signaling have been identified in T-ALL cells such as the Notch pathway. This plays an important role in leukemic cell growth, survival, and rewired metabolism in both T-ALL [117,118] and B-ALL cells [119]. Oncogenic Notch1 mutations are common in T-ALL and can be detected in 50% of cases [118]. Notch1 oncogenic activity controls a feed-forward loop that promotes cell growth by inducing c-MYC gene expression [118,120] and confer primary GC resistance [117]. Notch1 also regulates other survival signaling pathways such as PI3K-AKT [121], Bcl-2 [117] and IL-7Rα [122] pathways.
The Hh pathway is active in T-ALL cells and it crosstalks with Notch and GC signaling pathways [52]. A combination treatment consisting of a Notch inhibitor (γ-secretase inhibitor or GSI) and a Hh inhibitor (cyclopamine) downregulates the active fragment of Notch (intracellular NOTC, ICN) expression in suppressing the growth of T-ALL cell lines [123]. Moreover, GANT61, a Hh/GLI inhibitor, has demonstrated high cytotoxicity in Notch-dependent T-ALL [124]. Therefore, these studies suggest that crosstalk between the two signaling pathways, Hh and Notch, contribute to an additive or synergistic effect on the growth of ALL cells. It cannot be ruled out that part of the effects of GCs on the Hh pathway leading to GC-resistance are related to activation of the Notch pathway.
Other signaling pathways (PI3K/Akt/mTOR and MEK/ERK) interact with the Hh pathway. Inhibition of Akt or ERK pathways decreased GLI1 expression levels, suggesting that suggesting a positive regulation of GLI1 transcriptional activity and expression by these two pathways [125], the mechanism of which remains to be explored [116]. However, regarding ERK pathway, it has been shown that ERK2 phosphorylates GLI1 at three sites (Ser102, 116, 130), leading to weakening of GLI1 binding to SUFU, thereby increasing GLI1 transcriptional activity [126]. Thus, the activation of the Hh signaling pathway by GC may be involved in the activation of multiple cell survival signaling pathways.

4.7. Metabolic Reprogramming

In different human tissues, endogenous GC play an essential role in the regulation of energy metabolism in physiological or pathological conditions. Transcriptional profiling of GC-resistant leukemia cells revealed additional factors that may contribute to resistance: glutaminolysis, oxidative phosphorylation (OXPHOS) and glycolysis [89]. Considering glycolytic metabolism in cancer cell reprogramming, Olivas-Aguirre et al. recently reported evidence that GCs trigger a pro-survival mechanism in resistant cells. This mechanism involves a metabolic shift from glycolysis and glutaminolysis to increased lipolysis and fatty acid oxidation [56]. Accordingly, inhibition of this metabolic pathway greatly increased ALL cells sensitivity to GC and overcame GC resistance [56].

5. Inhibition of Paradoxical Signaling Nodes to Overcome GC Resistance

Overcoming resistance to GCs has become a necessity to improve ALL patients’ outcomes, but there are hurdles in finding effective approaches or treatments. It is increasingly accepted that GR genetic variations are not linked to GC resistance of ALL cells, but rather through deregulated signaling pathways [127,128,129]. These signaling pathways involved in the regulation of proliferation, differentiation, apoptosis, proteostasis, metabolism, autophagy, oncogenes and epigenetic modifications or tumor suppressors represent processes that may lead to therapy-induced resistance. Indeed, many functional studies as described in this report support the hypothesis that resistance to GCs is not only mediated by constitutive activity of these signaling pathways, but that they are also paradoxically activated by the GC themselves, thereby counteracting their therapeutic effect. Thus, it is not surprising that targeting these paradoxical pathways by promising small molecule inhibitors would overcome the resistance and increase the response of ALL cells to GCs as presented in Table 3. We present hereafter evidence for the modulation of these interconnected paradoxical pathways (identified in the previous section) as a mechanism for overcoming GC resistance.

5.1. Inhibition of Ca2+ and MAPK Signaling

In healthy lymphocytes as well as in ALL cells, Ca2+ signaling plays an important role in cell proliferation [53,54,55]; therefore, Ca2+-mediated survival can be targeted directly or indirectly to improve GC sensitivity and apoptosis. We have previously shown that GC-induced intracellular Ca2+ inhibition or chelation with Bapta-AM significantly potentiates, whereas the increase in cytosolic Ca2+ by thapsigargin suppresses the sensitivity of ALL cell lines and primary specimens to GC treatment [37]. We further demonstrated that a specific TRPC3 blocker, Pyr3, leads to a decrease in GC-induced Ca2+ signal, thereby promoting GC-mediated cell death [38,130]. Furthermore, Ca2+ signal induction is able to activate ERK1/2 protein and decrease GC responsiveness by limiting mitochondrial-mediated apoptotic signals [37]. Therefore, inhibition of MAPK pathways is expected to suppress ALL cell growth. For example, inhibition of a MAPK pathway (ERK1/2, MEK1/2, MEK2 or MEK4) increased ALL cells sensitivity to GC [37,72,99,131]. In addition, MEK overexpression induced ALL cells resistance to GC [131]. Furthermore, inhibition of JNK (another MAPK-signaling pathway), with SP600125 re-sensitizes GC response in ALL [50].

5.2. Inhibitor of PI3/AKT/mTOR Pathway

PI3K/AKT/mTOR modulation is expected to overcome resistance in ALL [48,86,132]. Several ALL cell lines and patients with ALL-derived xenograft samples treated with PI3K inhibitor AS605240 revealed a synergetic effect with GC [133]. Further experiments demonstrated that inhibition of AKT can reverse GC resistance by restoring GR translocation to the nucleus [11,132]. Akt2 inhibition has been shown to be more effective in restoring GC sensitivity than Akt1 inhibition, demonstrated by a higher synergistic effect with dexamethasone and less hepatic cytotoxicity [48]. Selective mTOR kinase modulators that target the catalytic subunit of mTor and lead to a pro-apoptotic and anti-proliferative phenotype [134]. As recently developed in ALL preclinical studies, the dual PI3K-mTOR inhibitor BEZ235 controls apoptosis markers [135] such as expression of pro-apoptotic BIM and anti-apoptotic MCL1, thereby restores GC sensitivity [136].

5.3. Inhibitor of IL-7R and BCL-2 Signaling

We previously saw that IL-7-mediated GC resistance involves an increase in pro-survival BCL-2 signaling via the IL-7R/JAK/STAT5 pathway (Figure 2) [39,75]. In this state of mind, inhibiting these actors in this pathway can directly antagonize GC-induced apoptosis. For example, disruption of the Sec61 translocon, which prevents IL-7R from reaching the membrane surface, overcomes GC resistance in ALL [137]. Moreover, JAK1/2 inhibition as well as BCL-2 inhibition can effectively overcome GC resistance in ALL [39,40,69,138,139]. A loss of BCL2 phosphorylation was correlated with JAK3 inhibition with tofacitinib. The synergy between tofacitinib and conventional chemotherapies, including GC, has been suggested as a good strategy for ALL, in vitro and in vivo [42]. Consistently, paraoxonase-2 silencing, which expression was correlated with reduction of Bcl-2 expression, inhibits tumor growth by sensitizing ALL cells to GC in vivo and in vitro [140].

5.4. Inhibition of Hedgehog Signaling

Recent report demonstrated that inhibition of Hedgehog signaling via GLI inhibitor GANT61 was shown to exert a synergistic anti-leukemic effect with GC in ALL cell lines and patient-derived xenografts samples [52]. Moreover, inhibition of Hedgehog activity increases ALL cells sensitivity to GC [51]. Likewise, NOTCH1 signaling promotes cell cycle progression and limits GC sensitivity [141]. Thus, inhibition of this pathway, through γ-secretase inhibitors (GSI), reverses GC resistance by upregulation of the GR and the pro-apoptotic BIM proteins [142,143].

5.5. Lck Inhibition

Lck inhibition has shown encouraging results. For example, pharmacological inhibitors such as dasatinib, nintedanib, WH-4-023 and bosutinib or specific Lck gene silencing, were shown to induce cell death in GC-resistant ALL cells, and combined treatment with GCs is able to reverse GC resistance in vitro, ex vivo, in vivo and in PDX samples [2,43,45,46,47]. This improvement in GC sensitivity is correlated with an increase in GRs, following inhibition of Lck, which means that Lck exerts a negative action on GR [46]. In addition, inhibition of Lck or aberrantly expressed Lck reverses GC insensitivity in chronic lymphocytic leukemia [107].

6. What Direction to Go in to Decipher this Paradox and Move the Field Forward

Although outcomes for children with ALL have dramatically improved over the past decades, some patients are treatment refractory and resistant to GCs. Many studies have focused on understanding the mechanisms of acquiring GC resistance observed during treatment, including genetic or epigenetic alterations in ALL cells [1,144]. Upfront detection of therapy-resistant pathways at the time of diagnosis may allow for treatment adaptation and relapse prevention. Indeed, treatment of ALL by GCs is also limited by primary resistance, i.e., from the upfront GC treatment [26,27]. Therefore, the initial behavior of cells in response to treatment is a major predictor of the efficacy of chemotherapy and the long-term outcomes of patients with B-ALL [145]. The phenomenon of resistance is responsible, in about 20–25% of patients, for relapse [146]. Resistance may justify therapeutic escalation to, for example, hematopoietic stem cell transplantation (HSCT) and/or chimeric antigen receptor (CAR) T cell transplantation. This resistance is observable from the first phase of treatment with GCs as monotherapy [35,36,40], suggesting the existence of intrinsic differences in GC sensitivity in ALL patients at the time of disease diagnosis. These differences can have long-term prognostic significance. This hypothesis is consolidated by the Berlin–Frankfurt–Münster (ALL-BFM) 95 trial, in which ALL patients were stratified into two groups following 7 days of prednisone monotherapy: those who had a prednisone poor response (PPR) and those who had a prednisone good response (PGR). PPR patients had an 8-year event-free survival rate of only 55.1%, as opposed to 81.3% for patients with a PGR [35]. Consequently, understanding the mechanistic basis for intrinsic differential resistance to GC, preventing or reversing it could mean a real breakthrough in the treatment of ALL. Among these mechanisms, attention should be paid to the cell survival pathways paradoxically activated by the GC themselves. Some recent studies have highlighted this paradoxical behavior of GC by the discovery of certain pathways such as Ca2+ signaling, IL-7R, PI3/AKT, MAPK/ERK and Hedgehog signaling. Still, further experiments are needed to identify these activated pathways from the upfront GC treatment. Then, it will be necessary to identify all the key molecules of these pathways that are paradoxically activated by the therapy before proceeding to development of a personalized therapy program in which patients whose cells strongly express these targets are treated by the combination of treatment with modulators of the identified pathway and first-line GC.

7. Conclusions

The therapy of ALL has been the subject of considerable efforts over the past decades. Despite this, many children and adult patients relapse. Resistance to corticosteroids is one of the main causes of major failure of the induction phase and relapse. This indicates an urgent need to understand the underlying resistance mechanisms and explore ways to circumvent them. Current efforts have provided valuable knowledge to understand these mechanisms and to develop small molecule inhibitors specifically to improve GC therapy. The best way to improve research on this subject would be to increase the understanding of cell survival pathways paradoxically activated by the GC themselves in order to prevent possible relapses. However, other escape mechanisms remain to be discovered. Focus on rational and feasible combinations of these pathways is essential and necessary to limit toxicity and to improve patients’ quality of life and the general status of patient outcomes.

Author Contributions

Conceptualization, S.A.-A.; bibliographic search and writing-original draft preparation, L.A. and S.A.-A.; writing-review and editing, P.S., J.-P.V. and S.A.-A.; supervision, S.A.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ligue Contre le Cancer Grant (AK/IP/BC/16094/16792), France, the Comité Départemental de la Ligue Contre le Cancer, Comité de Loire-Atlantique (IP/BC-17040), France and Association Vie et Espoir, Rouen, France.

Acknowledgments

Figures were created with BioRender.com (2023).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wei, G.; Twomey, D.; Lamb, J.; Schlis, K.; Agarwal, J.; Stam, R.W.; Opferman, J.T.; Sallan, S.E.; den Boer, M.L.; Pieters, R.; et al. Gene Expression-Based Chemical Genomics Identifies Rapamycin as a Modulator of MCL1 and Glucocorticoid Resistance. Cancer Cell 2006, 10, 331–342. [Google Scholar] [CrossRef] [PubMed]
  2. Serafin, V.; Capuzzo, G.; Milani, G.; Minuzzo, S.A.; Pinazza, M.; Bortolozzi, R.; Bresolin, S.; Porcù, E.; Frasson, C.; Indraccolo, S.; et al. Glucocorticoid Resistance Is Reverted by LCK Inhibition in Pediatric T-Cell Acute Lymphoblastic Leukemia. Blood 2017, 130, 2750–2761. [Google Scholar] [CrossRef] [PubMed]
  3. Bhojwani, D.; Yang, J.J.; Pui, C.-H. Biology of Childhood Acute Lymphoblastic Leukemia. Pediatr. Clin. N. Am. 2015, 62, 47–60. [Google Scholar] [CrossRef] [PubMed]
  4. Pui, C.-H.; Robison, L.L.; Look, A.T. Acute Lymphoblastic Leukaemia. Lancet 2008, 371, 1030–1043. [Google Scholar] [CrossRef]
  5. Rafei, H.; Kantarjian, H.M.; Jabbour, E.J. Recent Advances in the Treatment of Acute Lymphoblastic Leukemia. Leuk. Lymphoma 2019, 60, 2606–2621. [Google Scholar] [CrossRef]
  6. Hunger, S.P.; Lu, X.; Devidas, M.; Camitta, B.M.; Gaynon, P.S.; Winick, N.J.; Reaman, G.H.; Carroll, W.L. Improved Survival for Children and Adolescents with Acute Lymphoblastic Leukemia Between 1990 and 2005: A Report from the Children’s Oncology Group Listen to the Podcast by Dr Silverman at Www.Jco.Org/Podcasts. J. Clin. Oncol. 2012, 30, 1663–1669. [Google Scholar] [CrossRef]
  7. Pui, C.-H.; Yang, J.J.; Hunger, S.P.; Pieters, R.; Schrappe, M.; Biondi, A.; Vora, A.; Baruchel, A.; Silverman, L.B.; Schmiegelow, K.; et al. Childhood Acute Lymphoblastic Leukemia: Progress Through Collaboration. J. Clin. Oncol. 2015, 33, 2938–2948. [Google Scholar] [CrossRef]
  8. Chan, K.W. Acute Lymphoblastic Leukemia. Curr. Probl. Pediatr. Adolesc. Health Care 2002, 32, 40–49. [Google Scholar] [CrossRef]
  9. Van Vlierberghe, P.; Pieters, R.; Beverloo, H.B.; Meijerink, J.P.P. Molecular-Genetic Insights in Paediatric T-Cell Acute Lymphoblastic Leukaemia. Br. J. Haematol. 2008, 143, 153–168. [Google Scholar] [CrossRef]
  10. De Smedt, R.; Morscio, J.; Goossens, S.; Van Vlierberghe, P. Targeting Steroid Resistance in T-Cell Acute Lymphoblastic Leukemia. Blood Rev. 2019, 38, 100591. [Google Scholar] [CrossRef]
  11. Piovan, E.; Yu, J.; Tosello, V.; Herranz, D.; Ambesi-Impiombato, A.; DaSilva, A.C.; Sanchez-Martin, M.; Perez-Garcia, A.; Rigo, I.; Castillo, M.; et al. Direct Reversal of Glucocorticoid Resistance by AKT Inhibition in Acute Lymphoblastic Leukemia. Cancer Cell 2013, 24, 766–776. [Google Scholar] [CrossRef] [PubMed]
  12. Wittmann, M.; Riccardi, C.; Timmermans, S.; Souffriau, J.; Libert, C. A General Introduction to Glucocorticoid Biology. Front. Immunol. 2019, 1, 1545. [Google Scholar] [CrossRef]
  13. Tissing, W.J.E.; Meijerink, J.P.P.; den Boer, M.L.; Pieters, R. Molecular Determinants of Glucocorticoid Sensitivity and Resistance in Acute Lymphoblastic Leukemia. Leukemia 2003, 17, 17–25. [Google Scholar] [CrossRef] [PubMed]
  14. Inaba, H.; Pui, C.H. Glucocorticoid Use in Acute Lymphoblastic Leukaemia. Lancet Oncol. 2010, 11, 1096–1106. [Google Scholar] [CrossRef]
  15. Montgomery, B.; Cheng, H.H.; Drechsler, J.; Mostaghel, E.A. Glucocorticoids and Prostate Cancer Treatment: Friend or Foe? Asian J. Androl. 2014, 16, 354–358. [Google Scholar] [CrossRef]
  16. Möricke, A.; Zimmermann, M.; Valsecchi, M.G.; Stanulla, M.; Biondi, A.; Mann, G.; Locatelli, F.; Cazzaniga, G.; Niggli, F.; Aricò, M.; et al. Dexamethasone vs Prednisone in Induction Treatment of Pediatric ALL: Results of the Randomized Trial AIEOP-BFM ALL 2000. Blood 2016, 127, 2101–2112. [Google Scholar] [CrossRef]
  17. Uckun, F.M.; Gaynon, P.S.; Stram, D.O.; Sensel, M.G.; Sarquis, M.B.; Willoughby, M. Bone Marrow Leukemic Progenitor Cell Content in Pediatric T-Lineage Acute Lymphoblastic Leukemia Patients with an Isolated Extramedullary First Relapse. Leuk. Lymphoma 2001, 40, 279–285. [Google Scholar] [CrossRef]
  18. Uckun, F.M.; Gaynon, P.S.; Stram, D.O.; Sensel, M.G.; Sarquis, M.B.; Lazarus, K.H.; Willoughby, M. Paucity of Leukemic Progenitor Cells in the Bone Marrow of Pediatric B- Lineage Acute Lymphoblastic Leukemia Patients with an Isolated Extramedullary First Relapse. Clin. Cancer Res. 1999, 5, 2415–2420. [Google Scholar]
  19. Liu, X.; Zou, Y.; Chen, X.; Wang, S.; Guo, Y.; Yang, W.; Zhang, L.; Chen, Y.; Zhang, Y.; Zhu, X. Minimal Residual Disease Surveillance at Day 90 Predicts Long-Term Survival in Pediatric Patients with T-Cell Acute Lymphoblastic Leukemia. Leuk. Lymphoma 2020, 61, 3460–3467. [Google Scholar] [CrossRef]
  20. Bostrom, B.C.; Sensel, M.R.; Sather, H.N.; Gaynon, P.S.; La, M.K.; Johnston, K.; Erdmann, G.R.; Gold, S.; Heerema, N.A.; Hutchinson, R.J.; et al. Dexamethasone versus Prednisone and Daily Oral versus Weekly Intravenous Mercaptopurine for Patients with Standard-Risk Acute Lymphoblastic Leukemia: A Report from the Children’s Cancer Group. Blood 2003, 101, 3809–3817. [Google Scholar] [CrossRef]
  21. Larsen, E.C.; Devidas, M.; Chen, S.; Salzer, W.L.; Raetz, E.A.; Loh, M.L.; Mattano, L.A.; Cole, C.; Eicher, A.; Haugan, M.; et al. Dexamethasone and High-Dose Methotrexate Improve Outcome for Children and Young Adults with High-Risk B-Acute Lymphoblastic Leukemia: A Report From Children’s Oncology Group Study AALL0232. J. Clin. Oncol. 2016, 34, 2380–2388. [Google Scholar] [CrossRef] [PubMed]
  22. Domenech, C.; Suciu, S.; De Moerloose, B.; Mazingue, F.; Plat, G.; Ferster, A.; Uyttebroeck, A.; Sirvent, N.; Lutz, P.; Yakouben, K.; et al. Dexamethasone (6 Mg/M2/Day) and Prednisolone (60 Mg/M2/Day) Were Equally Effective as Induction Therapy for Childhood Acute Lymphoblastic Leukemia in the EORTC CLG 58951 Randomized Trial. Haematologica 2014, 99, 1220–1227. [Google Scholar] [CrossRef]
  23. Erlacher, M.; Knoflach, M.; Stec, I.E.M.; Böck, G.; Wick, G.; Wiegers, G.J. TCR Signaling Inhibits Glucocorticoid-Induced Apoptosis in Murine Thymocytes Depending on the Stage of Development. Eur. J. Immunol. 2005, 35, 3287–3296. [Google Scholar] [CrossRef] [PubMed]
  24. Jamieson, C.A.M.; Yamamoto, K.R. Crosstalk Pathway for Inhibition of Glucocorticoid-Induced Apoptosis by T Cell Receptor Signaling. Proc. Natl. Acad. Sci. USA 2000, 97, 7319–7324. [Google Scholar] [CrossRef]
  25. Tsitoura, D.C.; Rothman, P.B. Enhancement of MEK/ERK Signaling Promotes Glucocorticoid Resistance in CD4+ T Cells. J. Clin. Investig. 2004, 113, 619–627. [Google Scholar] [CrossRef] [PubMed]
  26. DeFranco, D.B. Role of Molecular Chaperones in Subnuclear Trafficking of Glucocorticoid Receptors. Kidney Int. 2000, 57, 1241–1249. [Google Scholar] [CrossRef] [PubMed]
  27. Clarisse, D.; Offner, F.; De Bosscher, K. Latest Perspectives on Glucocorticoid-Induced Apoptosis and Resistance in Lymphoid Malignancies. Biochim. Biophys. Acta Rev. Cancer 2020, 1874, 188430. [Google Scholar] [CrossRef]
  28. Brown, J.A.; Ferrando, A. Glucocorticoid Resistance in Acute Lymphoblastic Leukemia: BIM Finally. Cancer Cell 2018, 34, 869–871. [Google Scholar] [CrossRef]
  29. Schmidt, S.; Rainer, J.; Ploner, C.; Presul, E.; Riml, S.; Kofler, R. Glucocorticoid-Induced Apoptosis and Glucocorticoid Resistance: Molecular Mechanisms and Clinical Relevance. Cell Death Differ. 2004, 11, 45–55. [Google Scholar] [CrossRef]
  30. Mullighan, C.G.; Zhang, J.; Kasper, L.H.; Lerach, S.; Payne-Turner, D.; Phillips, L.A.; Heatley, S.L.; Holmfeldt, L.; Collins-Underwood, J.R.; Ma, J.; et al. CREBBP Mutations in Relapsed Acute Lymphoblastic Leukaemia. Nature 2011, 471, 235–239. [Google Scholar] [CrossRef]
  31. Nicolaides, N.C.; Charmandari, E. Novel Insights into the Molecular Mechanisms Underlying Generalized Glucocorticoid Resistance and Hypersensitivity Syndromes. Hormones 2017, 16, 124–138. [Google Scholar] [PubMed]
  32. Hala, M.; Hartmann, B.L.; Bock, G.; Geley, S.; Kofler, R. Glucocorticoid-Receptor-Gene Defects and Resistance to Glucocorticoid-Induced Apoptosis in Human Leukemic Cell Lines. Int. J. Cancer 1996, 68, 663–668. [Google Scholar] [CrossRef]
  33. Strasser-Wozak, E.M.C.; Hattmannstorfer, R.; Hála, M.; Hartmann, B.L.; Fiegl, M.; Geley, S.; Kofler, R. Splice Site Mutation in the Glucocorticoid Receptor Gene Causes Resistance to Glucocorticoid-Induced Apoptosis in a Human Acute Leukemic Cell Line. Cancer Res. 1995, 55, 348–353. [Google Scholar]
  34. Klumper, E.; Pieters, R.; Veerman, A.J.P.; Huismans, D.R.; Loonen, A.H.; Hählen, K.; Kaspers, G.J.L.; Van Wering, E.R.; Hartmann, R.; Henze, G. In Vitro Cellular Drug Resistance in Children with Relapsed/Refractory Acute Lymphoblastic Leukemia. Blood 1995, 86, 3861–3868. [Google Scholar] [CrossRef] [PubMed]
  35. Lauten, M.; Möricke, A.; Beier, R.; Zimmermann, M.; Stanulla, M.; Meissner, B.; Odenwald, E.; Attarbaschi, A.; Niemeyer, C.; Niggli, F.; et al. Prediction of Outcome by Early Bone Marrow Response in Childhood Acute Lymphoblastic Leukemia Treated in the ALL-BFM 95 Trial: Differential Effects in Precursor B-Cell and T-Cell Leukemia. Haematologica 2012, 97, 1048–1056. [Google Scholar] [CrossRef] [PubMed]
  36. Conter, V.; Bartram, C.R.; Valsecchi, M.G.; Schrauder, A.; Panzer-Grümayer, R.; Möricke, A.; Aricò, M.; Zimmermann, M.; Mann, G.; De Rossi, G.; et al. Molecular Response to Treatment Redefines All Prognostic Factors in Children and Adolescents with B-Cell Precursor Acute Lymphoblastic Leukemia: Results in 3184 Patients of the AIEOP-BFM ALL 2000 Study. Blood 2010, 115, 3206–3214. [Google Scholar] [CrossRef] [PubMed]
  37. Abdoul-Azize, S.; Dubus, I.; Vannier, J.P. Improvement of Dexamethasone Sensitivity by Chelation of Intracellular Ca2+ in Pediatric Acute Lymphoblastic Leukemia Cells through the Prosurvival Kinase ERK1/2 Deactivation. Oncotarget 2017, 8, 27339–27352. [Google Scholar] [CrossRef]
  38. Abdoul-Azize, S.; Buquet, C.; Vannier, J.-P.; Dubus, I. Pyr3, a TRPC3 Channel Blocker, Potentiates Dexamethasone Sensitivity and Apoptosis in Acute Lymphoblastic Leukemia Cells by Disturbing Ca2+ Signaling, Mitochondrial Membrane Potential Changes and Reactive Oxygen Species Production. Eur. J. Pharmacol. 2016, 784, 90–98. [Google Scholar] [CrossRef]
  39. Meyer, L.K.; Huang, B.J.; Delgado-Martin, C.; Roy, R.P.; Hechmer, A.; Wandler, A.M.; Vincent, T.L.; Fortina, P.; Olshen, A.B.; Wood, B.L.; et al. Glucocorticoids Paradoxically Facilitate Steroid Resistance in T Cell Acute Lymphoblastic Leukemias and Thymocytes. J. Clin. Investig. 2020, 130, 863–876. [Google Scholar] [CrossRef]
  40. Delgado-Martin, C.; Meyer, L.K.; Huang, B.J.; Shimano, K.A.; Zinter, M.S.; Nguyen, J.V.; Smith, G.A.; Taunton, J.; Winter, S.S.; Roderick, J.R.; et al. JAK/STAT Pathway Inhibition Overcomes IL7-Induced Glucocorticoid Resistance in a Subset of Human T-Cell Acute Lymphoblastic Leukemias. Leukemia 2017, 31, 2568–2576. [Google Scholar] [CrossRef]
  41. De Groot, A.P.; Saito, Y.; Kawakami, E.; Hashimoto, M.; Aoki, Y.; Ono, R.; Ogahara, I.; Fujiki, S.; Kaneko, A.; Sato, K.; et al. Targeting Critical Kinases and Anti-Apoptotic Molecules Overcomes Steroid Resistance in MLL-Rearranged Leukaemia. EBioMedicine 2021, 64, 103235. [Google Scholar] [CrossRef] [PubMed]
  42. Bodaar, K.; Yamagata, N.; Barthe, A.; Landrigan, J.; Chonghaile, T.N.; Burns, M.; Stevenson, K.E.; Devidas, M.; Loh, M.L.; Hunger, S.P.; et al. JAK3 Mutations and Mitochondrial Apoptosis Resistance in T-Cell Acute Lymphoblastic Leukemia. Leukemia 2022, 36, 1499–1507. [Google Scholar] [CrossRef] [PubMed]
  43. Shi, Y.; Beckett, M.C.; Blair, H.J.; Tirtakusuma, R.; Nakjang, S.; Enshaei, A.; Halsey, C.; Vormoor, J.; Heidenreich, O.; Krippner-Heidenreich, A.; et al. Phase II-like Murine Trial Identifies Synergy between Dexamethasone and Dasatinib in T-Cell Acute Lymphoblastic Leukemia. Haematologica 2021, 106, 1056–1066. [Google Scholar] [CrossRef] [PubMed]
  44. Ghosh, M.C.; Baatar, D.; Collins, G.; Carter, A.; Indig, F.; Biragyn, A.; Taub, D.D. Dexamethasone Augments CXCR4-Mediated Signaling in Resting Human T Cells via the Activation of the Src Kinase Lck. Blood 2009, 113, 575–584. [Google Scholar] [CrossRef]
  45. Fazio, G.; Bresolin, S.; Silvestri, D.; Quadri, M.; Saitta, C.; Vendramini, E.; Buldini, B.; Palmi, C.; Bardini, M.; Grioni, A.; et al. PAX5 Fusion Genes Are Frequent in Poor Risk Childhood Acute Lymphoblastic Leukaemia and Can Be Targeted with BIBF1120. EBioMedicine 2022, 83, 104224. [Google Scholar] [CrossRef]
  46. Jin, Q.; Gutierrez Diaz, B.; Pieters, T.; Zhou, Y.; Narang, S.; Fijalkwoski, I.; Borin, C.; Van Laere, J.; Payton, M.; Cho, B.-K.; et al. Oncogenic Deubiquitination Controls Tyrosine Kinase Signaling and Therapy Response in Acute Lymphoblastic Leukemia. Sci. Adv. 2022, 8, eabq8437. [Google Scholar] [CrossRef]
  47. Spijkers-Hagelstein, J.A.P.; Mimoso Pinhanços, S.; Schneider, P.; Pieters, R.; Stam, R.W. Src Kinase-Induced Phosphorylation of Annexin A2 Mediates Glucocorticoid Resistance in MLL-Rearranged Infant Acute Lymphoblastic Leukemia. Leukemia 2013, 27, 1063–1071. [Google Scholar] [CrossRef]
  48. Xie, M.; Yang, A.; Ma, J.; Wu, M.; Xu, H.; Wu, K.; Jin, Y.; Xie, Y. Akt2 Mediates Glucocorticoid Resistance in Lymphoid Malignancies through FoxO3a/Bim Axis and Serves as a Direct Target for Resistance Reversal. Cell Death Dis. 2019, 9, 1013. [Google Scholar] [CrossRef]
  49. Miller, A.L.; Garza, A.S.; Johnson, B.H.; Thompson, E.B. Pathway Interactions between MAPKs, MTOR, PKA, and the Glucocorticoid Receptor in Lymphoid Cells. Cancer Cell Int. 2007, 7, 3. [Google Scholar] [CrossRef]
  50. Nicholson, L.; Evans, C.A.; Matheson, E.; Minto, L.; Keilty, C.; Sanichar, M.; Case, M.; Schwab, C.; Williamson, D.; Rainer, J.; et al. Quantitative Proteomic Analysis Reveals Maturation as a Mechanism Underlying Glucocorticoid Resistance in B Lineage ALL and Re-Sensitization by JNK Inhibition. Br. J. Haematol. 2015, 171, 595–605. [Google Scholar] [CrossRef]
  51. Ji, Z.; Mei, F.C.; Johnson, B.H.; Thompson, E.B.; Cheng, X. Protein Kinase A, Not Epac, Suppresses Hedgehog Activity and Regulates Glucocorticoid Sensitivity in Acute Lymphoblastic Leukemia Cells. J. Biol. Chem. 2007, 282, 37370–37377. [Google Scholar] [CrossRef]
  52. Bongiovanni, D.; Tosello, V.; Saccomani, V.; Dalla Santa, S.; Amadori, A.; Zanovello, P.; Piovan, E. Crosstalk between Hedgehog Pathway and the Glucocorticoid Receptor Pathway as a Basis for Combination Therapy in T-Cell Acute Lymphoblastic Leukemia. Oncogene 2020, 39, 6544–6555. [Google Scholar] [CrossRef] [PubMed]
  53. Prevarskaya, N.; Skryma, R.; Shuba, Y. Calcium in Tumour Metastasis: New Roles for Known Actors. Nat. Rev. Cancer 2011, 11, 609–618. [Google Scholar] [CrossRef] [PubMed]
  54. Monteith, G.R.; McAndrew, D.; Faddy, H.M.; Roberts-Thomson, S.J. Calcium and Cancer: Targeting Ca2+ Transport. Nat. Rev. Cancer 2007, 7, 519–530. [Google Scholar] [CrossRef] [PubMed]
  55. Roderick, H.L.; Cook, S.J. Ca2+ Signalling Checkpoints in Cancer: Remodelling Ca2+ for Cancer Cell Proliferation and Survival. Nat. Rev. Cancer 2008, 8, 361–375. [Google Scholar] [CrossRef] [PubMed]
  56. Olivas-Aguirre, M.; Pérez-Chávez, J.; Torres-López, L.; Hernández-Cruz, A.; Pottosin, I.; Dobrovinskaya, O. Dexamethasone-Induced Fatty Acid Oxidation and Autophagy/Mitophagy Are Essential for T-ALL Glucocorticoid Resistance. Cancers 2023, 15, 445. [Google Scholar] [CrossRef] [PubMed]
  57. Guo, Y.; Hao, D.; Hu, H. High Doses of Dexamethasone Induce Endoplasmic Reticulum Stress-Mediated Apoptosis by Promoting Calcium Ion Influx-Dependent CHOP Expression in Osteoblasts. Mol. Biol. Rep. 2021, 48, 7841–7851. [Google Scholar] [CrossRef]
  58. Feske, S.; Skolnik, E.Y.; Prakriya, M. Ion Channels and Transporters in Lymphocyte Function and Immunity. Nat. Rev. Immunol. 2012, 12, 532–547. [Google Scholar] [CrossRef]
  59. Dobrovinskaya, O.; Delgado-Enciso, I.; Quintero-Castro, L.J.; Best-Aguilera, C.; Rojas-Sotelo, R.M.; Pottosin, I. Placing Ion Channels into a Signaling Network of T Cells: From Maturing Thymocytes to Healthy T Lymphocytes or Leukemic T Lymphoblasts. Biomed Res. Int. 2015, 2015, 750203. [Google Scholar] [CrossRef]
  60. Sarang, Z.; Gyurina, K.; Scholtz, B.; Kiss, C.; Szegedi, I. Altered Expression of Autophagy-Related Genes Might Contribute to Glucocorticoid Resistance in Precursor B-Cell-Type Acute Lymphoblastic Leukemia. Eur. J. Haematol. 2016, 97, 453–460. [Google Scholar] [CrossRef]
  61. Eylenstein, A.; Gehring, E.; Heise, N.; Shumilina, E.; Schmidt, S.; Szteyn, K.; Münzer, P.; Nurbaeva, M.K.; Eichenmüller, M.; Tyan, L.; et al. Stimulation of Ca2+-Channel Orai1/STIM1 by Serum- and Glucocorticoid-Inducible Kinase 1 (SGK1). FASEB J. 2011, 25, 2012–2021. [Google Scholar] [CrossRef] [PubMed]
  62. Lang, F.; Pelzl, L.; Hauser, S.; Hermann, A.; Stournaras, C.; Schöls, L. To Die or Not to Die SGK1-Sensitive ORAI/STIM in Cell Survival. Cell Calcium 2018, 74, 29–34. [Google Scholar] [CrossRef] [PubMed]
  63. Yu, W.; Honisch, S.; Schmidt, S.; Yan, J.; Schmid, E.; Alkahtani, S.; Alkahtane, A.A.; Alarifi, S.; Stournaras, C.; Lang, F. Chorein Sensitive Orai1 Expression and Store Operated Ca2+ Entry in Rhabdomyosarcoma Cells. Cell. Physiol. Biochem. 2016, 40, 1141–1152. [Google Scholar] [CrossRef] [PubMed]
  64. Lang, F.; Böhmer, C.; Palmada, M.; Seebohm, G.; Strutz-Seebohm, N.; Vallon, V. (Patho)Physiological Significance of the Serum- and Glucocorticoid-Inducible Kinase Isoforms. Physiol. Rev. 2006, 86, 1151–1178. [Google Scholar] [CrossRef]
  65. Lang, F.; Voelkl, J. Therapeutic Potential of Serum and Glucocorticoid Inducible Kinase Inhibition. Expert Opin. Investig. Drugs 2013, 22, 701–714. [Google Scholar] [CrossRef]
  66. Valle-Reyes, S.; Valencia-Cruz, G.; Liñan-Rico, L.; Pottosin, I.; Dobrovinskaya, O. Differential Activity of Voltage- and Ca2+-Dependent Potassium Channels in Leukemic T Cell Lines: Jurkat Cells Represent an Exceptional Case. Front. Physiol. 2018, 9, 499. [Google Scholar] [CrossRef]
  67. Valle, J.S.; Castellanos, R.; Olivas, M.Á.; Pottosin, I.; Dobrovinskaya, O.; Schnoor, M. Kv1.3 Channel Is a Potential Marker for B Acute Lymphoblastic Leukemia. FASEB J. 2020, 34, 1. [Google Scholar] [CrossRef]
  68. Henke, G.; Maier, G.; Wallisch, S.; Boehmer, C.; Lang, F. Regulation of the Voltage Gated K+ Channel Kv1.3 by the Ubiquitin Ligase Nedd4-2 and the Serum and Glucocorticoid Inducible Kinase SGK1. J. Cell. Physiol. 2004, 199, 194–199. [Google Scholar] [CrossRef]
  69. Ribeiro, D.; Melão, A.; van Boxtel, R.; Santos, C.I.; Silva, A.; Silva, M.C.; Cardoso, B.A.; Coffer, P.J.; Barata, J.T. STAT5 Is Essential for IL-7-Mediated Viability, Growth, and Proliferation of T-Cell Acute Lymphoblastic Leukemia Cells. Blood Adv. 2018, 2, 2199–2213. [Google Scholar] [CrossRef]
  70. Silva, A.; Laranjeira, A.B.A.; Martins, L.R.; Cardoso, B.A.; Demengeot, J.; Yunes, J.A.; Seddon, B.; Barata, J.T. IL-7 Contributes to the Progression of Human T-Cell Acute Lymphoblastic Leukemias. Cancer Res. 2011, 71, 4780–4789. [Google Scholar] [CrossRef]
  71. Jing, D.; Bhadri, V.A.; Beck, D.; Thoms, J.A.I.; Yakob, N.A.; Wong, J.W.H.; Knezevic, K.; Pimanda, J.E.; Lock, R.B. Opposing Regulation of BIM and BCL2 Controls Glucocorticoid-Induced Apoptosis of Pediatric Acute Lymphoblastic Leukemia Cells. Blood 2015, 125, 273–283. [Google Scholar] [CrossRef] [PubMed]
  72. Li, Y.; Buijs-Gladdines, J.G.C.A.M.; Canté-Barrett, K.; Stubbs, A.P.; Vroegindeweij, E.M.; Smits, W.K.; van Marion, R.; Dinjens, W.N.M.; Horstmann, M.; Kuiper, R.P.; et al. IL-7 Receptor Mutations and Steroid Resistance in Pediatric T Cell Acute Lymphoblastic Leukemia: A Genome Sequencing Study. PLoS Med. 2016, 13, e1002200. [Google Scholar] [CrossRef] [PubMed]
  73. Barata, J.T.; Cardoso, A.A.; Nadler, L.M.; Boussiotis, V.A. Interleukin-7 Promotes Survival and Cell Cycle Progression of T-Cell Acute Lymphoblastic Leukemia Cells by down-Regulating the Cyclin-Dependent Kinase Inhibitor P27kip1. Blood 2001, 98, 1524–1531. [Google Scholar] [CrossRef] [PubMed]
  74. Jiang, Q.; Li, W.Q.; Hofmeister, R.R.; Young, H.A.; Hodge, D.R.; Keller, J.R.; Khaled, A.R.; Durum, S.K. Distinct Regions of the Interleukin-7 Receptor Regulate Different Bcl2 Family Members. Mol. Cell. Biol. 2004, 24, 6501–6513. [Google Scholar] [CrossRef] [PubMed]
  75. Kontro, M.; Kuusanmäki, H.; Eldfors, S.; Burmeister, T.; Andersson, E.I.; Bruserud; Brümmendorf, T.H.; Edgren, H.; Gjertsen, B.T.; Itälä-Remes, M.; et al. Novel Activating STAT5B Mutations as Putative Drivers of T-Cell Acute Lymphoblastic Leukemia. Leukemia 2014, 28, 1738–1742. [Google Scholar] [CrossRef]
  76. De Smedt, R.; Morscio, J.; Reunes, L.; Roels, J.; Bardelli, V.; Lintermans, B.; Van Loocke, W.; Almeida, A.; Cheung, L.C.; Kotecha, R.S.; et al. Targeting Cytokine- and Therapy-Induced PIM1 Activation in Preclinical Models of T-Cell Acute Lymphoblastic Leukemia and Lymphoma. Blood 2020, 135, 1685–1695. [Google Scholar] [CrossRef]
  77. Lin, Y.-W.; Beharry, Z.M.; Hill, E.G.; Song, J.H.; Wang, W.; Xia, Z.; Zhang, Z.; Aplan, P.D.; Aster, J.C.; Smith, C.D.; et al. A Small Molecule Inhibitor of Pim Protein Kinases Blocks the Growth of Precursor T-Cell Lymphoblastic Leukemia/Lymphoma. Blood 2010, 115, 824–833. [Google Scholar] [CrossRef]
  78. Van der Zwet, J.C.G.; Buijs-Gladdines, J.G.C.A.M.; Cordo’, V.; Debets, D.O.; Smits, W.K.; Chen, Z.; Dylus, J.; Zaman, G.J.R.; Altelaar, M.; Oshima, K.; et al. MAPK-ERK Is a Central Pathway in T-Cell Acute Lymphoblastic Leukemia That Drives Steroid Resistance. Leukemia 2021, 35, 3394–3405. [Google Scholar] [CrossRef]
  79. Oliveira, M.L.; Akkapeddi, P.; Ribeiro, D.; Melão, A.; Barata, J.T. IL-7R-Mediated Signaling in T-Cell Acute Lymphoblastic Leukemia: An Update. Adv. Biol. Regul. 2019, 71, 88–96. [Google Scholar] [CrossRef]
  80. Barata, J.T.; Silva, A.; Brandao, J.G.; Nadler, L.M.; Cardoso, A.A.; Boussiotis, V.A. Activation of PI3K Is Indispensable for Interleukin 7-Mediated Viability, Proliferation, Glucose Use, and Growth of T Cell Acute Lymphoblastic Leukemia Cells. J. Exp. Med. 2004, 200, 659–669. [Google Scholar] [CrossRef]
  81. Zenatti, P.P.; Ribeiro, D.; Li, W.; Zuurbier, L.; Silva, M.C.; Paganin, M.; Tritapoe, J.; Hixon, J.A.; Silveira, A.B.; Cardoso, B.A.; et al. Oncogenic IL7R Gain-of-Function Mutations in Childhood T-Cell Acute Lymphoblastic Leukemia. Nat. Genet. 2011, 43, 932–941. [Google Scholar] [CrossRef] [PubMed]
  82. Canté-Barrett, K.; Spijkers-Hagelstein, J.A.P.; Buijs-Gladdines, J.G.C.A.M.; Uitdehaag, J.C.M.; Smits, W.K.; Van Der Zwet, J.; Buijsman, R.C.; Zaman, G.J.R.; Pieters, R.; Meijerink, J.P.P. MEK and PI3K-AKT Inhibitors Synergistically Block Activated IL7 Receptor Signaling in T-Cell Acute Lymphoblastic Leukemia. Leukemia 2016, 30, 1832–1843. [Google Scholar] [CrossRef] [PubMed]
  83. Van der Zwet, J.C.G.; Cordo’, V.; Buijs-Gladdines, J.G.C.A.M.; Hagelaar, R.; Smits, W.K.; Vroegindeweij, E.; Graus, L.T.M.; Poort, V.; Nulle, M.; Pieters, R.; et al. STAT5 Does Not Drive Steroid Resistance in T-Cell Acute Lymphoblastic Leukemia despite the Activation of BCL2 and BCLXL Following Glucocorticoid Treatment. Haematologica 2022, 108, 732–746. [Google Scholar] [CrossRef]
  84. Goossens, S.; Van Vlierberghe, P. Overcoming Steroid Resistance in T Cell Acute Lymphoblastic Leukemia. PLoS Med. 2016, 13, e1002208. [Google Scholar] [CrossRef] [PubMed]
  85. So, L.; Fruman, D.A. PI3K Signalling in B- and T-Lymphocytes: New Developments and Therapeutic Advances. Biochem. J. 2012, 442, 465–481. [Google Scholar] [CrossRef]
  86. Liu, P.; Cheng, H.; Roberts, T.M.; Zhao, J.J. Targeting the Phosphoinositide 3-Kinase Pathway in Cancer. Nat. Rev. Drug Discov. 2009, 8, 627–644. [Google Scholar] [CrossRef]
  87. Bornhauser, B.C.; Bonapace, L.; Lindholm, D.; Martinez, R.; Cario, G.; Schrappe, M.; Niggli, F.K.; Schäfer, B.W.; Bourquin, J.P. Low-Dose Arsenic Trioxide Sensitizes Glucocorticoid-Resistant Acute Lymphoblastic Leukemia Cells to Dexamethasone via an Akt-Dependent Pathway. Blood 2007, 110, 2084–2091. [Google Scholar] [CrossRef]
  88. Datta, S.R.; Dudek, H.; Xu, T.; Masters, S.; Haian, F.; Gotoh, Y.; Greenberg, M.E. Akt Phosphorylation of BAD Couples Survival Signals to the Cell-Intrinsic Death Machinery. Cell 1997, 91, 231–241. [Google Scholar] [CrossRef]
  89. Beesley, A.H.; Firth, M.J.; Ford, J.; Weller, R.E.; Freitas, J.R.; Perera, K.U.; Kees, U.R. Glucocorticoid Resistance in T-Lineage Acute Lymphoblastic Leukaemia Is Associated with a Proliferative Metabolism. Br. J. Cancer 2009, 100, 1926–1936. [Google Scholar] [CrossRef]
  90. Toscan, C.E.; Jing, D.; Mayoh, C.; Lock, R.B. Reversal of Glucocorticoid Resistance in Paediatric Acute Lymphoblastic Leukaemia Is Dependent on Restoring BIM Expression. Transl. Ther. Br. J. Cancer 2020, 122, 1769–1781. [Google Scholar] [CrossRef]
  91. Gutierrez, A.; Sanda, T.; Grebliunaite, R.; Carracedo, A.; Salmena, L.; Ahn, Y.; Dahlberg, S.; Neuberg, D.; Moreau, L.A.; Winter, S.S.; et al. High Frequency of PTEN, PI3K, and AKT Abnormalities in T-Cell Acute Lymphoblastic Leukemia. Blood 2009, 114, 647–650. [Google Scholar] [CrossRef] [PubMed]
  92. Arancibia, S.; Benítez, D.; Núñez, L.E.; Jewell, C.M.; Langjahr, P.; Candia, E.; Zapata-Torres, G.; Cidlowski, J.A.; González, M.-J.; Hermoso, M.A. Phosphatidylinositol 3-Kinase Interacts with the Glucocorticoid Receptor upon TLR2 Activation. J. Cell. Mol. Med. 2011, 15, 339–349. [Google Scholar] [CrossRef] [PubMed]
  93. Tsai, H.J.; Kobayashi, S.; Izawa, K.; Ishida, T.; Watanabe, T.; Umezawa, K.; Lin, S.F.; Tojo, A. Bioimaging Analysis of Nuclear Factor-ΚB Activity in Philadelphia Chromosome-Positive Acute Lymphoblastic Leukemia Cells Reveals Its Synergistic Upregulation by Tumor Necrosis Factor-α-Stimulated Changes to the Microenvironment. Cancer Sci. 2011, 102, 2014–2021. [Google Scholar] [CrossRef] [PubMed]
  94. Birkenkamp, K.U.; Coffer, P.J. Regulation of Cell Survival and Proliferation by the FOXO (Forkhead Box, Class O) Subfamily of Forkhead Transcription Factors. Biochem. Soc. Trans. 2003, 31, 292–297. [Google Scholar] [CrossRef]
  95. Ma, J.; Xie, Y.; Shi, Y.; Qin, W.; Zhao, B.; Jin, Y. Glucocorticoid-Induced Apoptosis Requires FOXO3A Activity. Biochem. Biophys. Res. Commun. 2008, 377, 894–898. [Google Scholar] [CrossRef]
  96. Kruth, K.A.; Fang, M.; Shelton, D.N.; Abu-Halawa, O.; Mahling, R.; Yang, H.; Weissman, J.S.; Loh, M.L.; Müschen, M.; Tasian, S.K.; et al. Suppression of B-Cell Development Genes Is Key to Glucocorticoid Efficacy in Treatment of Acute Lymphoblastic Leukemia. Blood 2017, 129, 3000–3008. [Google Scholar] [CrossRef]
  97. Weigelt, B.; Downward, J. Genomic Determinants of PI3K Pathway Inhibitor Response in Cancer. Front. Oncol. 2012, 2, 109. [Google Scholar] [CrossRef]
  98. Miller, A.L.; Webb, M.S.; Copik, A.J.; Wang, Y.; Johnson, B.H.; Kumar, R.; Thompson, E.B. P38 Mitogen-Activated Protein Kinase (MAPK) Is a Key Mediator in Glucocorticoid-Induced Apoptosis of Lymphoid Cells: Correlation between P38 MAPK Activation and Site-Specific Phosphorylation of the Human Glucocorticoid Receptor at Serine 211. Mol. Endocrinol. 2005, 19, 1569–1583. [Google Scholar] [CrossRef]
  99. Jones, C.L.; Gearheart, C.M.; Fosmire, S.; Delgado-Martin, C.; Evensen, N.A.; Bride, K.; Waanders, A.J.; Pais, F.; Wang, J.; Bhatla, T.; et al. MAPK Signaling Cascades Mediate Distinct Glucocorticoid Resistance Mechanisms in Pediatric Leukemia. Blood 2015, 126, 2202–2212. [Google Scholar] [CrossRef]
  100. Di, J.; Huang, H.; Qu, D.; Tang, J.; Cao, W.; Lu, Z.; Cheng, Q.; Yang, J.; Bai, J.; Zhang, Y.; et al. Rap2B Promotes Proliferation, Migration, and Invasion of Human Breast Cancer through Calcium-Related ERK1/2 Signaling Pathway. Sci. Rep. 2015, 5, 12363. [Google Scholar] [CrossRef]
  101. Cazzaniga, V.; Bugarin, C.; Bardini, M.; Giordan, M.; te Kronnie, G.; Basso, G.; Biondi, A.; Fazio, G.; Cazzaniga, G. LCK Over-Expression Drives STAT5 Oncogenic Signaling in PAX5 Translocated BCP-ALL Patients. Oncotarget 2015, 6, 1569–1581. [Google Scholar] [CrossRef] [PubMed]
  102. Bommhardt, U.; Schraven, B.; Simeoni, L. Beyond TCR Signaling: Emerging Functions of Lck in Cancer and Immunotherapy. Int. J. Mol. Sci. 2019, 20, 3500. [Google Scholar] [CrossRef] [PubMed]
  103. Accordi, B.; Espina, V.; Giordan, M.; VanMeter, A.; Milani, G.; Galla, L.; Ruzzene, M.; Sciro, M.; Trentin, L.; De Maria, R.; et al. Functional Protein Network Activation Mapping Reveals New Potential Molecular Drug Targets for Poor Prognosis Pediatric BCP-ALL. PLoS ONE 2010, 5, e13552. [Google Scholar] [CrossRef] [PubMed]
  104. Conter, V.; Valsecchi, M.G.; Buldini, B.; Parasole, R.; Locatelli, F.; Colombini, A.; Rizzari, C.; Putti, M.C.; Barisone, E.; Lo Nigro, L.; et al. Early T-Cell Precursor Acute Lymphoblastic Leukaemia in Children Treated in AIEOP Centres with AIEOP-BFM Protocols: A Retrospective Analysis. Lancet. Haematol. 2016, 3, e80–e86. [Google Scholar] [CrossRef] [PubMed]
  105. Bartis, D.; Boldizsár, F.; Szabó, M.; Pálinkás, L.; Németh, P.; Berki, T. Dexamethasone Induces Rapid Tyrosine-Phosphorylation of ZAP-70 in Jurkat Cells. J. Steroid Biochem. Mol. Biol. 2006, 98, 147–154. [Google Scholar] [CrossRef]
  106. Foà, R.; Bassan, R.; Vitale, A.; Elia, L.; Piciocchi, A.; Puzzolo, M.-C.; Canichella, M.; Viero, P.; Ferrara, F.; Lunghi, M.; et al. Dasatinib-Blinatumomab for Ph-Positive Acute Lymphoblastic Leukemia in Adults. N. Engl. J. Med. 2020, 383, 1613–1623. [Google Scholar] [CrossRef]
  107. Harr, M.W.; Caimi, P.F.; McColl, K.S.; Zhong, F.; Patel, S.N.; Barr, P.M.; Distelhorst, C.W. Inhibition of Lck Enhances Glucocorticoid Sensitivity and Apoptosis in Lymphoid Cell Lines and in Chronic Lymphocytic Leukemia. Cell Death Differ. 2010, 17, 1381–1391. [Google Scholar] [CrossRef]
  108. Palacios, E.H.; Weiss, A. Function of the Src-Family Kinases, Lck and Fyn, in T-Cell Development and Activation. Oncogene 2004, 23, 7990–8000. [Google Scholar] [CrossRef]
  109. Lewis, R.S. Calcium Signaling Mechanisms in T Lymphocytes. Annu. Rev. Immunol. 2003, 19, 497–521. [Google Scholar] [CrossRef]
  110. Crabtree, G.R. Generic Signals and Specific Minireview Outcomes: Signaling through Ca2+, Calcineurin, and NF-AT. Cell 1999, 96, 611–614. [Google Scholar] [CrossRef]
  111. Zhao, Y.; Tozawa, Y.; Iseki, R.; Mukai, M.; Iwata, M. Calcineurin Activation Protects T Cells from Glucocorticoid-Induced Apoptosis. J. Immunol. 1995, 154, 6346–6354. [Google Scholar] [CrossRef] [PubMed]
  112. Burns, M.A.; Liao, Z.W.; Yamagata, N.; Pouliot, G.P.; Stevenson, K.E.; Neuberg, D.S.; Thorner, A.R.; Ducar, M.; Silverman, E.A.; Hunger, S.P.; et al. Hedgehog Pathway Mutations Drive Oncogenic Transformation in High-Risk T-Cell Acute Lymphoblastic Leukemia. Leukemia 2018, 32, 2126–2137. [Google Scholar] [CrossRef] [PubMed]
  113. Dagklis, A.; Demeyer, S.; DeBie, J.; Radaelli, E.; Pauwels, D.; Degryse, S.; Gielen, O.; Vicente, C.; Vandepoel, R.; Geerdens, E.; et al. Hedgehog Pathway Activation in T-Cell Acute Lymphoblastic Leukemia Predicts Response to SMO and GLI1 Inhibitors. Blood 2016, 128, 2642–2654. [Google Scholar] [CrossRef] [PubMed]
  114. Regl, G.; Kasper, M.; Schnidar, H.; Eichberger, T.; Neill, G.W.; Philpott, M.P.; Esterbauer, H.; Hauser-Kronberger, C.; Frischauf, A.-M.; Aberger, F. Activation of the BCL2 Promoter in Response to Hedgehog/GLI Signal Transduction Is Predominantly Mediated by GLI2. Cancer Res. 2004, 64, 7724–7731. [Google Scholar] [CrossRef] [PubMed]
  115. Heine, V.M.; Rowitch, D.H. Hedgehog Signaling Has a Protective Effect in Glucocorticoid-Induced Mouse Neonatal Brain Injury through an 11betaHSD2-Dependent Mechanism. J. Clin. Investig. 2009, 119, 267–277. [Google Scholar] [CrossRef]
  116. Martelli, A.M.; Paganelli, F.; Truocchio, S.; Palumbo, C.; Chiarini, F.; McCubrey, J.A. Understanding the Roles of the Hedgehog Signaling Pathway during T-Cell Lymphopoiesis and in T-Cell Acute Lymphoblastic Leukemia (T-ALL). Int. J. Mol. Sci. 2023, 24, 2962. [Google Scholar] [CrossRef]
  117. Deftos, M.L.; He, Y.W.; Ojala, E.W.; Bevan, M.J. Correlating Notch Signaling with Thymocyte Maturation. Immunity 1998, 9, 777–786. [Google Scholar] [CrossRef]
  118. Palomero, T.; Lim, W.K.; Odom, D.T.; Sulis, M.L.; Real, P.J.; Margolin, A.; Barnes, K.C.; O’Neil, J.; Neuberg, D.; Weng, A.P.; et al. NOTCH1 Directly Regulates C-MYC and Activates a Feed-Forward-Loop Transcriptional Network Promoting Leukemic Cell Growth. Proc. Natl. Acad. Sci. USA 2006, 103, 18261–18266. [Google Scholar] [CrossRef]
  119. Kamdje, A.H.N.; Krampera, M. Notch Signaling in Acute Lymphoblastic Leukemia: Any Role for Stromal Microenvironment? Blood 2011, 118, 6506–6514. [Google Scholar] [CrossRef]
  120. Aster, J.C.; Pear, W.S.; Blacklow, S.C. Notch Signaling in Leukemia. Annu. Rev. Pathol. Mech. Dis. 2008, 3, 587–613. [Google Scholar] [CrossRef]
  121. Palomero, T.; Sulis, M.L.; Cortina, M.; Real, P.J.; Barnes, K.; Ciofani, M.; Caparros, E.; Buteau, J.; Brown, K.; Perkins, S.L.; et al. Mutational Loss of PTEN Induces Resistance to NOTCH1 Inhibition in T-Cell Leukemia. Nat. Med. 2007, 13, 1203–1210. [Google Scholar] [CrossRef] [PubMed]
  122. Cramer, S.D.; Aplan, P.D.; Durum, S.K. Therapeutic Targeting of IL-7Rα Signaling Pathways in ALL Treatment. Blood 2016, 128, 473–478. [Google Scholar] [CrossRef] [PubMed]
  123. Okuhashi, Y.; Itoh, M.; Nara, N.; Tohda, S. Effects of Combination of Notch Inhibitor plus Hedgehog Inhibitor or Wnt Inhibitor on Growth of Leukemia Cells. Anticancer Res. 2011, 31, 893–896. [Google Scholar] [PubMed]
  124. Tosello, V.; Bongiovanni, D.; Liu, J.; Pan, Q.; Yan, K.K.; Saccomani, V.; Van Trimpont, M.; Pizzi, M.; Mazzoni, M.; Dei Tos, A.P.; et al. Cross-Talk between GLI Transcription Factors and FOXC1 Promotes T-Cell Acute Lymphoblastic Leukemia Dissemination. Leukemia 2021, 35, 984–1000. [Google Scholar] [CrossRef]
  125. Hou, X.; Chen, X.; Zhang, P.; Fan, Y.; Ma, A.; Pang, T.; Song, Z.; Jin, Y.; Hao, W.; Liu, F.; et al. Inhibition of Hedgehog Signaling by GANT58 Induces Apoptosis and Shows Synergistic Antitumor Activity with AKT Inhibitor in Acute T Cell Leukemia Cells. Biochimie 2014, 101, 50–59. [Google Scholar] [CrossRef]
  126. Bardwell, A.J.; Wu, B.; Sarin, K.Y.; Waterman, M.L.; Atwood, S.X.; Bardwell, L. ERK2 MAP Kinase Regulates SUFU Binding by Multisite Phosphorylation of GLI1. Life Sci. Alliance 2022, 5, e202101353. [Google Scholar] [CrossRef]
  127. Tissing, W.J.E.; Meijerink, J.P.P.; Den Boer, M.L.; Brinkhof, B.; Van Rossum, E.F.C.; Van Wering, E.R.; Koper, J.W.; Sonneveld, P.; Pieters, R. Genetic Variations in the Glucocorticoid Receptor Gene Are Not Related to Glucocorticoid Resistance in Childhood Acute Lymphoblastic Leukemia. Clin. Cancer Res. 2005, 11, 6050–6056. [Google Scholar] [CrossRef]
  128. Bachmann, P.S.; Gorman, R.; MacKenzie, K.L.; Lutze-Mann, L.; Lock, R.B. Dexamethasone Resistance in B-Cell Precursor Childhood Acute Lymphoblastic Leukemia Occurs Downstream of Ligand-Induced Nuclear Translocation of the Glucocorticoid Receptor. Blood 2005, 105, 2519–2526. [Google Scholar] [CrossRef]
  129. Olivas-Aguirre, M.; Torres-López, L.; Pottosin, I.; Dobrovinskaya, O. Overcoming Glucocorticoid Resistance in Acute Lymphoblastic Leukemia: Repurposed Drugs Can Improve the Protocol. Front. Oncol. 2021, 11, 617937. [Google Scholar] [CrossRef]
  130. Abdoul-Azize, S.; Dubus, I.; Vannier, J.P. Modulation of Glucocorticoid Sensitivity in Acute Lymphoblastic Leukemia: Pyr3, a New Therapeutic Tool? Med. Sci. 2017, 33, 130–132. [Google Scholar] [CrossRef]
  131. Polak, A.; Kiliszek, P.; Sewastianik, T.; Szydłowski, M.; Jabłońska, E.; Białopiotrowicz, E.; Górniak, P.; Markowicz, S.; Nowak, E.; Grygorowicz, M.A.; et al. MEK Inhibition Sensitizes Precursor B-Cell Acute Lymphoblastic Leukemia (B-ALL) Cells to Dexamethasone through Modulation of MTOR Activity and Stimulation of Autophagy. PLoS ONE 2016, 11, e0155893. [Google Scholar] [CrossRef] [PubMed]
  132. Follini, E.; Marchesini, M.; Roti, G. Strategies to Overcome Resistance Mechanisms in T-Cell Acute Lymphoblastic Leukemia. Int. J. Mol. Sci. 2019, 20, 3021. [Google Scholar] [CrossRef] [PubMed]
  133. Silveira, A.B.; Laranjeira, A.B.A.; Rodrigues, G.O.L.; Leal, P.C.; Cardoso, B.A.; Barata, J.T.; Yunes, R.A.; Zanchin, N.I.T.; Brandalise, S.R.; Yunes, J.A. PI3K Inhibition Synergizes with Glucocorticoids but Antagonizes with Methotrexate in T-Cell Acute Lymphoblastic Leukemia. Oncotarget 2015, 6, 13105–13118. [Google Scholar] [CrossRef] [PubMed]
  134. Janes, M.R.; Limon, J.J.; So, L.; Chen, J.; Lim, R.J.; Chavez, M.A.; Vu, C.; Lilly, M.B.; Mallya, S.; Ong, S.T.; et al. Effective and Selective Targeting of Leukemia Cells Using a TORC1/2 Kinase Inhibitor. Nat. Med. 2010, 16, 205–213. [Google Scholar] [CrossRef] [PubMed]
  135. Chiarini, F.; Grimaldi, C.; Ricci, F.; Tazzari, P.L.; Evangelisti, C.; Ognibene, A.; Battistelli, M.; Falcieri, E.; Melchionda, F.; Pession, A.; et al. Activity of the Novel Dual Phosphatidylinositol 3-Kinase/Mammalian Target of Rapamycin Inhibitor NVP-BEZ235 against T-Cell Acute Lymphoblastic Leukemia. Cancer Res. 2010, 70, 8097–8107. [Google Scholar] [CrossRef]
  136. Hall, C.P.; Reynolds, C.P.; Kang, M.H. Modulation of Glucocorticoid Resistance in Pediatric T-Cell Acute Lymphoblastic Leukemia by Increasing BIM Expression with the PI3K/MTOR Inhibitor BEZ235. Clin. Cancer Res. 2016, 22, 621–632. [Google Scholar] [CrossRef]
  137. Meyer, L.K.; Delgado-Martin, C.; Sharp, P.P.; Huang, B.J.; McMinn, D.; Vincent, T.L.; Ryan, T.; Horton, T.M.; Wood, B.L.; Teachey, D.T.; et al. Inhibition of the Sec61 Translocon Overcomes Cytokine-Induced Glucocorticoid Resistance in T-Cell Acute Lymphoblastic Leukaemia. Br. J. Haematol. 2022, 198, 137–141. [Google Scholar] [CrossRef]
  138. Tian, Y.; Ai, H.; Ji, X.; Wei, X.D.; Song, Y.P.; Yin, Q.S. Efficacy Analysis of Venetoclax Combined with TKI and Dexamethasone-Containing Low-Dose Chemotherapy for Relapsed/Refractory Ph+acute B-Lymphoblastic Leukemia. Zhonghua Yi Xue Za Zhi 2022, 102, 745–748. [Google Scholar] [CrossRef]
  139. Kawashima-Goto, S.; Imamura, T.; Tomoyasu, C.; Yano, M.; Yoshida, H.; Fujiki, A.; Tamura, S.; Osone, S.; Ishida, H.; Morimoto, A.; et al. BCL2 Inhibitor (ABT-737): A Restorer of Prednisolone Sensitivity in Early T-Cell Precursor-Acute Lymphoblastic Leukemia with High MEF2C Expression? PLoS ONE 2015, 10, e0132926. [Google Scholar] [CrossRef]
  140. Hui, P.Y.; Chen, Y.H.; Qin, J.; Jiang, X.H. PON2 Blockade Overcomes Dexamethasone Resistance in Acute Lymphoblastic Leukemia. Hematology 2022, 27, 32–42. [Google Scholar] [CrossRef]
  141. Cialfi, S.; Palermo, R.; Manca, S.; Checquolo, S.; Bellavia, D.; Pelullo, M.; Quaranta, R.; Dominici, C.; Gulino, A.; Screpanti, I.; et al. Glucocorticoid Sensitivity of T-Cell Lymphoblastic Leukemia/Lymphoma Is Associated with Glucocorticoid Receptor-Mediated Inhibition of Notch1 Expression. Leukemia 2013, 27, 485–488. [Google Scholar] [CrossRef]
  142. Real, P.J.; Ferrando, A.A. NOTCH Inhibition and Glucocorticoid Therapy in T-Cell Acute Lymphoblastic Leukemia. Leukemia 2009, 23, 1374–1377. [Google Scholar] [CrossRef] [PubMed]
  143. Real, P.J.; Tosello, V.; Palomero, T.; Castillo, M.; Hernando, E.; De Stanchina, E.; Sulis, M.L.; Barnes, K.; Sawai, C.; Homminga, I.; et al. γ-Secretase Inhibitors Reverse Glucocorticoid Resistance in T Cell Acute Lymphoblastic Leukemia. Nat. Med. 2008, 15, 50–58. [Google Scholar] [CrossRef]
  144. Bachmann, P.S.; Gorman, R.; Papa, R.A.; Bardell, J.E.; Ford, J.; Kees, U.R.; Marshall, G.M.; Lock, R.B. Divergent Mechanisms of Glucocorticoid Resistance in Experimental Models of Pediatric Acute Lymphoblastic Leukemia. Cancer Res. 2007, 67, 4482–4490. [Google Scholar] [CrossRef] [PubMed]
  145. Gao, J.; Liu, W.J. Prognostic Value of the Response to Prednisone for Children with Acute Lymphoblastic Leukemia: A Meta-Analysis. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 7858–7866. [Google Scholar] [CrossRef]
  146. Irving, J.A.E. Towards an Understanding of the Biology and Targeted Treatment of Paediatric Relapsed Acute Lymphoblastic Leukaemia. Br. J. Haematol. 2016, 172, 655–666. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Proposed model for ALL cell fate following GC treatment. In this model, GC, by directly regulating crucial death or survival genes, leads to apoptotic cell death, but at the same time, paradoxically promotes ALL cell survival via several pathways (for other details see text).
Figure 1. Proposed model for ALL cell fate following GC treatment. In this model, GC, by directly regulating crucial death or survival genes, leads to apoptotic cell death, but at the same time, paradoxically promotes ALL cell survival via several pathways (for other details see text).
Cancers 15 02812 g001
Figure 2. Proposed model for the mechanism by which GC induce their own resistance ALL cells. Dex treatment leads to GR translocation, which induces a transcriptional program resulting in lymphoid cell death (therapeutic effect). Along with this pathway, Dex also provokes cytosolic Ca2+ release as well as IL-7R upregulation and subsequent triggering of a pro-survival activity (paradoxical effect). This latter pathway may blunt the anticancer efficacy of chemotherapy (for other details see text).
Figure 2. Proposed model for the mechanism by which GC induce their own resistance ALL cells. Dex treatment leads to GR translocation, which induces a transcriptional program resulting in lymphoid cell death (therapeutic effect). Along with this pathway, Dex also provokes cytosolic Ca2+ release as well as IL-7R upregulation and subsequent triggering of a pro-survival activity (paradoxical effect). This latter pathway may blunt the anticancer efficacy of chemotherapy (for other details see text).
Cancers 15 02812 g002
Table 1. GC used in leukemia therapy.
Table 1. GC used in leukemia therapy.
NameGC PotencyMineral Corticoid PotencyPlasma Half-lifeDosing in TreatmentReferences
Prednisone3.5–40.812–36 h40–60 mg/m2[14,15]
Dexamethasone25–80036–54 h6–10 mg/m2[14]
Table 2. Clinical trials utilizing GC.
Table 2. Clinical trials utilizing GC.
TrialsnGCDoseDuration of TreatmentDiseaseStudy Arm/GroupResultsRefs

NCT00613457

2039

Dex

Pred

10 mg/m2/day


60 mg/m2/day

14 days


20 days

ALL


Pred vs. Dex
Dex reduced the incidence of better salvageable relapses
Significant survival benefit from dex only for patients with T-cell ALL





[16]


NCT03390387


4000
Dex


Pred
6 mg/m2/day


60 mg/m2/day
28 days or day 1–15 and 22–29

28 days



ALL
Dex intermittent vs Dex continue vs. Pred


Recruiting


NA

NCT00002816
Phase 3


120


Dex

-

-

Relapsed ALL
Early relapse vs. late relapse
for drugs association
After reinduction, LPC counts were lower than in patients treated for an overt BM first relapse.

[17,18]
NCT00707083
Phase 3

2231

Dex

-

Days 1–5 and 29–33

ALL
Bone marrow suppression and liver toxicity
[19]


(CCG)-1922


1060
Dex


Pred
6 mg/m2/day
40 mg/m2/day
28 days
28 days

ALL
6 MP + Oral Pred vs. 6 MP + Intravenous Pred vs. 6 MP + Oral dex vs. 6 MP + Intravenous Dex
Dex provided a 34% reduction in risk of relapse


[20]


AALL0232



3154
Dex


Pred
10 mg/m2/day
60 mg/m2/day
14 days
28 days

High risk B-ALL


Pred vs. Dex
Higher rate of subsequent osteonecrosis with Dex-treated patient

[21]


NCT00003728

1947
Dex


Pred
6 mg/m2/day
60 mg/m2/day

-

ALL



Pred vs. Dex
EVS similar
Decreased cumulative incidence of CNS relapse with dex.

[22]
NCT01324180
14Dex10 mg/m2/day-Relapsed/ refractory ALL

Met + VPLD

-
NA

NCT03613428

12

Pred

1 mg/kg

28 days

Relapse/refractory T-ALL

Rux + Pred

-

NA

NCT03817320



31


Dex


10 mg/m2/day


14 days

Relapsed /refractory ALL


Ixa + VXLD


Recruiting


NA
n: number of patients; Dex: Dexamethasone; Pred: Prednisone; Met: Metformin; Rux: Ruxolitinib; VLPD: Vincristine, Dexamethasone, Doxorubicin, and PEG-asparaginase; Ixa: Ixazomib; VXLD: Vincristine, Dexamethasone, Asparaginase, and Doxorubicin; EFS: Event free survival; CNS: central nervous system; NA: no publication available; LPC: leukemic progenitor cell.
Table 3. GC-activated signaling pathways in ALL whose inhibition overcomes GC resistance or increases GC sensitivity.
Table 3. GC-activated signaling pathways in ALL whose inhibition overcomes GC resistance or increases GC sensitivity.
PathwaysDrugActivation ModeInhibitor (s)Leukemic CellsConsequencesReferences

Ca2+ signaling

Dex

intracellular release
BAPTA-AM
Pyr3
thapsigargin
Reh

Nalm-6

increases GC sensitivity

[37,38]

IL7R signaling

Dex

IL-7R upregulation

Ruxolitinib
JAK3i

CCRF-CEM
Patient samples
PDX samples

overcomes GC resistance

[39,40]

BCL2 signaling

Dex

BCL2 upregulation

ABT-199
Ruxolitinib
Tofacitinib

PDX samples
Patient samples

overcomes GC resistance

[39,41,42]



LCK signaling



Dex/Pred


LCK phosphorylation

Dasatinib
Bosutinib
Nintedanib
WH-4-023
shRNA
Patient samples
CCRF-CEM
Jurkat
TALL-1
SUPT1
LK203
PDX samples



overcomes GC resistance



[2,43,44,45,46,47]


AKT signaling


Dex


AKT phosphorylation


MK2206
Akt inhibitor IV

Patient samples
CCRF-CEM
MOLT3
PF382



increases GC sensitivity


[11,48]

ERK signaling

JNK signaling


Dex

ERK, JNK
phosphorylation

U0126

SP600125
ip

CEM-C1-15
R3F9


overcomes GC resistance


[49,50]

Hedgehog signaling

Dex
Hh activation via GLI1 and PTCH mRNAsmPKI
GANT61
CEM-C7-14
T-ALL cell lines
PDX samples


increases GC sensitivity


[51,52]
Ca2+: Calcium; ip: JNK inhibitory peptide; Dex: dexamethasone; Pred: prednisolone; PDX: Patient-derived xenografts.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Angot, L.; Schneider, P.; Vannier, J.-P.; Abdoul-Azize, S. Beyond Corticoresistance, A Paradoxical Corticosensitivity Induced by Corticosteroid Therapy in Pediatric Acute Lymphoblastic Leukemias. Cancers 2023, 15, 2812. https://doi.org/10.3390/cancers15102812

AMA Style

Angot L, Schneider P, Vannier J-P, Abdoul-Azize S. Beyond Corticoresistance, A Paradoxical Corticosensitivity Induced by Corticosteroid Therapy in Pediatric Acute Lymphoblastic Leukemias. Cancers. 2023; 15(10):2812. https://doi.org/10.3390/cancers15102812

Chicago/Turabian Style

Angot, Laure, Pascale Schneider, Jean-Pierre Vannier, and Souleymane Abdoul-Azize. 2023. "Beyond Corticoresistance, A Paradoxical Corticosensitivity Induced by Corticosteroid Therapy in Pediatric Acute Lymphoblastic Leukemias" Cancers 15, no. 10: 2812. https://doi.org/10.3390/cancers15102812

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop