Next Article in Journal
Sugar Fructose Triggers Gut Dysbiosis and Metabolic Inflammation with Cardiac Arrhythmogenesis
Next Article in Special Issue
How to Improve Prognostication in Acute Myeloid Leukemia with CBFB-MYH11 Fusion Transcript: Focus on the Role of Molecular Measurable Residual Disease (MRD) Monitoring
Previous Article in Journal
Postnatal Expression Profile of MicroRNAs Associated with Cardiovascular Diseases in 3- to 11-Year-Old Preterm-Born Children
Previous Article in Special Issue
Pharmacological Inhibition of WIP1 Sensitizes Acute Myeloid Leukemia Cells to the MDM2 Inhibitor Nutlin-3a
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RSK Isoforms in Acute Myeloid Leukemia

Department of Pediatrics, Stanford University School of Medicine, Stanford, CA 94305, USA
*
Author to whom correspondence should be addressed.
Biomedicines 2021, 9(7), 726; https://doi.org/10.3390/biomedicines9070726
Submission received: 28 May 2021 / Revised: 20 June 2021 / Accepted: 22 June 2021 / Published: 24 June 2021

Abstract

:
Ribosomal S6 Kinases (RSKs) are a group of serine/threonine kinases that function downstream of the Ras/Raf/MEK/ERK signaling pathway. Four RSK isoforms are directly activated by ERK1/2 in response to extracellular stimuli including growth factors, hormones, and chemokines. RSKs phosphorylate many cytosolic and nuclear targets resulting in the regulation of diverse cellular processes such as cell proliferation, survival, and motility. In hematological malignancies such as acute myeloid leukemia (AML), RSK isoforms are highly expressed and aberrantly activated resulting in poor outcomes and resistance to chemotherapy. Therefore, understanding RSK function in leukemia could lead to promising therapeutic strategies. This review summarizes the current information on human RSK isoforms and discusses their potential roles in the pathogenesis of AML and mechanism of pharmacological inhibitors.

1. Introduction

The Ras-mitogen-activated protein kinase (MAPK) pathway is involved in the regulation of normal cell proliferation, survival, growth, and differentiation [1,2]. More than 30% of all human cancers are associated with abnormal control of this signaling network, resulting in gain of function and subsequent extracellular signal-regulated kinase (ERK) hyperactivation [3]. The Ras-MAPK pathway is initiated by a ligand binding to the receptor tyrosine kinase (RTK) receptor, followed by docking adaptor proteins such as growth factor receptor-bound protein 2 (GRB2) and Son of Sevenless (SOS), leading to activation of the associated Rat sarcoma (Ras) and recruitment of Raf. Raf then activates downstream mitogen-activated protein kinases kinase (MEK1/2) and ERK1/2 (Figure 1) [2].
RSKs are a group of serine/threonine kinases that function in the MAPK signaling cascade and are the direct downstream effectors of ERK1/2. Four RSK isoforms are directly activated by ERK1/2 in response to extracellular stimuli including growth factors, hormones, and chemokines [4,5]. RSKs phosphorylate many cytosolic and nuclear targets resulting in the regulation of diverse cellular processes such as cell proliferation, survival, and motility. Therefore, the RSK isoforms represent attractive therapeutic targets for cancer [6,7]. Here we review the general roles of RSK isoforms and discuss their potential roles in AML and current pharmacological tools to inhibit their function.

2. RSK Isoforms

Four RSK isoforms (RSK1-4) have been identified in humans (Figure 2). They share 75–80% homology in amino acid sequence and several conserved functional motifs. Two distinct functional kinase domains are connected by conserved linker regions (turn motif and hydrophobic motif) and are flanked by N- and C-terminal tails [8,9]. The C-terminal kinase domain (CTKD) is involved in auto-phosphorylation of N-terminal kinase domain (NTKD) [9,10,11], and the NTKD is responsible for phosphorylation of downstream substrates. The C-terminal tail contains two conserved motifs: the kinase interaction motif (KIM) for ERK1/2 docking site [12,13,14] and the type 1 PDZ domain-binding motif for the interaction with PDZ domain-containing proteins [15]. RSK3 possesses a potential nuclear localization signal in N-terminal tail.
Inactive RSK isoforms are localized in cytoplasm. Upon stimulation, RSKs are phosphorylated by ERK1/2 and translocate into nucleus for the induction of immediate-early gene expression. RSK4 is distinct from RSK1-3 as it is constitutively active, predominantly localized in cytosol, and exhibits growth-factor-independent kinase activity [16].
RSK isoforms are ubiquitously expressed in every human tissue and brain region with the exception of RSK4, which is mostly expressed during embryonic development [4,17]. This supports the current view that RSKs are functionally redundant. However, tissue-specific variations in expression levels of RSK isoforms have also been reported, suggesting that they might have isoform specificity in mediating distinct cellular functions [4,18,19]. RSK1 is predominantly expressed in the lung, bone marrow, and T cells; RSK2 is more abundant in T cells, lymph nodes, and prostate. RSK3 has high expression in the lung, brain, spinal cord, and retina. RSK4 transcript is much lower than that of other isoforms.

3. Regulation of RSK Isoforms

RSKs are directly or indirectly phosphorylated by several kinases in the Ras/Raf/MEK/ERK signal cascade. All human RSK isoforms have six conserved phosphorylation sites: Ser221, Thr359, Ser363, Ser380, Thr573, and Ser732, respective to RSK1 (Figure 3) [20].
In quiescent cells, RSKs are in complex with inactive ERK1/2 through kinase interaction motif (KIM) ERK-docking site. Following the stimulation of cells, ERK1/2 phosphorylate Thr573 within the CTKD activation loop leading to full activation of CTKD [12,21]. ERK1/2 might also phosphorylate Ser363 and Thr359 in the conserved linker region. The fully activated CTKD then auto-phosphorylates Ser380 within the hydrophobic motif of the linker region [11], promoting recruitment of 3′-phosphoinositide-dependent kinases-1 (PDK1) [22]. After binding, PDK1 phosphorylates Ser221 within the NTKD [23,24], leading to the full activation of NTKD and subsequent phosphorylation of the substrates [25]. NTKD also auto-phosphorylates Ser732 located next to ERK1/2 docking site, decreasing the ERK affinity for RSKs in a negative feedback loop [14]. De-phosphorylation of RSKs by protein phosphatases PP2Cdelta has been also reported as the other inactivation mechanism of RSK isoforms [26].
More studies provide evidence for alternative RSK2 activation mechanisms through the p38 MAPK and fibroblast growth factor receptor-3 (FGFR3) [27,28,29]. p38 MAPK-regulated kinases MK2/3 (MAPKAPK2/3) function as potential CTKD activation in dendritic cells [27]. FGFR3 and Scr family kinases phosphorylate RSK2 at Tyr529 to promote inactive ERK recruitment and Tyr707 to potentially disrupt the auto-inhibitory motif [28,29,30].

4. The Function of RSK Isoforms

Analysis of RSK1 substrates by using synthetic peptide libraries has identified the minimum consensus sequences (Arg/Lys-Xaa-Arg-Xaa-Xaa-pSer/Thr or Arg-Arg-Xaa-pSer/Thr). Although RSK1 preferentially phosphorylates this motif, there are diverse substrates containing different targeting sequences from this motif [2,31].
Through regulating extensive downstream substrates in transcriptional gene expression and directly affecting protein function, RSKs are involved in various cellular processes such as protein synthesis, proliferation, cell survival, and motility (Figure 4) [4,5,18]. In addition, RSKs have critical roles in cancer initiation by regulating cell proliferation and survival as well as cancer development/progression by regulating metastasis, cancer stem cell, and chemoresistance [6].

4.1. Protein Synthesis

Many findings support that RSKs are involved in translational regulation through the phosphorylation of several ribosome-associated proteins. One such way is through modulating the mammalian target of rapamycin (mTOR) pathway. RSKs phosphorylate tuberous sclerosis complex 2 (TSC2) and Raptor to promote mTOR complex 1 (mTORC1)-mediated translation [32,33,34,35]. RSKs also phosphorylate and inhibit glycogen synthase kinase-3 (GSK3), causing the stimulation of the translation initiation factor eIF2B [36]. Independent from the mTOR pathway, RSKs directly phosphorylate the eIF4B to stimulate the translation initiation of growth- and survival-related transcripts [37,38]. Recently, RSKs are also reported to promote the degradation of programmed cell death 4 (PDCD4), a negative regulator of the eIF4A helicase [39,40]. Moreover, RSKs facilitate the assembly of the translation preinitiation complex and increase cap-dependent translation by phosphorylating the 40 s ribosomal protein s6 (RPS6) [41].

4.2. Cell Cycle and Proliferation

There is much evidence that RSKs promote cell proliferation through modulating the components of cell cycle machinery. RSKs regulate G1 progression by phosphorylating c-Fos, GSK3, and p27Kip1. RSK2-dependent stabilization of c-Fos is essential to increase the cyclin D1 expression and thereby facilitate G1-S phase progression [42,43,44]. Phosphorylation on p27Kip1 by RSK1 suppresses its inhibition activity on the cyclin E/A-CDK2 complexes [45,46]. GSK3 normally targets c-Myc and cyclin D1 degradation. Therefore, inhibition of GSK3 activity by RSKs facilitates G1 phase progression [47,48]. RSKs also promote G2-M transition by phosphorylating and activating CDC25C involved in CDK1-cyclin B complex activation [49].
RSKs have been shown to promote cell proliferation through phosphorylating Max dimerization protein-1 (Mad1), a suppressor of Myc-mediated cell proliferation and transformation. Phosphorylated Mad1 undergoes proteasomal degradation and thereby alleviating its suppression of Myc [50]. RSKs also phosphorylate cAMP response element-binding protein (CREB), which is a critical transcription factor for cell proliferation and survival through regulation of c-Fos expression [51].

4.3. Cell Survival

RSK-mediated phosphorylation involves the inactivation of pro-apoptotic proteins and the activation of transcription factors controlling cell survival promotion. In response to growth factors, RSKs directly phosphorylate pro-apoptotic proteins BCL2-associated agonist of cell death (BAD) and death-associated protein kinase (DAPK) [1,52,53,54]. This results in their decreased activities and increased cell survival. In addition, RSKs inhibit caspase activity by phosphorylating the CCAAT/enhancer binding protein beta (C/EBPβ) [55]. Phosphorylated C/EBPβ creates a functional XEVD caspase inhibitory box to inhibit caspases 1 and 8 activities. In addition to the function of cell proliferation, RSK-activated CREB promotes cell survival by increasing the transcription of pro-survival genes such as B-cell lymphoma 2 (BCL2), B-cell lymphoma-extra large (BCL-XL), and myeloid cell leukemia-1 (MCL-1) [1,56].

4.4. Cell Motility and Metastasis

RSKs are involved in cell migration, invasion, and metastasis in various cell types through both transcriptional and protein regulation [7]. RSKs activated by ERK1/2 phosphorylate the transcriptional factor Fos-related antigen 1 (FRA1) to promote the expression of pro-metastatic genes such as Integrin, Laminin332, matrix metalloproteinase (MMP), and Rac1 [57]. RSKs also phosphorylate Filamin A and thereby promote cell migration and suppress integrin activation and cell adhesion [58,59]. Finally, RSKs phosphorylate and inactivate the SH3 domain-containing protein (SH3P2), a negative regulator of cell motility [60].

4.5. Cancers

The aberrant expression and activity of RSKs has been associated with uncontrolled proliferation and prolonged survival in several cancer types. In addition, RSKs are found to phosphorylate cell cycle checkpoint components such as checkpoint kinase 1 (Chk1) and Mer11 to suppress DNA damage signaling and possibly increase cancer cell chemoresistance [61,62].
There is evidence that RSK isoforms have functional differences, especially in cancer [6]. RSK1/2 promote cancer growth, proliferation, and survival. The expression or activation of RSK1/2 appears to be increased in many cancers including lung cancer [63,64], head and neck squamous cell carcinoma (HNSCC) [65], breast cancer [40,66], prostate cancer [67], leukemia [68], melanoma [39,69], multiple myeloma [28], and glioblastoma [58,70]. The expression and activation of RSKs in lung cancer inhibits cell death through inactivation of a pro-apoptotic protein BAD [64]. The loss of RSK2 function leads to decreased phosphorylation of CREB and Hsp27, and thereby substantially reduces invasiveness of HNSCC cells [65].
In contrast, RSK3/4 may act conversely as tumor suppressors. Reduced expression of RSK3/4 has been observed in various cancer types such as ovarian cancer [71], colorectal cancer [72], acute myeloid leukemia [73], and breast cancer [74,75]. Low expression of RSK4 is predictive of a poor prognosis in patients with colorectal cancer [72]. In ovarian cancer cell lines, overexpression of RSK3 was found to decease their proliferation [71]. Overexpression of RSK4 decreased the proliferation of breast cancer by accumulating cells in the G0/G1 phase [75].
Moreover, RSK4 has been shown to participate in p53-dependent cell growth arrest [76] and in oncogene-induced cellular senescence in colon and renal cell carcinomas [77]. However, it has also been reported that RSK4 is overexpressed in >50% of primary malignant lung cancers [64], and RSK3/4 can mediate tumor resistance to PI3 kinase inhibitors in breast cancer [78]. Overall, more in-depth studies are required to fully understand the role and associated molecular mechanisms of action of the different RSK isoforms in cancer.

5. RSK Isoforms in AML

AML is a genetically and phenotypically heterogeneous hematological malignancy characterized by the accumulation of immature myeloid blasts with peripheral blood cytopenia [79,80]. AML patients have an overall survival of less than 65% in children and 40% in adults. The current treatment options, including intensive chemotherapy and stem cell transplantation, are associated with significant morbidity and mortality [81]. Thus, it is critical to develop more effective and less toxic therapies for AML.
The Ras/Raf/MEK/ERK pathway has been reported to be constitutively activated in more than 50% of AML and acute lymphocytic leukemia (ALL) cases [82,83]. Thus, RSK isoforms play an important role in AML pathogenesis and progression. In addition, it has been observed that RSK1/2 are the predominant isoforms expressed in AML cells, whereas RSK4 has shown significantly lower expression in AML patients compared to healthy people. This suggests that downregulated RSK4 expression may lead to leukemia or negatively affect the prognosis of patients with AML [73]. We discuss below the roles of RSKs in AML pathogenesis (Figure 5) and how RSKs could be a therapeutic target for AML treatment.

5.1. RSKs and CREB in AML Cell Survival

Our group has previously observed that approximately 60% of AML patients express CREB at high levels, and this is associated with an increased risk of relapse and decreased event-free survival [84,85]. Similarly, the expression and phosphorylation levels of RSKs are significantly increased in pediatric AML patients associated with poor survival [86,87]. Furthermore, we have observed that RSKs phosphorylate CREB on Ser133, and then phosphorylated CREB mediates proliferation and survival of myeloid cells through induced expression of Bcl-2, cyclin A, and cyclin D [84,86]. In addition, we have recently established that RSK inhibition inhibits AML cell proliferation through the regulation of mitotic exit [87]. A potent RSK inhibitor, BI-D1870, increases metaphase arrest by preventing the metaphase/anaphase transition, followed by induced apoptosis of AML patient cells through impeded association of cell division cycle 20 (CDC20) with anaphase promoting complex/cyclosome (APC/C) and increased mitotic arrest deficient 1 (MAD2) and CDC20 binding. Moreover, BI-D1870 treatment potentiates the anti-leukemic activity of vincristine via synergistically increased mitotic arrest and apoptosis in AML cells. Therefore, our findings suggest a novel therapy that overcomes vincristine resistance to AML cells.

5.2. RSKs in FLT3-ITD+ Cell Survival

FLT3 is a receptor-tyrosine kinase expressed on hematopoietic progenitor cells and plays an important role in proliferation, survival, and differentiation of these cells [88,89]. The internal tandem duplication mutation in FLT3 (FLT3-ITD) is the most frequent mutation in AML found in 25–30% of cases and associated with a poor prognosis [90]. FLT3-ITD leads to constitutive activation of the Ras/Raf/MEK/ERK pathway [82,83]. As a downstream regulator of this pathway, RSKs have essential roles in the pathogenesis and myeloid lineage determination of FLT3/ITD-induced hematopoietic transformation. Activated RSK1 phosphorylates and inactivates pro-apoptotic BAD protein, preventing the apoptosis of BaF/FLT3-ITD cells [91]. Inhibition of RSK1 expression reduces BAD phosphorylation resulting in the induced apoptosis of MV4-11, a cell line harboring FLT3-ITD. Targeting RSK2 by FMK, a RSK inhibitor, attenuates cell viability and induces significant apoptosis in human primary FLT3-ITD+ leukemic cells [68]. These findings suggest that combined inhibition of FLT3 and RSKs may be a viable therapeutic strategy to cure AML patients with FLT3-ITD.
In addition, recent studies have reported that RSKs have pro-survival functions as a new target of Pim2 kinase in relapsed FLT3-ITD+ AML cells [92,93]. Pim2 is a downstream target of FLT3-ITD+ AML cells and directly contributes to FLT3 inhibitor resistance [93]. Ectopic expression of RSK2 rescues the viability of Pim2-depleted cells through the regulation of Bax expression. These support the involvement of RSK2 in AML cell survival as a downstream of Pim2 and a novel therapeutic strategy against therapy-resistant FLT3-ITD+ AML.
Furthermore, it has been reported that FLT3-ITD activates RSK1 to enhance proliferation and survival of AML cells by activating mTORC1 and eIF4 [94]. Activated RSK1/2 via FLT3-ITD and MEK/ERK pathways phosphorylate TSC2 and eIF4B in cooperation with Pim2, thus activating mTORC1/S6K/4EBP1 pathway resulting in enhanced proliferation.

5.3. Alternative RSK Isoform Activation in FGFR3-Activated Cells

It has been shown that RSK2 has a role in hematopoietic transformation of AML and multiple myeloma via an alternative mechanism of RSK activation [28,30]. FGFR3 directly phosphorylates Tyr529 on RSK2, which facilitates inactive ERK1/2 binding to RSK2, and consequently phosphorylates and activates RSK2 [28]. FGFR3 additionally phosphorylates Tyr707 on RSK2 that may disrupt the auto-inhibitory αL-helix motif on C-terminal [30]. Targeted inhibition of RSK2 effectively induced apoptosis in FGFR3-expressing myeloma cells, suggesting that RSK2 is a critical signaling effector in FGFR3-mediated hematopoietic transformation.

5.4. RSKs in the Resistance to As2O3 Treatment

Arsenic trioxide (As2O3) treatment is an effective therapy for acute promyelocytic leukemia (APL) which is a subtype of AML, but shows no significant clinical activity in other non-APL subtype refractory or relapsed AML cases [95]. Recent studies have implicated that RSK1 is involved in the resistance of AML to As2O3 [96,97]. RSK1 is phosphorylated and activated during As2O3 treatment in different AML cell lines. This suggests that RSK1 counteracts As2O3-dependent anti-leukemic response by being activated in a negative feedback regulatory manner. Combined treatment of RSK1 inhibitor with As2O3 was found to result in more potent suppression of leukemic cells. These results suggest that RSK1 is a potentially important target to enhance the anti-leukemic properties of As2O3.

5.5. RSKs with Inhibition of SHH Signaling

Sonic Hedgehog (SHH) signaling is implicated in drug resistance for a range of human cancers [98]. A recent study showed that RSK inhibition overcomes resistance to inhibition of the SHH pathway in pediatric medulloblastoma [99]. In addition, SHH signaling has been found to play a role in the self-renewal of leukemia stem cells (LSCs) for CML and multiple myeloma [100,101,102]. These findings suggest that the combination of RSK and SHH inhibitors may be a complementary anti-leukemia strategy especially through regulating LSCs.

6. Current RSK Inhibitors

As RSKs play important roles in various cancers including AML through disease-related signaling pathways, RSK inhibition provides a promising therapeutic strategy for many diseases. So far, various RSK inhibitors are reported to target the NTKD or CTKD on RSK isoforms (Table 1) [103]. However, none of them is currently useful for in vivo use in cancer models. Furthermore, current RSK inhibitors target more than one RSK isoforms, suggesting the limitation of their efficacy as anticancer agents. As isoform-specific function of RSK is suggested in various cancers, more efficient and specific RSK inhibitors need to be developed for molecular therapy medicine in the future.

6.1. SL0101

The flavonol rhamnoside SL0101 is a cell-permeable kaemperfol glycoside isolated from the tropical plant Fosteronia refracta. SL0101 is the first identified pan-RSK inhibitor. This is shown to target the NTKD of RSK1 and RSK2 in the nanomolar range (IC50 of 89 nM at 100μM ATP for RSK2) while having no significant effects on other AGC kinases [57,104]. SL0101 treatment has been shown to impair the growth of MCF7 breast cancer cell but not of MCF-10A normal breast epithelial cells [104]. However, this has shown a higher EC50 in vivo and lower stability. Although several analogues have been developed [105], poor pharmacokinetic (PK) properties and off-target effects have limited their development.

6.2. BI-D1870

The dihydropteridinone BI-D1870 is the pan-RSK inhibitor as an ATP antagonist to target NTKD [106]. BI-D1870 completely abrogates activities of all RSK isoforms and has selectivity for RSKs relative to other AGC kinases [107]. However, BI-D1870 also significantly inhibits polo like kinase 1 (PLK1) with similar potency and Aurora B, maternal embryonic leucine zipper kinase (MELK), PIM3, mammalian sterile 20-like kinase 2 (MST2), and GSK3β activities at 10- to 100-fold higher concentration [107].
We have observed that 5μM of BI-D1870 specifically inhibits the proliferation of AML primary cells as well as HL60 AML cell line [87]. Among four RSK isoforms, BI-D1870 shows 2-fold lower potency toward RSK3 and RSK4 as compared to RSK1 and RSK2 (IC50 of 15 nM for RSK4, 18 nM for RSK3, 24 nM for RSK2, and 31 nM of RSK1) [106]. This suggests the structural divergence of RSK isoforms that could be exploited for design of isoform-specific inhibitors. Similar with SL0101, BI-D1870 has shortcomings including poor PK profiles and non-specific interaction. There are a few structure–activity relationship (SRA) studies based on BI-D1870 to improve the limitations: substituted pteridinones such as the difluorophenyl pyridine derivatives (LJH685, LJI308) and a series of substituted pteridinones and pyrimidines [108].

6.3. LJH685 and LJI308

LJH685 and LJI308 are difluorophenyl pyridine derivatives of BI-D1870. These have IC50 of 4–13 nM range against all RSK isoforms and, more importantly, have much fewer off-target effects than BI-D1870 [109]. However, it has been reported that LJH685 exhibits high clearance, short plasma half-life, and mild tissue distribution in rats, which should be addressed in further studies [110].

6.4. BIX02565

BIX02565 has been reported to target RSK2 with IC50 of 1.1 nM [111]. It is unclear whether BIX02565 inhibits other RSK isoforms. However, this also has poor selectivity as it also inhibits leucine rich repeat kinase 2 (LRRK2), rearranged during transfection (RET), ded2-like kinase 2 (CLK2), FLT3, and platelet-derived growth factor receptor (PDGFR) [112]. Furthermore, BIX02565 has shown to cause severe decrease in mean heart rate and arterial pressure in a mouse model [113].

6.5. BRD7389

BRD7389 is a RSK inhibitor that induces insulin expression and differentiation of pancreatic alpha into beta cells [114,115]. However, this inhibits RSKs with relatively higher IC50 of 1.5 μM for RSK1, 2.4 μM for RSK2, and 1.2 μM for RSK3 [114]. Moreover, BRD7389 also inhibits CDK5, death-associated protein kinase-related apaptosis-inducing protein kinase 1 (DRAK1), FLT2, PIM1, and protein kinase G 1 α (PKG1α) in very lower IC50 of 2.8–6.5 μM.

6.6. FMK

Pyrrolopyrimidine FMK (fluoromethyl ketone) is an irreversible RSK inhibitor that binds to cysteine residue on the ATP-binding pocket of CTKD [116]. This inhibits RSK phosphorylation on Ser386 and downstream singling with IC50 of 15 nM for RSK2 [117]. However, FMK also inhibits Lck, Src, EphA2, and S6K1 with greater concentration [117]. Besides, FMK inhibits only the activating process of RSK, as CTKD is not necessary once the NTKD is active through other kinases such as p38.

6.7. PMD-026

PMD-026 is the first orally bioavailable small molecule inhibitor targeting RSK. It has been developed to treat triple negative breast cancer (TNBC) and is currently in Phase1/1b clinical trial for advanced breast cancer [118]. PMD-026 demonstrates high specificity for the four RSK isoforms in vitro (IC50 0.7–2 nM) with good selectivity as well as in vivo efficacy in mouse xenograft tumor models of TNBC. Moreover, PMD-026 did not cause any apparent cardiotoxicity, neutropenia, or ocular toxicity in mice and dogs, setting this inhibitor apart from other RSK inhibitors. In addition, preclinical data show that PMD-026 is effective alone or in combination with conventional chemotherapies. PMD-026 in combination with chemotherapy has the potential to become a platform technology for a wide range of refractory cancers.

7. Conclusions

The roles of RSKs in diverse cellular processes are context-dependent and complex. RSKs have important roles in AML pathogenesis, so they are attractive therapeutic targets for AML. However, current RSK inhibitors are not useful as cancer therapies due to limitations such as off-target effects and poor PK properties. Emerging evidence suggests that RSKs have isoform-specific functions. Thus, a critical goal for future research is to develop not only more RSK inhibitors, but isoform-selective inhibitors. Further studies may aid in the identification of RSK substrates and evaluate the role of RSK isoforms and RSK inhibition as anti-leukemic strategy.

Author Contributions

M.Y., J.O.G., K.M. and K.M.S. wrote and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Stanford Maternal Child Health Research Institute and Hyundai Hope on Wheels Grant. K.M.S. is funded by the Pediatric Cancer Research Foundation, Department of Defense, NIDDK/NIH, Leukemia & Lymphoma Society, and Cure Childhood Cancer.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bonni, A.; Brunet, A.; West, A.E.; Datta, S.R.; Takasu, M.A.; Greenberg, M.E. Cell survival promoted by the Ras-MAPK signaling pathway by transcription-dependent and -independent mechanisms. Science 1999, 286, 1358–1362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Cargnello, M.; Roux, P.P. Activation and function of the MAPKs and their substrates, the MAPK-activated protein kinases. Microbiol. Mol. Biol. Rev. 2011, 75, 50–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hoshino, R.; Chatani, Y.; Yamori, T.; Tsuruo, T.; Oka, H.; Yoshida, O.; Shimada, Y.; Ari-i, S.; Wada, H.; Fujimoto, J.; et al. Constitutive activation of the 41-/43-kDa mitogen-activated protein kinase signaling pathway in human tumors. Oncogene 1999, 18, 813–822. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Romeo, Y.; Zhang, X.; Roux, P.P. Regulation and function of the RSK family of protein kinases. Biochem. J. 2012, 441, 553–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Anjum, R.; Blenis, J. The RSK family of kinases: Emerging roles in cellular signalling. Nat. Rev. Mol. Cell Biol. 2008, 9, 747–758. [Google Scholar] [CrossRef]
  6. Romeo, Y.; Roux, P.P. Paving the way for targeting RSK in cancer. Expert Opin. Ther. Targets. 2011, 15, 5–9. [Google Scholar] [CrossRef] [PubMed]
  7. Sulzmaier, F.J.; Ramos, J.W. RSK isoforms in cancer cell invasion and metastasis. Cancer Res. 2013, 73, 6099–6105. [Google Scholar] [CrossRef] [Green Version]
  8. Jones, S.W.; Erikson, E.; Blenis, J.; Maller, J.L.; Erikson, R.L. A Xenopus ribosomal protein S6 kinase has two apparent kinase domains that are each similar to distinct protein kinases. Proc. Natl. Acad. Sci. USA 1988, 85, 3377–3381. [Google Scholar] [CrossRef] [Green Version]
  9. Fisher, T.L.; Blenis, J. Evidence for two catalytically active kinase domains in pp90rsk. Mol. Cell Biol. 1996, 16, 1212–1219. [Google Scholar] [CrossRef] [Green Version]
  10. Bjorbaek, C.; Zhao, Y.; Moller, D.E. Divergent functional roles for p90rsk kinase domains. J. Biol. Chem. 1995, 270, 18848–18852. [Google Scholar] [CrossRef] [Green Version]
  11. Vik, T.A.; Ryder, J.W. Identification of serine 380 as the major site of autophosphorylation of Xenopus pp90rsk. Biochem. Biophys. Res. Commun. 1997, 235, 398–402. [Google Scholar] [CrossRef]
  12. Smith, J.A.; Poteet-Smith, C.E.; Malarkey, K.; Sturgill, T.W. Identification of an extracellular signal-regulated kinase (ERK) docking site in ribosomal S6 kinase, a sequence critical for activation by ERK in vivo. J. Biol. Chem. 1999, 274, 2893–2898. [Google Scholar] [CrossRef] [Green Version]
  13. Gavin, A.C.; Nebreda, A.R. A MAP kinase docking site is required for phosphorylation and activation of p90(rsk)/MAPKAP kinase-1. Curr. Biol. 1999, 9, 281–284. [Google Scholar] [CrossRef] [Green Version]
  14. Roux, P.P.; Richards, S.A.; Blenis, J. Phosphorylation of p90 ribosomal S6 kinase (RSK) regulates extracellular signal-regulated kinase docking and RSK activity. Mol. Cell Biol. 2003, 23, 4796–4804. [Google Scholar] [CrossRef] [Green Version]
  15. Thomas, G.M.; Rumbaugh, G.R.; Harrar, D.B.; Huganir, R.L. Ribosomal S6 kinase 2 interacts with and phosphorylates PDZ domain-containing proteins and regulates AMPA receptor transmission. Proc. Natl. Acad. Sci. USA 2005, 102, 15006–15011. [Google Scholar] [CrossRef] [Green Version]
  16. Dummler, B.A.; Hauge, C.; Silber, J.; Yntema, H.G.; Kruse, L.S.; Kofoed, B.; Hemmings, B.A.; Alessi, D.R.; Frodin, M. Functional characterization of human RSK4, a new 90-kDa ribosomal S6 kinase, reveals constitutive activation in most cell types. J. Biol. Chem. 2005, 280, 13304–13314. [Google Scholar] [CrossRef] [Green Version]
  17. Zeniou, M.; Ding, T.; Trivier, E.; Hanauer, A. Expression analysis of RSK gene family members: The RSK2 gene, mutated in Coffin-Lowry syndrome, is prominently expressed in brain structures essential for cognitive function and learning. Hum. Mol. Genet. 2002, 11, 2929–2940. [Google Scholar] [CrossRef]
  18. Carriere, A.; Ray, H.; Blenis, J.; Roux, P.P. The RSK factors of activating the Ras/MAPK signaling cascade. Front. Biosci. 2008, 13, 4258–4275. [Google Scholar] [CrossRef]
  19. Lara, R.; Seckl, M.J.; Pardo, O.E. The p90 RSK family members: Common functions and isoform specificity. Cancer Res. 2013, 73, 5301–5308. [Google Scholar] [CrossRef] [Green Version]
  20. Dalby, K.N.; Morrice, N.; Caudwell, F.B.; Avruch, J.; Cohen, P. Identification of regulatory phosphorylation sites in mitogen-activated protein kinase (MAPK)-activated protein kinase-1a/p90rsk that are inducible by MAPK. J. Biol. Chem. 1998, 273, 1496–1505. [Google Scholar] [CrossRef] [Green Version]
  21. Sutherland, C.; Alterio, J.; Campbell, D.G.; Le Bourdelles, B.; Mallet, J.; Haavik, J.; Cohen, P. Phosphorylation and activation of human tyrosine hydroxylase in vitro by mitogen-activated protein (MAP) kinase and MAP-kinase-activated kinases 1 and 2. Eur. J. Biochem. 1993, 217, 715–722. [Google Scholar] [CrossRef]
  22. Frodin, M.; Jensen, C.J.; Merienne, K.; Gammeltoft, S. A phosphoserine-regulated docking site in the protein kinase RSK2 that recruits and activates PDK1. EMBO J. 2000, 19, 2924–2934. [Google Scholar] [CrossRef]
  23. Jensen, C.J.; Buch, M.B.; Krag, T.O.; Hemmings, B.A.; Gammeltoft, S.; Frodin, M. 90-kDa ribosomal S6 kinase is phosphorylated and activated by 3-phosphoinositide-dependent protein kinase-1. J. Biol. Chem. 1999, 274, 27168–27176. [Google Scholar] [CrossRef] [Green Version]
  24. Richards, S.A.; Fu, J.; Romanelli, A.; Shimamura, A.; Blenis, J. Ribosomal S6 kinase 1 (RSK1) activation requires signals dependent on and independent of the MAP kinase ERK. Curr. Biol. 1999, 9, 810–820. [Google Scholar] [CrossRef] [Green Version]
  25. Frodin, M.; Antal, T.L.; Dummler, B.A.; Jensen, C.J.; Deak, M.; Gammeltoft, S.; Biondi, R.M. A phosphoserine/threonine-binding pocket in AGC kinases and PDK1 mediates activation by hydrophobic motif phosphorylation. EMBO J. 2002, 21, 5396–5407. [Google Scholar] [CrossRef] [Green Version]
  26. Doehn, U.; Gammeltoft, S.; Shen, S.H.; Jensen, C.J. p90 ribosomal S6 kinase 2 is associated with and dephosphorylated by protein phosphatase 2Cdelta. Biochem. J. 2004, 382 Pt 2, 425–431. [Google Scholar] [CrossRef] [Green Version]
  27. Zaru, R.; Ronkina, N.; Gaestel, M.; Arthur, J.S.; Watts, C. The MAPK-activated kinase Rsk controls an acute Toll-like receptor signaling response in dendritic cells and is activated through two distinct pathways. Nat. Immunol. 2007, 8, 1227–1235. [Google Scholar] [CrossRef] [PubMed]
  28. Kang, S.; Dong, S.; Gu, T.L.; Guo, A.; Cohen, M.S.; Lonial, S.; Khoury, H.J.; Fabbro, D.; Gilliland, D.G.; Bergsagel, P.L.; et al. FGFR3 activates RSK2 to mediate hematopoietic transformation through tyrosine phosphorylation of RSK2 and activation of the MEK/ERK pathway. Cancer Cell. 2007, 12, 201–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Kang, S.; Dong, S.; Guo, A.; Ruan, H.; Lonial, S.; Khoury, H.J.; Gu, T.L.; Chen, J. Epidermal growth factor stimulates RSK2 activation through activation of the MEK/ERK pathway and src-dependent tyrosine phosphorylation of RSK2 at Tyr-529. J. Biol. Chem. 2008, 283, 4652–4657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Kang, S.; Elf, S.; Dong, S.; Hitosugi, T.; Lythgoe, K.; Guo, A.; Ruan, H.; Lonial, S.; Khoury, H.J.; Williams, I.R.; et al. Fibroblast growth factor receptor 3 associates with and tyrosine phosphorylates p90 RSK2, leading to RSK2 activation that mediates hematopoietic transformation. Mol. Cell Biol. 2009, 29, 2105–2117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Lara, R.; Mauri, F.A.; Taylor, H.; Derua, R.; Shia, A.; Gray, C.; Nicols, A.; Shiner, R.J.; Schofield, E.; Bates, P.A.; et al. An siRNA screen identifies RSK1 as a key modulator of lung cancer metastasis. Oncogene 2011, 30, 3513–3521. [Google Scholar] [CrossRef]
  32. Carriere, A.; Cargnello, M.; Julien, L.A.; Gao, H.; Bonneil, E.; Thibault, P.; Roux, P.P. Oncogenic MAPK signaling stimulates mTORC1 activity by promoting RSK-mediated raptor phosphorylation. Curr. Biol. 2008, 18, 1269–1277. [Google Scholar] [CrossRef] [Green Version]
  33. Nandagopal, N.; Roux, P.P. Regulation of global and specific mRNA translation by the mTOR signaling pathway. Translation 2015, 3, e983402. [Google Scholar] [CrossRef] [Green Version]
  34. Rolfe, M.; McLeod, L.E.; Pratt, P.F.; Proud, C.G. Activation of protein synthesis in cardiomyocytes by the hypertrophic agent phenylephrine requires the activation of ERK and involves phosphorylation of tuberous sclerosis complex 2 (TSC2). Biochem. J. 2005, 388 Pt 3, 973–984. [Google Scholar] [CrossRef] [Green Version]
  35. Roux, P.P.; Ballif, B.A.; Anjum, R.; Gygi, S.P.; Blenis, J. Tumor-promoting phorbol esters and activated Ras inactivate the tuberous sclerosis tumor suppressor complex via p90 ribosomal S6 kinase. Proc. Natl. Acad. Sci. USA 2004, 101, 13489–13494. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, X.; Janmaat, M.; Beugnet, A.; Paulin, F.E.; Proud, C.G. Evidence that the dephosphorylation of Ser(535) in the epsilon-subunit of eukaryotic initiation factor (eIF) 2B is insufficient for the activation of eIF2B by insulin. Biochem. J. 2002, 367 Pt 2, 475–481. [Google Scholar] [CrossRef]
  37. Shahbazian, D.; Roux, P.P.; Mieulet, V.; Cohen, M.S.; Raught, B.; Taunton, J.; Hershey, J.W.; Blenis, J.; Pende, M.; Sonenberg, N. The mTOR/PI3K and MAPK pathways converge on eIF4B to control its phosphorylation and activity. EMBO J. 2006, 25, 2781–2791. [Google Scholar] [CrossRef]
  38. Shahbazian, D.; Parsyan, A.; Petroulakis, E.; Topisirovic, I.; Martineau, Y.; Gibbs, B.F.; Svitkin, Y.; Sonenberg, N. Control of cell survival and proliferation by mammalian eukaryotic initiation factor 4B. Mol. Cell Biol. 2010, 30, 1478–1485. [Google Scholar] [CrossRef] [Green Version]
  39. Galan, J.A.; Geraghty, K.M.; Lavoie, G.; Kanshin, E.; Tcherkezian, J.; Calabrese, V.; Jeschke, G.R.; Turk, B.E.; Ballif, B.A.; Blenis, J.; et al. Phosphoproteomic analysis identifies the tumor suppressor PDCD4 as a RSK substrate negatively regulated by 14-3-3. Proc. Natl. Acad. Sci. USA 2014, 111, E2918–E2927. [Google Scholar] [CrossRef] [Green Version]
  40. Cuesta, R.; Holz, M.K. RSK-mediated down-regulation of PDCD4 is required for proliferation, survival, and migration in a model of triple-negative breast cancer. Oncotarget 2016, 7, 27567–27583. [Google Scholar] [CrossRef]
  41. Roux, P.P.; Shahbazian, D.; Vu, H.; Holz, M.K.; Cohen, M.S.; Taunton, J.; Sonenberg, N.; Blenis, J. RAS/ERK signaling promotes site-specific ribosomal protein S6 phosphorylation via RSK and stimulates cap-dependent translation. J. Biol. Chem. 2007, 282, 14056–14064. [Google Scholar] [CrossRef] [Green Version]
  42. David, J.P.; Mehic, D.; Bakiri, L.; Schilling, A.F.; Mandic, V.; Priemel, M.; Idarraga, M.H.; Reschke, M.O.; Hoffmann, O.; Amling, M.; et al. Essential role of RSK2 in c-Fos-dependent osteosarcoma development. J. Clin. Investig. 2005, 115, 664–672. [Google Scholar] [CrossRef] [Green Version]
  43. Chen, R.H.; Juo, P.C.; Curran, T.; Blenis, J. Phosphorylation of c-Fos at the C-terminus enhances its transforming activity. Oncogene 1996, 12, 1493–1502. [Google Scholar]
  44. Murphy, L.O.; Smith, S.; Chen, R.H.; Fingar, D.C.; Blenis, J. Molecular interpretation of ERK signal duration by immediate early gene products. Nat. Cell Biol. 2002, 4, 556–564. [Google Scholar] [CrossRef]
  45. Fujita, N.; Sato, S.; Tsuruo, T. Phosphorylation of p27Kip1 at threonine 198 by p90 ribosomal protein S6 kinases promotes its binding to 14-3-3 and cytoplasmic localization. J. Biol. Chem. 2003, 278, 49254–49260. [Google Scholar] [CrossRef] [Green Version]
  46. Larrea, M.D.; Hong, F.; Wander, S.A.; da Silva, T.G.; Helfman, D.; Lannigan, D.; Smith, J.A.; Slingerland, J.M. RSK1 drives p27Kip1 phosphorylation at T198 to promote RhoA inhibition and increase cell motility. Proc. Natl. Acad. Sci. USA 2009, 106, 9268–9273. [Google Scholar] [CrossRef] [Green Version]
  47. Diehl, J.A.; Cheng, M.; Roussel, M.F.; Sherr, C.J. Glycogen synthase kinase-3beta regulates cyclin D1 proteolysis and subcellular localization. Genes Dev. 1998, 12, 3499–3511. [Google Scholar] [CrossRef] [Green Version]
  48. Sears, R.; Nuckolls, F.; Haura, E.; Taya, Y.; Tamai, K.; Nevins, J.R. Multiple Ras-dependent phosphorylation pathways regulate Myc protein stability. Genes Dev. 2000, 14, 2501–2514. [Google Scholar] [CrossRef] [Green Version]
  49. Wu, C.F.; Liu, S.; Lee, Y.C.; Wang, R.; Sun, S.; Yin, F.; Bornmann, W.G.; Yu-Lee, L.Y.; Gallick, G.E.; Zhang, W.; et al. RSK promotes G2/M transition through activating phosphorylation of Cdc25A and Cdc25B. Oncogene 2014, 33, 2385–2394. [Google Scholar] [CrossRef] [Green Version]
  50. Zhu, J.; Blenis, J.; Yuan, J. Activation of PI3K/Akt and MAPK pathways regulates Myc-mediated transcription by phosphorylating and promoting the degradation of Mad1. Proc. Natl. Acad. Sci. USA 2008, 105, 6584–6589. [Google Scholar] [CrossRef] [Green Version]
  51. Bruning, J.C.; Gillette, J.A.; Zhao, Y.; Bjorbaeck, C.; Kotzka, J.; Knebel, B.; Avci, H.; Hanstein, B.; Lingohr, P.; Moller, D.E.; et al. Ribosomal subunit kinase-2 is required for growth factor-stimulated transcription of the c-Fos gene. Proc. Natl. Acad. Sci. USA 2000, 97, 2462–2467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Shimamura, A.; Ballif, B.A.; Richards, S.A.; Blenis, J. Rsk1 mediates a MEK-MAP kinase cell survival signal. Curr. Biol. 2000, 10, 127–135. [Google Scholar] [CrossRef] [Green Version]
  53. Bialik, S.; Kimchi, A. DAP-kinase as a target for drug design in cancer and diseases associated with accelerated cell death. Semin. Cancer Biol. 2004, 14, 283–294. [Google Scholar] [CrossRef] [PubMed]
  54. Anjum, R.; Roux, P.P.; Ballif, B.A.; Gygi, S.P.; Blenis, J. The tumor suppressor DAP kinase is a target of RSK-mediated survival signaling. Curr. Biol. 2005, 15, 1762–1767. [Google Scholar] [CrossRef] [Green Version]
  55. Buck, M.; Poli, V.; Hunter, T.; Chojkier, M. C/EBPbeta phosphorylation by RSK creates a functional XEXD caspase inhibitory box critical for cell survival. Mol. Cell. 2001, 8, 807–816. [Google Scholar] [CrossRef]
  56. Xing, J.; Ginty, D.D.; Greenberg, M.E. Coupling of the RAS-MAPK pathway to gene activation by RSK2, a growth factor-regulated CREB kinase. Science 1996, 273, 959–963. [Google Scholar] [CrossRef]
  57. Doehn, U.; Hauge, C.; Frank, S.R.; Jensen, C.J.; Duda, K.; Nielsen, J.V.; Cohen, M.S.; Johansen, J.V.; Winther, B.R.; Lund, L.R.; et al. RSK is a principal effector of the RAS-ERK pathway for eliciting a coordinate promotile/invasive gene program and phenotype in epithelial cells. Mol. Cell. 2009, 35, 511–522. [Google Scholar] [CrossRef] [Green Version]
  58. Gawecka, J.E.; Young-Robbins, S.S.; Sulzmaier, F.J.; Caliva, M.J.; Heikkila, M.M.; Matter, M.L.; Ramos, J.W. RSK2 protein suppresses integrin activation and fibronectin matrix assembly and promotes cell migration. J. Biol. Chem. 2012, 287, 43424–43437. [Google Scholar] [CrossRef] [Green Version]
  59. Woo, M.S.; Ohta, Y.; Rabinovitz, I.; Stossel, T.P.; Blenis, J. Ribosomal S6 kinase (RSK) regulates phosphorylation of filamin A on an important regulatory site. Mol. Cell Biol. 2004, 24, 3025–3035. [Google Scholar] [CrossRef] [Green Version]
  60. Tanimura, S.; Hashizume, J.; Kurosaki, Y.; Sei, K.; Gotoh, A.; Ohtake, R.; Kawano, M.; Watanabe, K.; Kohno, M. SH3P2 is a negative regulator of cell motility whose function is inhibited by ribosomal S6 kinase-mediated phosphorylation. Genes Cells 2011, 16, 514–526. [Google Scholar] [CrossRef]
  61. Chen, C.; Zhang, L.; Huang, N.J.; Huang, B.; Kornbluth, S. Suppression of DNA-damage checkpoint signaling by Rsk-mediated phosphorylation of Mre11. Proc. Natl. Acad. Sci. USA 2013, 110, 20605–20610. [Google Scholar] [CrossRef] [Green Version]
  62. Ray-David, H.; Romeo, Y.; Lavoie, G.; Deleris, P.; Tcherkezian, J.; Galan, J.A.; Roux, P.P. RSK promotes G2 DNA damage checkpoint silencing and participates in melanoma chemoresistance. Oncogene 2013, 32, 4480–4489. [Google Scholar] [CrossRef] [Green Version]
  63. Abdulrahman, N.; Jaballah, M.; Poomakkoth, N.; Riaz, S.; Abdelaziz, S.; Issa, A.; Mraiche, F. Inhibition of p90 ribosomal S6 kinase attenuates cell migration and proliferation of the human lung adenocarcinoma through phospho-GSK-3beta and osteopontin. Mol. Cell Biochem. 2016, 418, 21–29. [Google Scholar] [CrossRef]
  64. Poomakkoth, N.; Issa, A.; Abdulrahman, N.; Abdelaziz, S.G.; Mraiche, F. p90 ribosomal S6 kinase: A potential therapeutic target in lung cancer. J. Transl. Med. 2016, 14, 14. [Google Scholar] [CrossRef] [Green Version]
  65. Kang, S.; Elf, S.; Lythgoe, K.; Hitosugi, T.; Taunton, J.; Zhou, W.; Xiong, L.; Wang, D.; Muller, S.; Fan, S.; et al. p90 ribosomal S6 kinase 2 promotes invasion and metastasis of human head and neck squamous cell carcinoma cells. J. Clin. Investig. 2010, 120, 1165–1177. [Google Scholar] [CrossRef] [Green Version]
  66. Smith, J.A.; Poteet-Smith, C.E.; Xu, Y.; Errington, T.M.; Hecht, S.M.; Lannigan, D.A. Identification of the first specific inhibitor of p90 ribosomal S6 kinase (RSK) reveals an unexpected role for RSK in cancer cell proliferation. Cancer Res. 2005, 65, 1027–1034. [Google Scholar]
  67. Clark, D.E.; Errington, T.M.; Smith, J.A.; Frierson, H.F.; Weber, M.J., Jr.; Lannigan, D.A. The serine/threonine protein kinase, p90 ribosomal S6 kinase, is an important regulator of prostate cancer cell proliferation. Cancer Res. 2005, 65, 3108–3116. [Google Scholar] [CrossRef] [Green Version]
  68. Elf, S.; Blevins, D.; Jin, L.; Chung, T.W.; Williams, I.R.; Lee, B.H.; Lin, J.X.; Leonard, W.J.; Taunton, J.; Khoury, H.J.; et al. p90RSK2 is essential for FLT3-ITD- but dispensable for BCR-ABL-induced myeloid leukemia. Blood. 2011, 117, 6885–6894. [Google Scholar] [CrossRef] [Green Version]
  69. Romeo, Y.; Moreau, J.; Zindy, P.J.; Saba-El-Leil, M.; Lavoie, G.; Dandachi, F.; Baptissart, M.; Borden, K.L.B.; Meloche, S.; Roux, P.P. RSK regulates activated BRAF signalling to mTORC1 and promotes melanoma growth. Oncogene 2013, 32, 2917–2926. [Google Scholar] [CrossRef] [Green Version]
  70. Sulzmaier, F.J.; Young-Robbins, S.; Jiang, P.; Geerts, D.; Prechtl, A.M.; Matter, M.L.; Kesari, S.; Ramos, J.W. RSK2 activity mediates glioblastoma invasiveness and is a potential target for new therapeutics. Oncotarget 2016, 7, 79869–79884. [Google Scholar] [CrossRef] [Green Version]
  71. Bignone, P.A.; Lee, K.Y.; Liu, Y.; Emilion, G.; Finch, J.; Soosay, A.E.; Charnock, F.M.; Beck, S.; Dunham, I.; Mungall, A.J.; et al. RPS6KA2, a putative tumour suppressor gene at 6q27 in sporadic epithelial ovarian cancer. Oncogene 2007, 26, 683–700. [Google Scholar] [CrossRef] [Green Version]
  72. Cai, J.; Ma, H.; Huang, F.; Zhu, D.; Zhao, L.; Yang, Y.; Bi, J.; Zhang, T. Low expression of RSK4 predicts poor prognosis in patients with colorectal cancer. Int. J. Clin. Exp. Pathol. 2014, 7, 4959–4970. [Google Scholar]
  73. Rafiee, M.; Keramati, M.R.; Ayatollahi, H.; Sadeghian, M.H.; Barzegar, M.; Asgharzadeh, A.; Alinejad, M. Down-Regulation of Ribosomal S6 kinase RPS6KA6 in Acute Myeloid Leukemia Patients. Cell J. 2016, 18, 159–164. [Google Scholar]
  74. Li, Q.; Jiang, Y.; Wei, W.; Ji, Y.; Gao, H.; Liu, J. Frequent epigenetic inactivation of RSK4 by promoter methylation in cancerous and non-cancerous tissues of breast cancer. Med. Oncol. 2014, 31, 793. [Google Scholar] [CrossRef]
  75. Thakur, A.; Sun, Y.; Bollig, A.; Wu, J.; Biliran, H.; Banerjee, S.; Sarkar, F.H.; Liao, D.J. Anti-invasive and antimetastatic activities of ribosomal protein S6 kinase 4 in breast cancer cells. Clin. Cancer Res. 2008, 14, 4427–4436. [Google Scholar] [CrossRef] [Green Version]
  76. Berns, K.; Hijmans, E.M.; Mullenders, J.; Brummelkamp, T.R.; Velds, A.; Heimerikx, M.; Kerkhoven, R.M.; Madiredjo, M.; Nijkamp, W.; Weigelt, B.; et al. A large-scale RNAi screen in human cells identifies new components of the p53 pathway. Nature 2004, 428, 431–437. [Google Scholar] [CrossRef] [Green Version]
  77. Lopez-Vicente, L.; Armengol, G.; Pons, B.; Coch, L.; Argelaguet, E.; Lleonart, M.; Hernandez-Losa, J.; de Torres, I.; Ramon y Cajal, S. Regulation of replicative and stress-induced senescence by RSK4, which is down-regulated in human tumors. Clin. Cancer Res. 2009, 15, 4546–4553. [Google Scholar] [CrossRef] [Green Version]
  78. Serra, V.; Eichhorn, P.J.; Garcia-Garcia, C.; Ibrahim, Y.H.; Prudkin, L.; Sanchez, G.; Rodriguez, O.; Anton, P.; Parra, J.L.; Marlow, S.; et al. RSK3/4 mediate resistance to PI3K pathway inhibitors in breast cancer. J. Clin. Investig. 2013, 123, 2551–2563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Kavanagh, S.; Murphy, T.; Law, A.; Yehudai, D.; Ho, J.M.; Chan, S.; Schimmer, A.D. Emerging therapies for acute myeloid leukemia: Translating biology into the clinic. JCI Insight 2017, 2, e95679. [Google Scholar] [CrossRef] [Green Version]
  80. Saygin, C.; Carraway, H.E. Emerging therapies for acute myeloid leukemia. J. Hematol. Oncol. 2017, 10, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Dombret, H.; Gardin, C. An update of current treatments for adult acute myeloid leukemia. Blood 2016, 127, 53–61. [Google Scholar] [CrossRef] [PubMed]
  82. Ricciardi, M.R.; McQueen, T.; Chism, D.; Milella, M.; Estey, E.; Kaldjian, E.; Sebolt-Leopold, J.; Konopleva, M.; Andreeff, M. Quantitative single cell determination of ERK phosphorylation and regulation in relapsed and refractory primary acute myeloid leukemia. Leukemia 2005, 19, 1543–1549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kornblau, S.M.; Womble, M.; Qiu, Y.H.; Jackson, C.E.; Chen, W.; Konopleva, M.; Estey, E.H.; Andreeff, M. Simultaneous activation of multiple signal transduction pathways confers poor prognosis in acute myelogenous leukemia. Blood 2006, 108, 2358–2365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Shankar, D.B.; Cheng, J.C.; Kinjo, K.; Federman, N.; Moore, T.B.; Gill, A.; Rao, N.P.; Landaw, E.M.; Sakamoto, K.M. The role of CREB as a proto-oncogene in hematopoiesis and in acute myeloid leukemia. Cancer Cell. 2005, 7, 351–362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Cho, E.C.; Mitton, B.; Sakamoto, K.M. CREB and leukemogenesis. Crit. Rev. Oncog. 2011, 16, 37–46. [Google Scholar] [CrossRef] [Green Version]
  86. Kwon, E.M.; Raines, M.A.; Blenis, J.; Sakamoto, K.M. Granulocyte-macrophage colony-stimulating factor stimulation results in phosphorylation of cAMP response element-binding protein through activation of pp90RSK. Blood 2000, 95, 2552–2558. [Google Scholar] [CrossRef]
  87. Chae, H.D.; Dutta, R.; Tiu, B.; Hoff, F.W.; Accordi, B.; Serafin, V.; Youn, M.; Huang, M.; Sumarsono, N.; Davis, K.L.; et al. RSK inhibitor BI-D1870 inhibits acute myeloid leukemia cell proliferation by targeting mitotic exit. Oncotarget 2020, 11, 2387–2403. [Google Scholar] [CrossRef]
  88. Hospital, M.A.; Green, A.S.; Maciel, T.T.; Moura, I.C.; Leung, A.Y.; Bouscary, D.; Tamburini, J. FLT3 inhibitors: Clinical potential in acute myeloid leukemia. Onco Targets Ther. 2017, 10, 607–615. [Google Scholar] [CrossRef] [Green Version]
  89. Meshinchi, S.; Appelbaum, F.R. Structural and functional alterations of FLT3 in acute myeloid leukemia. Clin. Cancer Res. 2009, 15, 4263–4269. [Google Scholar] [CrossRef] [Green Version]
  90. Levis, M.; Small, D. FLT3: ITDoes matter in leukemia. Leukemia 2003, 17, 1738–1752. [Google Scholar] [CrossRef] [Green Version]
  91. Yang, X.; Liu, L.; Sternberg, D.; Tang, L.; Galinsky, I.; DeAngelo, D.; Stone, R. The FLT3 internal tandem duplication mutation prevents apoptosis in interleukin-3-deprived BaF3 cells due to protein kinase A and ribosomal S6 kinase 1-mediated BAD phosphorylation at serine 112. Cancer Res. 2005, 65, 7338–7347. [Google Scholar] [CrossRef] [Green Version]
  92. Hospital, M.A.; Jacquel, A.; Mazed, F.; Saland, E.; Larrue, C.; Mondesir, J.; Birsen, R.; Green, A.S.; Lambert, M.; Sujobert, P.; et al. RSK2 is a new Pim2 target with pro-survival functions in FLT3-ITD-positive acute myeloid leukemia. Leukemia 2018, 32, 597–605. [Google Scholar] [CrossRef]
  93. Green, A.S.; Maciel, T.T.; Hospital, M.A.; Yin, C.; Mazed, F.; Townsend, E.C.; Pilorge, S.; Lambert, M.; Paubelle, E.; Jacquel, A.; et al. Pim kinases modulate resistance to FLT3 tyrosine kinase inhibitors in FLT3-ITD acute myeloid leukemia. Sci. Adv. 2015, 1, e1500221. [Google Scholar] [CrossRef] [Green Version]
  94. Watanabe, D.; Nogami, A.; Okada, K.; Akiyama, H.; Umezawa, Y.; Miura, O. FLT3-ITD Activates RSK1 to Enhance Proliferation and Survival of AML Cells by Activating mTORC1 and eIF4B Cooperatively with PIM or PI3K and by Inhibiting Bad and BIM. Cancers 2019, 11, 1827. [Google Scholar] [CrossRef] [Green Version]
  95. Parmar, S.; Rundhaugen, L.M.; Boehlke, L.; Riley, M.; Nabhan, C.; Raji, A.; Frater, J.L.; Tallman, M.S. Phase II trial of arsenic trioxide in relapsed and refractory acute myeloid leukemia, secondary leukemia and/or newly diagnosed patients at least 65 years old. Leuk. Res. 2004, 28, 909–919. [Google Scholar] [CrossRef]
  96. Galvin, J.P.; Altman, J.K.; Szilard, A.; Goussetis, D.J.; Vakana, E.; Sassano, A.; Platanias, L.C. Regulation of the kinase RSK1 by arsenic trioxide and generation of antileukemic responses. Cancer Biol. Ther. 2013, 14, 411–416. [Google Scholar] [CrossRef]
  97. Beauchamp, E.M.; Kosciuczuk, E.M.; Serrano, R.; Nanavati, D.; Swindell, E.P.; Viollet, B.; O’Halloran, T.V.; Altman, J.K.; Platanias, L.C. Direct binding of arsenic trioxide to AMPK and generation of inhibitory effects on acute myeloid leukemia precursors. Mol. Cancer Ther. 2015, 14, 202–212. [Google Scholar] [CrossRef] [Green Version]
  98. Taipale, J.; Beachy, P.A. The Hedgehog and Wnt signalling pathways in cancer. Nature 2001, 411, 349–354. [Google Scholar] [CrossRef]
  99. Pambid, M.R.; Berns, R.; Adomat, H.H.; Hu, K.; Triscott, J.; Maurer, N.; Zisman, N.; Ramaswamy, V.; Hawkins, C.E.; Taylor, M.D.; et al. Overcoming resistance to Sonic Hedgehog inhibition by targeting p90 ribosomal S6 kinase in pediatric medulloblastoma. Pediatr. Blood Cancer. 2014, 61, 107–115. [Google Scholar] [CrossRef] [Green Version]
  100. Lin, T.L.; Levy, M.Y. Acute myeloid leukemia: Focus on novel therapeutic strategies. Clin. Med. Insights Oncol. 2012, 6, 205–217. [Google Scholar] [CrossRef] [Green Version]
  101. Zhao, C.; Chen, A.; Jamieson, C.H.; Fereshteh, M.; Abrahamsson, A.; Blum, J.; Kwon, H.Y.; Kim, J.; Chute, J.P.; Rizzieri, D.; et al. Hedgehog signalling is essential for maintenance of cancer stem cells in myeloid leukaemia. Nature 2009, 458, 776–779. [Google Scholar] [CrossRef] [Green Version]
  102. Peacock, C.D.; Wang, Q.; Gesell, G.S.; Corcoran-Schwartz, I.M.; Jones, E.; Kin, J.; Devereux, W.L.; Rhodes, J.T.; Huff, C.A.; Beachy, P.A.; et al. Hedgehog signaling maintains a tumor stem cell compartment in multiple myeloma. Proc. Natl. Acad. Sci. USA 2007, 104, 4048–4053. [Google Scholar] [CrossRef] [Green Version]
  103. Nguyen, T.L. Targeting RSK: An overview of small molecule inhibitors. Anticancer Agents Med. Chem. 2008, 8, 710–716. [Google Scholar] [CrossRef]
  104. Smith, J.A.; Maloney, D.J.; Hecht, S.M.; Lannigan, D.A. Structural basis for the activity of the RSK-specific inhibitor, SL0101. Bioorg. Med. Chem. 2007, 15, 5018–5034. [Google Scholar] [CrossRef]
  105. Hilinski, M.K.; Mrozowski, R.M.; Clark, D.E.; Lannigan, D.A. Analogs of the RSK inhibitor SL0101: Optimization of in vitro biological stability. Bioorg. Med. Chem. Lett. 2012, 22, 3244–3247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Sapkota, G.P.; Cummings, L.; Newell, F.S.; Armstrong, C.; Bain, J.; Frodin, M.; Grauert, M.; Hoffmann, M.; Schnapp, G.; Steegmaier, M.; et al. BI-D1870 is a specific inhibitor of the p90 RSK (ribosomal S6 kinase) isoforms in vitro and in vivo. Biochem. J. 2007, 401, 29–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Bain, J.; Plater, L.; Elliott, M.; Shpiro, N.; Hastie, C.J.; McLauchlan, H.; Klevernic, I.; Arthur, J.S.; Alessi, D.R.; Cohen, P. The selectivity of protein kinase inhibitors: A further update. Biochem. J. 2007, 408, 297–315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Casalvieri, K.A.; Matheson, C.J.; Backos, D.S.; Reigan, P. Substituted pteridinones as p90 ribosomal S6 protein kinase (RSK) inhibitors: A structure-activity study. Bioorg. Med. Chem. 2020, 28, 115303. [Google Scholar] [CrossRef]
  109. Aronchik, I.; Appleton, B.A.; Basham, S.E.; Crawford, K.; Del Rosario, M.; Doyle, L.V.; Estacio, W.F.; Lan, J.; Lindvall, M.K.; Luu, C.A.; et al. Novel potent and selective inhibitors of p90 ribosomal S6 kinase reveal the heterogeneity of RSK function in MAPK-driven cancers. Mol. Cancer Res. 2014, 12, 803–812. [Google Scholar] [CrossRef] [Green Version]
  110. Jain, R.; Mathur, M.; Lan, J.; Costales, A.; Atallah, G.; Ramurthy, S.; Subramanian, S.; Setti, L.; Feucht, P.; Warne, B.; et al. Discovery of Potent and Selective RSK Inhibitors as Biological Probes. J. Med. Chem. 2015, 58, 6766–6783. [Google Scholar] [CrossRef]
  111. Kirrane, T.M.; Boyer, S.J.; Burke, J.; Guo, X.; Snow, R.J.; Soleymanzadeh, L.; Swinamer, A.; Zhang, Y.; Madwed, J.B.; Kashem, M.; et al. Indole RSK inhibitors. Part 2: Optimization of cell potency and kinase selectivity. Bioorg. Med. Chem. Lett. 2012, 22, 738–742. [Google Scholar] [CrossRef]
  112. Boyer, S.J.; Burke, J.; Guo, X.; Kirrane, T.M.; Snow, R.J.; Zhang, Y.; Sarko, C.; Soleymanzadeh, L.; Swinamer, A.; Westbrook, J.; et al. Indole RSK inhibitors. Part 1: Discovery and initial SAR. Bioorg. Med. Chem. Lett. 2012, 22, 733–737. [Google Scholar] [CrossRef]
  113. Fryer, R.M.; Muthukumarana, A.; Chen, R.R.; Smith, J.D.; Mazurek, S.N.; Harrington, K.E.; Dinallo, R.M.; Burke, J.; DiCapua, F.M.; Guo, X.; et al. Mitigation of off-target adrenergic binding and effects on cardiovascular function in the discovery of novel ribosomal S6 kinase 2 inhibitors. J. Pharmacol. Exp. Ther. 2012, 340, 492–500. [Google Scholar] [CrossRef] [Green Version]
  114. Fomina-Yadlin, D.; Kubicek, S.; Walpita, D.; Dancik, V.; Hecksher-Sorensen, J.; Bittker, J.A.; Sharifnia, T.; Shamji, A.; Clemons, P.A.; Wagner, B.K.; et al. Small-molecule inducers of insulin expression in pancreatic alpha-cells. Proc. Natl. Acad. Sci. USA 2010, 107, 15099–15104. [Google Scholar] [CrossRef] [Green Version]
  115. Park, Y.S.; Cho, N.J. EGFR and PKC are involved in the activation of ERK1/2 and p90 RSK and the subsequent proliferation of SNU-407 colon cancer cells by muscarinic acetylcholine receptors. Mol. Cell Biochem. 2012, 370, 191–198. [Google Scholar] [CrossRef]
  116. Cohen, M.S.; Zhang, C.; Shokat, K.M.; Taunton, J. Structural bioinformatics-based design of selective, irreversible kinase inhibitors. Science 2005, 308, 1318–1321. [Google Scholar] [CrossRef] [Green Version]
  117. Cohen, M.S.; Hadjivassiliou, H.; Taunton, J. A clickable inhibitor reveals context-dependent autoactivation of p90 RSK. Nat. Chem Biol. 2007, 3, 156–160. [Google Scholar] [CrossRef] [Green Version]
  118. Dunn, S.E.; Jayanthan, A.; Huynh, M.-M.; Flahive, E.; Pambid, M.R.; Dorr, A.; Los, G. Abstract P3-10-10: PMD-026, a first-in-class oral p90 ribosomal S6 kinase (RSK) inhibitor for triple negative breast cancer (TNBC). Cancer Res 2020, 80 (Suppl. S4), P3-10-10–P13-10-10. [Google Scholar]
Figure 1. A schematic model for RSK activation. When RTK is stimulated by growth factor, it activates the docking proteins GRB2 and SOS. SOS triggers Ras to exchange guanosine diphosphate (GDP) to guanosine triphosphate (GTP) and then to become activated. Ras activates Raf kinases, which phosphorylate MEK1/2, ERK1/2, and RSK1-4. Then, RSKs phosphorylate various downstream substrates to mediate diverse cellular processes.
Figure 1. A schematic model for RSK activation. When RTK is stimulated by growth factor, it activates the docking proteins GRB2 and SOS. SOS triggers Ras to exchange guanosine diphosphate (GDP) to guanosine triphosphate (GTP) and then to become activated. Ras activates Raf kinases, which phosphorylate MEK1/2, ERK1/2, and RSK1-4. Then, RSKs phosphorylate various downstream substrates to mediate diverse cellular processes.
Biomedicines 09 00726 g001
Figure 2. The domain structure of four RSK isoforms in human. RSKs exhibit two functional domains NTKD and CTKD, which are connected by a linker region. RSKs have six conserved phosphorylation sites. C-terminal tail contains an ERK1/2-docking domain called KIM motif and PDZ-binding motif.
Figure 2. The domain structure of four RSK isoforms in human. RSKs exhibit two functional domains NTKD and CTKD, which are connected by a linker region. RSKs have six conserved phosphorylation sites. C-terminal tail contains an ERK1/2-docking domain called KIM motif and PDZ-binding motif.
Biomedicines 09 00726 g002
Figure 3. A schematic model of RSK1 activating process. Extracellular signals activate ERK1/2, which phosphorylates Thr573 in CTKD. The activated CTKD autophosphorylates Ser380 at the hydrophobic motif. In addition, ERK1/2 might carry out the phosphorylation for Thr359 and Ser363. A constitutively active Ser/Thr kinase PDK1 binds at the phosphorylated Ser380 and phosphorylates Ser211 in the NTKD, which leads to the full activation of RSKs and subsequent phosphorylation of the various substrates. This is followed by the phosphorylation of Ser749 by the NTKD, decreasing the ERK1/2 affinity for RSKs in a negative feedback loop.
Figure 3. A schematic model of RSK1 activating process. Extracellular signals activate ERK1/2, which phosphorylates Thr573 in CTKD. The activated CTKD autophosphorylates Ser380 at the hydrophobic motif. In addition, ERK1/2 might carry out the phosphorylation for Thr359 and Ser363. A constitutively active Ser/Thr kinase PDK1 binds at the phosphorylated Ser380 and phosphorylates Ser211 in the NTKD, which leads to the full activation of RSKs and subsequent phosphorylation of the various substrates. This is followed by the phosphorylation of Ser749 by the NTKD, decreasing the ERK1/2 affinity for RSKs in a negative feedback loop.
Biomedicines 09 00726 g003
Figure 4. Molecular targets and cellular function of RSK isoforms. RSKs control diverse cellular processes by regulating extensive downstream targets.
Figure 4. Molecular targets and cellular function of RSK isoforms. RSKs control diverse cellular processes by regulating extensive downstream targets.
Biomedicines 09 00726 g004
Figure 5. Functional mechanisms of RSK isoforms in AML. RSKs control survival and apoptosis of leukemic cells through aberrantly activated upstream signaling.
Figure 5. Functional mechanisms of RSK isoforms in AML. RSKs control survival and apoptosis of leukemic cells through aberrantly activated upstream signaling.
Biomedicines 09 00726 g005
Table 1. Summary of pharmacological inhibitors of RSK isoforms.
Table 1. Summary of pharmacological inhibitors of RSK isoforms.
InhibitorsStructureTargetIC50EC50
SL0101 Biomedicines 09 00726 i001CKTD89 nM in RSK250 μM
BI-D1870 Biomedicines 09 00726 i002CKTD31 nM in RSK1
24 nM in RSK2
18 nM in RSK3
15 nM in RSK4
1 μM
LJH685 Biomedicines 09 00726 i003CKTD6 nM in RSK1
5 nM in RSK2
4 nM in RSK3
730–790 nM
LJI308 Biomedicines 09 00726 i004CKTD6 nM in RSK1
4 nM in RSK2
13 nM in RSK3
200–300 nM
BIX02565 Biomedicines 09 00726 i005CKTD1 nM in RSK2N.A.
BRD7389 Biomedicines 09 00726 i006CKTD1.5 μM in RSK1
2.4 μM in RSK2
1.2 μM in RSK3
N.A.
FMK Biomedicines 09 00726 i007NKTD15 nM in RSK2200 nM
PM-026 0.7–2 nM in RSK2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Youn, M.; Gomez, J.O.; Mark, K.; Sakamoto, K.M. RSK Isoforms in Acute Myeloid Leukemia. Biomedicines 2021, 9, 726. https://doi.org/10.3390/biomedicines9070726

AMA Style

Youn M, Gomez JO, Mark K, Sakamoto KM. RSK Isoforms in Acute Myeloid Leukemia. Biomedicines. 2021; 9(7):726. https://doi.org/10.3390/biomedicines9070726

Chicago/Turabian Style

Youn, Minyoung, Jesus Omar Gomez, Kailen Mark, and Kathleen M. Sakamoto. 2021. "RSK Isoforms in Acute Myeloid Leukemia" Biomedicines 9, no. 7: 726. https://doi.org/10.3390/biomedicines9070726

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