Abstract
Over the last decade, the CMGC kinase DYRK2 has been reported as a tumor suppressor across various cancers triggering major antitumor and proapoptotic signals in breast, colon, liver, ovary, brain, and lung cancers, with lower DYRK2 expression correlated with poorer prognosis in patients. Contrary to this, various medicinal chemistry studies reported robust antiproliferative properties of DYRK2 inhibitors, whereas unbiased ‘omics’ and genome-wide association study-based studies identified DYRK2 as a highly overexpressed kinase in various patient tumor samples. A major paradigm shift occurred in the last 4 years when DYRK2 was found to regulate proteostasis in cancer via a two-pronged mechanism. DYRK2 phosphorylated and activated the 26S proteasome to enhance degradation of misfolded/tumor-suppressor proteins while also promoting the nuclear stability and transcriptional activity of its substrate, heat-shock factor 1 triggering protein folding. Together, DYRK2 regulates proteostasis and promotes protumorigenic survival for specific cancers. Indeed, potent and selective small-molecule inhibitors of DYRK2 exhibit in vitro and in vivo anti-tumor activity in triple-negative breast cancer and myeloma models. However, with conflicting and contradictory reports across different cancers, the overarching role of DYRK2 remains enigmatic. Specific cancer (sub)types coupled to spatiotemporal interactions with substrates could decide the procancer or anticancer role of DYRK2. The current review aims to provide a balanced and critical appreciation of the literature to date, highlighting top substrates such as p53, c-Myc, c-Jun, heat-shock factor 1, proteasome, or NOTCH1, to discuss DYRK2 inhibitors available to the scientific community and to shed light on this duality of protumorigenic and antitumorigenic roles of DYRK2.
Keywords: protein kinase, kinase inhibitor, proteasome, stress, phosphorylation, E3 ligase, proteostasis
Abbreviations: CMGC, Cyclin-dependent kinases, Mitogen-activated protein kinases, Glycogen synthase kinases, and CDC-like kinases; CML, chronic myeloid leukemia; DYRK, Dual-specificity tYrosine phosphorylation–Regulated Kinase; EMT, epithelial–mesenchymal transition; GEMMs, genetically engineered mouse models; HIPK2, homeodomain-interacting protein kinase; HSF1, heat-shock factor 1; IFN, interferon; MM, multiple myeloma; NAPA, N-terminal autophosphorylation accessory; NOTCH1, neurogenic locus notch homolog protein 1; NSCLC, non–small-cell lung cancer; PEST, Pro-Glu-Ser-Thr; Ser62, serine62; Ser727, serine727; STAT3, signal transducer and activator of transcription 3; TBK1, TANK-binding kinase 1; TCGA, The Cancer Genome Atlas; Thr58, Threonine58; TNBC, triple-negative breast cancer
Protein kinase DYRK2 is a member of the Dual-specificity tYrosine phosphorylation–Regulated Kinase (DYRK) family, which in turn belongs to the Cyclin-dependent kinases, Mitogen-activated protein kinases, Glycogen synthase kinases, and CDC-like kinases (CMGC) superfamily within the kinase complement of the human genome (1). The DYRK family consists of 5 members divided into two classes: Class I is comprised of DYRK1A and DYRK1B, whereas class II is comprised of DYRK2, DYRK3, and DYRK4 (Fig. 1A). DYRK2 is a class II DYRK that exhibits various structural features such as the NAPA or N-terminal autophosphorylation accessory domains (yellow/orange), DYRK-homology domain (green), activation loop segment (purple), nuclear localization sequence (red), the CMGC family–specific insert domain (gray) (Fig. 1B) most of which are conserved across the DYRK family (2). In DYRK2, specific loss-of-function mutations have been reported in cancer (Fig. 1B), which affect either the activity of the kinase or impede its ability to form functional complexes with interactors (3). In fact, phosphoproteomics studies show that these cancer mutations significantly alter substrate specificity of DYRK2 in cells (3). Class I DYRKs exhibit two distinct nuclear localization sequences and a stretch of polyserine and polyproline (PEST, Pro-Glu-Ser-Thr) domain with no distinct NAPA domains as in class II paralogues (Fig. 1C). Despite subtle structural differences between class I and II members, all DYRK isoforms exhibit a highly conserved autophosphorylation-mediated activation mechanism (4). During translation, hydroxylation of a highly conserved proline residue (proline441 for hDYRK2) on the inert/nascent kinase domain of DYRKs triggers a tyrosine autophosphorylation event within the activation loop (tyrosine382 for hDYRK2), which leads to conversion of the inactive to the active conformation of the kinase (5). In the fully active form, the DYRK transition from a tyrosine-phosphorylating kinase to a serine-/threonine-directed kinase, thus acquiring the label ‘dual specificity’ (2, 4). The NAPA and DYRK-homology domain domains are thought to promote the structural integrity of the nascent kinase enough to execute the indispensable autophosphorylation event (2). The CMGC-specific insert is conserved across the CMGC kinase superfamily and is proposed to play important roles in stabilization of the tertiary structure of the kinase and promoting complex formation with interactors/substrates (6).
Like most CMGC kinases, DYRKs have an amino acid motif of preference on their substrates. DYRKs prefer an arginine (R) at the −3 position of the phosphoserine/threonine residue along with a strong preference for a proline (P) at +1: Rxx(pS/T)P motif (7). Being a preferred motif across all members, redundancies have been observed wherein multiple CMGC kinases phosphorylate the same site on the substrate (reviewed in Boni et al. [8]). Although both the −3 R and +1 P are strongly preferred, some DYRK substrates lack the +1 P such as histone H3 for DYRK1A (9), 26S proteasome regulatory subunit 6B RPT3 (10), and heat-shock factor 1 (HSF1) (11) for DYRK2, whereas the −3 R is lacking on multiple DYRK2 substrates such as p53 (12, 13), c-Jun, c-Myc (14), and SIAH2 (15). For those lacking the +1 P, the substrates have exhibited no redundant kinases within the related CMGC superfamily thus far (10). Among the DYRKs, DYRK2 often functions in tandem with related CMGC kinase, GSK3, in sequentially phosphorylating various substrates (7, 9, 16, 17). DYRK2 provides a priming phosphorylation for further GSK3 activity (7, 9, 16, 17). DYRK2 has been identified in all eukaryotes (10, 18), and interestingly across all orthologues, the conserved biological function of the DYRK2 isoform is regulation of cell division and/or tissue development (10, 18). A recent work has shown that DYRK2 is an essential kinase during embryogenesis, and mouse embryos with homozygous deletion of DYRK2 exhibit stunted development and pups die just before birth (19).
Of all the DYRK isoforms, DYRK2 is the only member that functions as a kinase activity–independent scaffold for an E3 ubiquitin ligase complex (20, 21, 22, 23, 24). DYRK2 is an integral part of the EDVP (EDD [ubiquitin protein ligase] + DDB1 [damage-specific DNA-binding protein] + VPRBP [HIV-1 Vpr-binding protein]) E3 ubiquitin ligase complex that carries out phosphorylation-mediated degradation of various cell cycle components to ensure smooth transition of G2/M stages of cell cycle (20, 21, 22, 23, 24). Some of the cancer mutations in Figure 1B are thought to affect efficient EDVP complex formation (3). Thus, over the past few decades, many groups have identified various molecular mechanism and substrates for DYRK2 playing diverse roles in cellular growth, proliferation, and developmental processes with a focal point being its role in cancer (10, 25, 26, 27, 28).
Besides DYRK2, the other DYRK isoforms, especially the class I's, have a long history in the field of cancer. Although DYRK1B has an overall protumorigenic role specifically in pancreatic and ovarian cancers, DYRK1A exhibits a more controversial role with reports of both protumorigenic and antitumorigenic mechanism in different cancers (reviewed in Boni et al. [8]). Within the class II DYRKs, very little is known about DYRK3 and DYRK4 with limited literature pointing to a more protumorigenic role for both (29, 30). DYRK2, on the other hand, is the most extensively studied class II isoform, and the high-profile substrates reported, such as p53 (12, 13), c-Jun (14), c-Myc (14), NOTCH1 (31), HSF1 (11), 26S proteasome (10, 25, 26), and SIAH2 (15, 31), have brought the kinase to the forefront of oncology research. For the past 2 decades, multiple studies have reported an overarching tumor suppressor role of DYRK2 across various cancers (reviewed in Yoshida and Yoshida [27]), with antitumorigenic roles including regulation of cell cycle, apoptosis, epithelial–mesenchymal transition (EMT), cancer stemness, and antimetastatic roles (reviewed in Yoshida and Yoshida [27]). On the other hand, since 2016, multiple studies report major protumorigenic roles of DYRK2 (10, 11, 25, 26), and a few studies have identified DYRK2 as a possible cancer driver (32, 33, 34). Furthermore, mRNA expression analyses from The Cancer Genome Atlas (TCGA) tumors along with matched normal controls reveal that the majority of cancers have higher median expression of DYRK2 than adjacent normal tissues (26), and a similar pattern has been shown for DYRK2 protein levels in some tumor types (11, 26). All of these data suggest that DYRK2 might be an excellent potential drug target. With such high profile, oncology-related substrates, could the function of DYRK2 differ based on cancer type or cell type? To shed some light onto this question, this review will re-examine the current literature on the role of DYRK2 in cancer and follow up with existing knowledge of small-molecule inhibitors developed to target DYRK2.
DYRK2 regulates proteostasis: an oncogenic role
DYRK2 maintains proteostasis of cancer cells by regulating two major players of the proteotoxic response pathway, which promotes the proper folding and/or degradation of proteins (Fig. 2). More than 90% of all solid human tumors carry numerous aberrations in chromosomes, referred to as aneuploidy (35). As a result of their severe aneuploidy, cancer cells are exposed to proteotoxic stress that increases the amount of toxic, unfolded proteins in the cell (36, 37). To survive proteotoxic stress, cancer cells can either increase protein folding capacity (controlled by the transcription factor HSF1) or increase the degradation of the misfolded/aggregated proteins (via the 26S proteasome and/or autophagy). DYRK2 phosphorylates and activates both HSF1 and the 26S proteasome and thereby activates the proteotoxic stress pathway promoting tumorigenesis in cancers such as triple-negative breast cancer (TNBC) and multiple myeloma (MM).
DYRK2 regulates 26S proteasome function
In 2016, an RNAi kinase screen identified DYRK2 as a kinase-regulating 26S proteasome activity (10). The study showed that DYRK2 depletion either by si/shRNA or CRISPR/Cas9 KO led to a 30 to 40% decrease in proteasome activity (10). The mature 26S proteasome is a complex of more than 30 distinct subunits that catalyzes 80% of eukaryotic protein degradation and harbors three distinct peptidase activities in the core subunit (chymotryptic, tryptic, and caspase-like) (38, 39). Besides the core of the proteasome, the complex also consists of the 19S regulatory subunit that binds to ubiquitylated proteins, whereas a six-membered ATPase ring hydrolyzes the protein into a polypeptide chain for entry into the peptidase core for degradation (40). Interestingly, DYRK2 phosphorylates the Rpt3 subunit on the ATPase ring of the 19S subunit of the proteasome on an evolutionarily conserved Thr25 site (10). Rpt3 pT25 had been previously reported by Steve Gygi's group in their 2008 work on quantitative phosphoproteomics of mitosis (41), but the function of the phosphorylation was not known. A phospho-specific antibody generated against pT25 Rpt3 showed that the site is dynamically upregulated during G2/M stage of the cell cycle and that serum starvation leads to loss of Thr25 phosphorylation (10). Furthermore, CRISPR/Cas9 knock-in of a phospho-deficient Thr25Ala on Rpt3 mimics the DYRK2 KO phenotypes in cells wherein there is a delay in mitotic progression, slower cell proliferation rates, and inhibition of all three peptidase activities of the 26S proteasome (10). The 26S proteasome degrades nearly 80% of all eukaryotic proteins, and hence, a 30% loss in activity leads to significant proteotoxic stress and consequent cell death in breast cancer cells (10). Intriguingly, DYRK2 KO cells were significantly more sensitive to the proteasome inhibitor, bortezomib, suggesting DYRK2 could be a possible therapeutic target for treatment of cancer (10). Indeed, in an ectopic nude mouse xenograft model, DYRK2 KO and T25A Rpt3 knock-in cells were less efficient in generating a tumor as compared to parental cells (10). This study further established the DYRK2-proteasome axis as potentially tumor promoting because higher expression of DYRK2 significantly correlated with higher mortality and poorer relapse-free survival in patients with breast cancer (10). In fact, inhibition or genetic depletion of DYRK2 tipped the scales of proteostasis in TNBC and MM cells. DYRK2 mRNA levels are higher in newly diagnosed and relapsed MM than normal donors (26). In fact, mice bearing syngrafted/xenografted myeloma cells with genetic depletion of DYRK2 exhibit significantly slower myeloma disease progression and reduced bone degeneration (26). Furthermore, bortezomib-resistant RPMI8226 myeloma cells express higher protein levels of DYRK2 than nonresistant RPMI8226 (26), suggesting that DYRK2 might play a role in driving drug resistance in some myeloma cases. The potent and selective DYRK2 inhibitor, LDN192960, induces cytotoxicity in myeloma cells both in vitro and in vivo with minimal off-target effects (26). The fact that the DYRK2 inhibitor alleviates myeloma burden in vivo suggests DYRK2 could indeed be a viable in vivo target for myeloma therapeutics. Resistance to proteasome inhibitors have been reported in patients, and this is either brought about by cancer mutations in the proteasome core or via upregulation of HSF1-mediated proteotoxic response pathway.
DYRK2 phosphorylates HSF1 and modulates proteotoxic response
The transcription factor HSF1 is the master regulator of proteotoxic stress responses and supports oncogenesis by helping cancer cells cope with the proteotoxic stress associated with both aneuploidy and oncogenic mutations. This has been demonstrated by the reduced susceptibility of Hsf1-KO mice to tumor formation driven either by Ras/p53 mutations or by chemical carcinogens (42, 43). Furthermore, high levels of HSF1 expression associate with poor outcome of various cancers (44). Upon proteotoxic stress, HSF1 is activated, translocates to the nucleus (45), and initiates the transcription of heat-shock proteins. Heat-shock proteins then function as molecular chaperones, protecting cells against proteotoxic stress by assisting in protein folding (46). HSF1 activity and stability are tightly controlled by multiple post-translational modifications (47). Among these, phosphorylation of serine 320 and serine 326 is associated with stability and nuclear accumulation followed by enhanced transcriptional activity of HSF1 (48, 49, 50). DYRK2 positively regulates HSF1 nuclear stability and activity, by phosphorylating it at Ser320 and Ser326 in TNBC cells (11). Indeed, DYRK2-depleted TNBC cells were far more sensitive to heat shock–mediated proteotoxic stress than parental cells, thus corroborating that DYRK2 plays a major role in maintaining proteostasis in TNBC cells. This link between DYRK2 and HSF1 is also observed in TNBC tumor samples, wherein a marked correlation was observed between high DYRK2 levels and high nuclear HSF1 levels.
The HSF1 pathway and the proteasome are not just two of the main pathways maintaining cell proteostasis, but they are interconnected and can compensate for each other. As mentioned before, proteasome inhibitors lead to the activation of HSF1 (Fig. 2) in an effort to protect the cell against the accumulation of toxic proteins (51, 52). The cytoprotective response mediated by HSF1 counteracts the cytotoxic effect of proteasome inhibitors (51, 52, 53), and thus, HSF1 inhibition might be effective to overcome proteasome inhibitor resistance in cancer cells. In that sense, a DYRK2 inhibitor induced cytotoxicity even in MM cells resistant to proteasome inhibitors (25, 26), suggesting that in fact DYRK2 inhibition might be targeting different complementary pathways. This observation was further echoed by a recent study showing that MM cells were extremely sensitive to increased temperatures and heat shock (54). In fact, combining heat shock with proteasome inhibitors led to higher accumulation of misfolded proteins leading to acute proteotoxic stress and apoptosis in the myeloma cells (54). Because cancer cells harbor significantly higher misfolded proteins than normal cells, targeting DYRK2 could indeed tip the scales for proteostasis in malignant cells and provide a significant therapeutic window for targeting specific cancers. This is indeed the case because normal/noncancerous cells were far more resistant to DYRK2 inhibitors (25, 26). Thus, targeting DYRK2 can significantly affect proteostasis (Fig. 2) via perturbation of both HSF1 and 26S proteasome activity leading to cancer cell death.
Hence, in the context of TNBC and MM, DYRK2 plays an overarching role as an oncogenic kinase and a potential therapeutic target.
DYRK2-p53 tumor suppressor link
A major molecular mechanism by which DYRK2 has been reported to exhibit the antitumorigenic role is via phosphorylation of tumor suppressor p53 on serine46 (Ser46). Upon genotoxic stress, energy stress, or heat shock, multiple CMGC kinases such as homeodomain-interacting protein kinase (HIPK2), mitogen-activated protein kinase p38α, and DYRK2 can phosphorylate p53 on Ser46, which triggers transcription of proapoptotic genes leading to cell death or cell senescence (reviewed in Liebl and Hofmann [55]). Upon DNA damage, DYRK2 is phosphorylated by ataxia-telangiectasia mutated kinase, which protects DYRK2 from proteasomal degradation leading to its nuclear accumulation where it phosphorylates p53 on Ser46 and promotes its transcriptional tumor suppressor activity (12, 13). Although phosphorylated Ser46 on p53 is indeed a marker for its tumor suppressor role, DYRK2 by no means is the exclusive kinase here. With multiple kinases including PKCδ, HIPK2, ataxia-telangiectasia mutated kinase, and p38α phosphorylating Ser46 upon genotoxic stress (55), it is hard to decipher to what extent DYRK2 contributes to this tumor suppressor role. Furthermore, p53 is mutated or truncated in a vast number of solid tumors and cancer patients with altered p53 exhibit significantly poorer survival (56, 57). Mutated p53 often exhibits stoichiometrically lower phosphoSer46 (58) and has been reported to trigger pro-oncogenic functions upon phosphorylation (59). This suggests that p53 phosphorylated on Ser46 serves as a tumor suppressor only in the few percentage of cancers containing WT p53 where patients exhibit better chances of survival.
Multiple publications carrying out sequencing or immunohistochemistry to study mRNA/protein levels of DYRK2 have suggested that DYRK2 is a tumor suppressor in colorectal (60, 61, 62, 63), liver (64), brain (65), and lung cancers (66, 67) and that the kinase promotes chemosensitivity in ovarian cancer (68). However, ovarian, liver, brain, lung, and colorectal cancers exhibit some of the highest mutations and variant allele frequencies in p53 compared with other cancer types (56, 57, 69). Thus, it is unclear to what extent DYRK2's phosphorylation of p53 could play as a tumor-suppressive role in these solid tumors exhibiting p53 mutation or loss. Furthermore, in endothelial cells, the pan-DYRK inhibitor, harmine (albeit with possible off-target effects), promotes p53 phosphorylation on Ser15, Ser20, and Ser37 (70), leading to higher p53 protein levels upon DNA damage (70, 71). Seemingly, in this case, DYRK2 inhibition led to tumor suppression. Hence, it is also important to decipher the molecular functions of DYRK2 in noncancer models or as a potential cancer driver. A recent unbiased deep multiomics study looking at the proteome, phosphoproteome, and transcriptome of murine high-grade brain cancer glioma model reported 41 kinases including DYRK2 exhibiting higher activity and rewired substrate signaling (34). Furthermore, the glioma murine model was generated by intracranial implantation of genetically engineered p53 null astrocytes, thus making the tumor-suppressor role of DYRK2-p53 axis highly untenable in this model.
Besides solid tumors, chronic myeloid leukemia (CML) cell lines exhibit significantly lower protein levels of DYRK2 than other hematological cancer cell lines (72). Interestingly, transcriptional upregulation of DYRK2 inhibits survival and self-renewal of CML stem/progenitor cells via c-Myc depletion and p53 activation (72). This tumor-suppressor role of DYRK2 seems to be CML specific because all other leukemia subtypes tested exhibited naturally elevated protein levels of DYRK2 at basal conditions (72), suggesting alternate driving mechanisms for tumorigenesis. Hematological malignancies exhibit fewer p53 mutations/loss (56, 57), and hence, DYRK2 could indeed be a tumor suppressor in specific subtypes such as CML. On a similar note, silencing DYRK2 has been reported to increase cell proliferation and reverse cell adhesion–mediated drug resistance in non-Hodgkin's lymphoma cell lines (73). Intriguingly, MM cells are highly sensitive to DYRK2 inhibition irrespective of p53 status (25, 26). Because DYRK2 inhibition in myeloma tips the balance of proteotoxic stress (26, 54), all cells whether p53 WT (such as cell line MM.1S) or mutated (such as cell lines RPMI8226 and U266B1) die (26). This suggests that in myeloma, the role of DYRK2 as an oncogenic driver probably plays a far greater role than its tumor-suppressor function potentiated by p53 phosphorylation. Hence, stratification of cancer subtypes before assigning molecular functions to DYRK2 is important. However, DYRK2 has tumor-suppressor mechanisms beyond p53 involvement, and it is important to investigate the diverse mechanisms at play to derive a larger perspective (Fig. 3).
Other molecular mechanisms linking DYRK2 and cancer
Various p53-independent tumor-suppressor mechanisms have been reported for DYRK2, while other substrates point to an oncogenic role. Each mechanism focuses on specific cancer types and subtypes (Fig. 3 and Table 1). The main substrates and mechanisms are critically presented below.
Table 1.
Cancer | Molecular partner | Phosphosite | Reported role | Pending question/issue |
---|---|---|---|---|
Breast | 26S proteasome/RPT3 (10, 25, 26) | Thr25 | Oncogenic; regulates proteostasis | Data specific for triple-negative breast cancer subtype. |
HSF1 (11) | Ser320, Ser326 | Data specific for triple-negative breast cancer subtype; possible redundancy with other kinases. | ||
SNAIL (83) | Ser104 | Tumor suppressor; EMT downregulation | Data heavily reliant on ectopic overexpression; possible redundancies with other kinases | |
NOTCH1 (31) | Thr2512 | Tumor suppressor; reduces invasion | Redundancies with other kinases; C-Myc pSer62 can be oncogenic; redundancies with other kinases; small sample sizes used for in vivo work. | |
C-Myc (14) | Ser62 | |||
C-Jun (14) | Ser243 | Possible redundancies with other DYRK family kinases | ||
p53 (12, 13) | Ser46 | Tumor suppressor; proapoptotic upon genotoxic stress | Most tumors are p53 mutated/null and mutated p53 becomes oncogenic; DYRK2 mRNA strongly overexpressed in breast cancer overall. | |
CDK14 (129) | n/a | Tumor suppressor; reduces invasion and proliferation | No specific mechanism reported on how DYRK2 regulates CDK14 transcription. | |
Lung | SIAH2 (86) | Ser16, Thr26, Ser28, Ser68, Thr119 | Tumor suppressor; modulates hypoxia response pathways | DYRK2 is strongly overexpressed in lung adenocarcinoma; redundancies with other kinases |
Ovarian | With EDVP/EDD only (89) | n/a | Oncogenic; degrades proapoptotic MOAP1; chemoresistance | Phosphorylated substrate (if any) not established. |
SNAIL (68) | Ser104 | Tumor-suppressor; EMT downregulation; chemoresistance | Data heavily reliant on ectopic overexpression; possible redundancies with other CMGC kinases; 2 cell lines used only. | |
Brain | PI3K/AKT/GSK3β (65) | n/a | Tumor-suppressor; EMT downregulation | DYRK2 mRNA is strongly overexpressed in glioma; DYRK inhibitors kill glioma cells. |
Unknown (unbiased multiomics) (34) | Oncogenic; DYRK2 increased activity and rewired signaling | Multiomics data derived from p53 null murine glioma models; no direct mechanism studied. | ||
Multiple myeloma | 26S proteasome/RPT3 (25, 26) | Thr25 | Oncogenic; regulates proteostasis | n/a |
Leukemia | p53/c-myc/KLF4 (72) | n/a | Tumor suppressor; reduces cancer stemness | Data specific for chronic myeloid leukemia subtype. |
Liver | p53/c-myc (62, 64, 130) | n/a | Tumor suppressor; EMT downregulation; reduces invasion; chemoresistance | p53 and c-Myc have extensive oncogenic mutations reported in liver cancer. |
LNC-HC/hsa-miR-183-5p (104) | Tumor suppressor; transcriptional upregulation of DYRK2 | Multiple tumor suppressors upregulated including DYRK2; no specific DYRK2 mechanism reported. | ||
Colorectal | p53 (12) | Ser46 | Tumor suppressor; proapoptotic upon genotoxic stress | Most colorectal tumors are p53 mutated/null, and mutated p53 becomes oncogenic; DYRK inhibitor promotes p53 phosphorylation. |
DNMT1 (103) | n/a | Tumor suppressor; epigenetic downregulation of DYRK2 | No specific DYRK2 mechanism reported. | |
miR-622 (63) | Tumor suppressor; EMT downregulation; reduces invasion | |||
Lymphoma | CDK2/p27Kip1 (73) | n/a | Tumor suppressor; EMT downregulation; chemoresistance | Data specific for non-Hodgkin's lymphoma subtype. |
CMGC, Cyclin-dependent kinases, Mitogen-activated protein kinases, Glycogen synthase kinases, and CDC-like kinases; DYRK, Dual-specificity tYrosine phosphorylation–Regulated Kinase; EMT, epithelial–mesenchymal transition; HSF1, heat-shock factor 1; MOAP, modulator of apoptosis protein 1; NOTCH1, neurogenic locus notch homolog protein 1; n/a, not directly reported.
Also, refer Figure 3.
c-Myc
DYRK2 has been reported to exhibit a p53-independent tumor-suppressor role by phosphorylating c-Myc on serine62 (Ser62) (14). c-Myc is a major proto-oncogenic transcription factor known to be overexpressed and mutated in various cancers (74). Post-translational modifications of c-Myc have been a topic of much debate over the past 30 years in which sequential phosphorylation of Ser62 and Threonine58 (Thr58) seems to play major roles in c-Myc transactivation (75, 76). The consensus in the field is that Thr58 is a GSK3-phosphorylation site while Ser62 seems to be the priming site for GSK3 activity, and similar to phosphorylation of p53 at Ser46, various CMGC kinases have been proposed (75), including DYRK2 to phosphorylate c-Myc. Dual phosphorylation of c-Myc on Thr58 and Ser62 triggers binding to an E3 ubiquitin ligase SCF-Fbxw7 (Skp-Cullen-F-box) and consequent proteasomal degradation of c-Myc (14), thus leading to the proposed tumor suppressor role of DYRK2. As stated previously, transcriptional upregulation of DYRK2 in CML promotes c-Myc degradation (72). In Burkitt lymphoma, nearly 60% of patients exhibit mutation of the GSK3 site, Thr58 (77), whereas primary cells exhibit lower levels of Thr58 phosphorylation (76), thus suggesting a tumor-suppressor role of GSK3 in this context. However, increased Ser62 phosphorylation has been observed in immortalized cells compared with primary cells (76), and monophosphorylation of Ser62 has been linked to c-Myc stabilization and higher transcriptional activity in multiple studies (78, 79). This indicates a similar conundrum as observed with the p53 Ser46 site wherein multiple kinases and diverse cancer subtypes exhibit altered mechanisms of action of major cancer-associated genes such as p53 and c-Myc.
c-Jun
A similar story is observed in case of c-Jun wherein two phosphorylation sites serine249 (Ser249: a bona fide GSK3 site) and Ser243 (reported to be phosphorylated by DYRK2) have been reported (14, 80). c-Jun is a transcription factor with established oncogenic roles (80). Similar to c-Myc, the E3 ubiquitin ligase SCF-Fbxw7 degrades c-Jun upon dual phosphorylation of Ser249 and Ser243 (81). Unlike c-Myc, Ser243 on c-Jun could be a DYRK-specific site because a previous study has elegantly ruled out most of the other CMGC kinase families (80). The same study did, however, observe redundancies between DYRK1A and DYRK2 for Ser243 on c-Jun in vitro (80), which is not surprising because the site is a +1P. In fact, dephosphorylation of Ser243 enhances c-Jun transcriptional activity in patients with cervical cancer exhibiting lower phosphoSer243 c-Jun in their tumors (82). Although the question of intra-DYRK redundancy remains, phosphoSer243 on c-Jun could indeed be a tumor-suppressor marker in specific cancers.
SNAIL
DYRK2 has also been reported to phosphorylate the zinc finger domain containing protein SNAIL that plays essential roles during development by triggering EMT (68, 83). DYRK2 knockdown led to upregulation of mesenchymal markers with consequent downregulation of epithelial E-cadherin mRNA in colon cancer (61) and promotion of proliferation and migration of glioma cells (65). This observation was consistent with other studies reporting downregulation of DYRK2 in metastatic colorectal secondary tumors found in the liver (62). SNAIL has been reported to be overexpressed in specific cancers and promote oncogenic progression by promoting EMT, invasion, and metastasis (83). DYRK2 phosphorylates serine104 on SNAIL that provides a priming site for GSK3, triggering the phosphorylation-mediated degradation of SNAIL (68, 83). This mechanism of antitumorigenic activity by DYRK2 is thought to promote chemosensitivity for ovarian cancer cells (68). A follow-up study reports that the DYRK2-mediated degradation of SNAIL is in fact reversed by p38α kinase (84). Although an interesting molecular mechanism, in both studies, DYRK2 ectopic overexpression has been carried out to justify the phosphorylation. Overexpression of CMGC kinases often leads to nonphysiological false-positive subcellular localizations and substrate identifications because of redundancy and high affinity for +1 P sites and hence further tools need to be used to confirm the DYRK2–SNAIL mechanism.
SIAH2
Seven In Absentia Homolog 2 or SIAH2 is an E3 ubiquitin ligase that plays a major role in targeted degradation of various proteins playing essential roles in regulating hypoxia (85). SIAH2 specifically regulates hypoxic tumor microenvironment by downregulation of key kinases in the Hippo signaling pathway (85). Furthermore, higher expression of SIAH2 is observed in lung cancer (86) and it plays oncogenic roles in castration-resistant prostate cancer (87). Interestingly, DYRK2 phosphorylates SIAH2 on 5 residues Ser16, Thr26, Ser28, Ser68, and Thr119. These modifications alter its subcellular localization thereby rewiring SIAH2 substrate specificity (15). SIAH2, in turn, is capable of degrading DYRK2 in specific cancers thereby triggering a protumorigenic hypoxic microenvironment (15, 85, 86). Some kinase redundancy has been observed wherein p38α kinase is capable of phosphorylating SIAH2 on same sites as DYRK2 (88); however, the DYRK2–SIAH2 link points to an interesting interplay between a kinase and a ubiquitin ligase regulating each other and thereby balancing protumorigenic and antitumorigenic roles.
EDVP E3 ubiquitin ligase
As stated previously, DYRK2 forms a kinase-independent scaffold for the EDVP E3 ligase complex and a recent study has reported loss-of-function point mutations of DYRK2 in cancer, which largely alters the interactome and substrate specificity of DYRK2 (3). The recurrent mutations (Fig. 1B) are thought to alter activity and/or formation of the EDVP complex (3). As part of the EDVP complex, DYRK2 phosphorylates and triggers degradation of multiple substrates such as katanin p60 (KATNA1) (23), telomerase reverse transcriptase (TERT) (22), and centrosome protein 110 (CP110) (21). Phosphorylation-mediated degradation of these substrates are required for proper cell cycle transitions especially the G2/M stage. Cancer mutations could result in incomplete EDVP complex formation, and incomplete EDVP can exhibit oncogenic prosurvival role because DYRK2+EDD alone degrades the proapoptotic factor modulator of apoptosis protein 1 independently of DDB1 and VPRBP in ovarian cancer (89). The substrates of DYRK2–EDVP exhibit both protumorigenic and antitumorigenic roles in various cancers thus adding further complexity. Ovarian cancer patients with higher levels of KATNA1 exhibit better overall survival (90); higher CP110 can decrease breast cancer cell invasion (91), yet lung cancer tissue expresses higher CP110 than the normal lung (92), while TERT is largely oncogenic (93). Thus, DYRK2–EDVP functions are tumor specific.
STAT3
DYRK2 has been reported to phosphorylate signal transducer and activator of transcription 3 (STAT3) in vitro (94). STAT3 is a transcription factor with both oncogenic and tumor-suppressor roles including regulation of tumor microenvironments (reviewed in Galoczova et al. [95]). STAT3 is phosphorylated on various residues upon interleukin/cytokine stimulation, and the phosphorylation on serine727 (Ser727) is thought to be an oncogenic biomarker in some subtypes of breast cancer (96). Although Ser727 is thought to promote the transcriptional activity of STAT3 (95), various kinases (CMGC family and beyond) have been reported to target Ser727 which is a +1P site (94). Thus, it is very difficult to dissect the importance of DYRK2 alone in driving phosphoSer727-mediated STAT3 activity.
TBK1
TANK-binding kinase 1 (TBK1) is an important upstream regulator of innate immune transcription pathways triggering type I interferon (IFN) translation and signaling in response to pathogens (97). DYRK2 phosphorylates TBK1 at serine527, which leads to phosphorylation-mediated degradation of TBK1 and downregulation of type I IFN signaling upon viral infection (97). Besides infections, the elevated presence of type I IFN correlates with a favorable prognosis in patients with different cancers (98, 99). In fact, reduced IFN-related gene expression leads to an immunosuppressive tumor microenvironment resulting in immunotherapy resistance in many solid tumors (99). Thus, DYRK2-mediated downregulation of IFN signaling could play a major oncogenic role in triggering immunotherapy resistance in various cancers. However, the study reporting DYRK2 as the upstream kinase of TBK1 relies on ectopic overexpression of DYRK2 to demonstrate direct phosphorylation of a canonical +1P motif (97). There could be redundancies with other CMGC kinases at that site which needs to be addressed more thoroughly.
NOTCH1
In response to chemotherapeutic agents, DYRK2 facilitates phosphorylation-mediated degradation of neurogenic locus notch homolog protein 1 (NOTCH1), which acts as an antiproliferative mechanism in breast cancer cells (31). NOTCH1 is a single transmembrane receptor and triggers intracellular signaling via binding to specific ligands (31). DYRK2 phosphorylates NOTCH1 on threonine2512 (Thr2512), which is a +1P site. However, NOTCH1 exhibits both tumor suppressor and oncogenic roles on a cancer-type basis (100). Interestingly, Thr2512 lies in the intracellular carboxy-terminal region of NOTCH1 that exhibits a PEST domain. The PEST region is the target of multiple CMGC kinases such as DYRK1A, HIPK2, CDKs, and GSK3, which triggers hyperphosphorylation and proteasomal degradation of NOTCH1 (reviewed in Lee et al. [101]). Thus, the redundancy conundrum remains to be solved to understand the function of NOTCH1's phosphorylation by DYRK2.
Transcriptional/epigenetic mechanisms
Besides modulating substrate phosphorylations, transcriptional and epigenetic mechanisms of DYRK2 regulation have also been proposed for some cancers. Specifically, the downregulation of DYRK2's gene expression has been linked to increased stemness in breast cancer (102) and CML (72) via upregulation of transcription factor Krüppel-like factor 4. DYRK2 expression was also downregulated transcriptionally by DNA methyltransferase 1 in colon cancer (103). The DYRK2 promoter region exhibited a higher level of methylation in cancer tissues than healthy tissues while treatment of cells with hypomethylating drug 5-azacytidine increased DYRK2 mRNA and protein levels (103). Furthermore, DYRK2 was reported to downregulate oncogenic miR-622 expression and reverse invasion of cancer cells (63), whereas long noncoding RNA long noncoding RNA derived from hepatocytes inhibits the proliferation of liver cancer cells by rescuing the expression of DYRK2 (104).
To reiterate, multiple molecular mechanisms have been proposed for DYRK2, and each mechanism is cancer-type or subtype specific (Fig. 3 and Table 1). The controversial role of DYRK2 is best highlighted in breast and lung cancers.
DYRK2 and breast cancer: a major controversy
TNBC
Various studies have focused on the role of DYRK2 in TNBC (10, 11, 25, 26). These studies revealed that both mRNA and protein levels of DYRK2 were higher in TNBC tumors than adjacent normal breast tissues (26). Complementing this information, a recent study with 715 samples of patients with breast cancer have shown that high protein levels of nuclear DYRK2 were associated with significantly reduced cancer survival and a shorter time to recurrence specifically within the TNBC subtype cohort (11). To test the potential therapeutic value of targeting DYRK2 in TNBC, three studies have compared the ability of parental and DYRK2-deficient TNBC cell lines to produce tumors in vivo (10, 25, 26). Crispr/Cas9-mediated DYRK2 deletion in MDA-MB-231 or MDA-MB-468 cells showed that tumors derived from TNBC–DYRK2–deficient cells had significantly slower growth rates and lower tumor burden than those derived from their parental cells. Importantly, two studies have shown that treatment with the DYRK2 inhibitors, curcumin and LDN192960, impaired growth of established TNBC tumors (25, 26). In contrast with these findings, other studies have used MDA-MB-231–derived xenografts and reported DYRK2 control EMT by degrading SNAIL (83) and promoting transcription factor Krüppel-like factor 4 expression (102), thereby functioning as a tumor suppressor. Both the studies used a DYRK2 overexpression system to show that higher DYRK2 decreased tumor formation. One study reported that mice xenografted with DYRK2-overexpressing MDA-MB-231 cells showed few metastatic lesions and a prolonged survival compared with those injected with control cells (83). In a second study, the authors compared the number of tumors produced by injecting increasing numbers of MDA-MB-231 cells with or without overexpressed DYRK2 (102). The authors used a sample size of n = 6 mice per condition and show that the total number of tumors derived from DYRK2-overexpressing cells was marginally lower than controls (102). This is in sharp contrast to others reporting DYRK2 depletion reduces proliferation and tumor formation potential of MDA-MB-231 cells (10, 11, 25, 26, 105). Some of these discrepancies might be due to the differential approaches used (DYRK2 knockdown/KO versus overexpression systems) or due to underpowered sample sizes. Furthermore, a phosphotyrosine proteomics study in TNBC cells reported that DYRK2 was among the top 5 phosphorylated proteins observed in aggressive basal-like TNBC cells (105). Because there is no evidence of the activation loop tyrosine exhibiting altered stoichiometric phosphorylation, the high levels of phosphorylation observed could be due to higher DYRK2 protein levels. In fact, siRNA knockdown of DYRK2 in basal-like TNBC MDA-MB-231 and HCC1395 cells lead to reduced proliferation, invasion, and colony formation potential of the cells (105).
Other breast cancer subtypes
Multiple studies looking at the role of DYRK2 in breast cancer have used the hormone receptor–positive and HER2-negative MCF7 cell line for xenograft studies. In the main study that supports the tumor-suppressor role of DYRK2 in breast cancer, the group identified DYRK2 as a priming kinase for c-Jun and c-Myc (14). In this study, the authors used a sample size of n = 3 mice per condition and carried out an orthotopic mammary-fat-pad breast cancer xenograft comparing MCF7 control cells and stable DYRK2 knockdown cells to investigate their ability to produce tumors (14). They found that DYRK2 knockdown cells clearly produce bigger tumors. Furthermore, DYRK2 knockdown cells showed higher invasion potential in vivo in an intracardiac injection model (n = 6 mice per condition). The same shRNA DYRK2 depleted cells were used in other studies as well to report the various tumor-suppressor roles of DYRK2 (102, 106). From the study with 715 samples of patients with breast cancer, no correlation was observed between DYRK2 expression and poor outcome in any of the receptor-positive breast cancer subtypes (11). However, TCGA data suggest that mRNA expression of DYRK2 is higher in breast invasive carcinoma and that higher DYRK2 expression correlates with poor survival in overall patients with breast cancer (8, 10, 26). Because mRNA and protein levels sometimes do not correlate, larger analysis looking at DYRK2 protein levels are needed to reach a finite conclusion.
The best way forward is to generate a conditional lox-cre mouse model for DYRK2 and generate hemizygous/homozygous deletion of DYRK2 in different subtypes of breast cancer genetically engineered mouse models (GEMMs) (107). Comparative tumor growth in the DYRK2 null versus parental GEMM over different subtypes would be a good way of addressing the pending questions on role of DYRK2 in breast cancer.
DYRK2 in lung cancer: unresolved issues
In 2003, the chromosome 12 region 12q13-14 was found to be amplified in adenocarcinomas of the lung and esophagus, and one of the resident genes, DYRK2, was significantly overexpressed in tumor samples as compared with normal tissues (33). In fact, DYRK2 exhibited the highest mRNA overexpression and highest copy numbers in tumors compared with normal tissue and other genes located in the 12q13-14 chromosomal region, suggesting that the overexpression of DYRK2 is the driving force behind the amplicon (33). This is reiterated in the TCGA lung adenocarcinoma and esophageal cancer cohort wherein tumor samples expressed higher DYRK2 mRNA than normal tissue (8). However, two independent studies report that higher protein or mRNA expression of DYRK2 is a favorable marker in pulmonary adenocarcinoma (66) and non–small-cell lung cancer (NSCLC) (67). In fact, pulmonary adenocarcinoma patients with higher DYRK2 expression exhibited a substantially higher 5-year survival than the group with lower DYRK2 expression. The higher DYRK2 levels associating with negative lymphatic invasion (66). Although the response rates to chemotherapy between the DYRK2-positive and DYRK2-negative patients were not different, patients with DYRK2+ tumors in recurrent NSCLC were suggested to have better outcome with chemotherapy (67). Mechanistically, in lung adenocarcinoma and squamous-cell lung cancer, E3 ubiquitin ligase SIAH2 targets DYRK2 for proteasomal degradation (86). SIAH2 protein and mRNA levels were found to be higher in samples of patients with lung cancer and exhibited a negative correlation with DYRK2 expression (86). Overall, the exact role of DYRK2 in lung neoplasia is still up for debate. Hence, using a similar strategy as suggested previously to generate conditional DYRK2 depletion in genetically engineered lung cancer mouse models for NSCLC, squamous-cell lung cancer, and other subtypes (108) could provide more clarity to this debate.
As reported previously, various global unbiased studies in various cancers have reported DYRK2 as a potential cancer driver with increased copy numbers, overexpression, and higher activity (32, 33, 34). On a similar note, a study using integrated high-resolution microarray analysis of gene copy number and expression in head and neck squamous-cell carcinoma cells reported that DYRK2 had the highest copy number and clear overexpression when compared with other genes in the 12q chromosomal amplicon (109). Furthermore, transcriptomics of blood identified DYRK2 as 1 of 10 potential prognostic biomarkers elevated in high-grade precancerous cervical lesions (110). Thus, unbiased identification of DYRK2 as a protein/kinase involved in potential protumorigenic role along with its substrates such as p53, c-Myc, and c-Jun further fuels the need to stratify cancers into subtypes before embarking on DYRK2 molecular studies. This duality of protumorigenic and antitumorigenic roles has been reported for the paralogue DYRK1A as well (111, 112) (Fig. 3 and Table 1), and hence, there is a clear precedence for such controversial roles in the DYRK family. One way of deconvoluting cancer-type and cell-type functions of a controversial kinase is by generating further tools such as potent and specific small-molecule kinase inhibitors.
Small-molecule inhibitors of DYRK2
Over the past three decades, various studies have been carried out to identify small-molecule inhibitors of kinases leading to the development of worldwide clinical trials and highly successful therapeutic targets and treatment options (113, 114). For the DYRKs, more than 60 reported small-molecule inhibitors have been published or are available in the public domain. ChEMBL (https://www.ebi.ac.uk/chembl) predicts that there are >1500 potential small molecules that can bind and possibly inhibit DYRK2, including established anticancer drugs sunitinib, erlotinib, afatinib, ruxolitinib, and crizotinib. A significant effort has been focused on development of DYRK1A small-molecule inhibitors because DYRK1A has established roles in neurodegenerative disorders. Consequently, early on the only available DYRK2 inhibitors were those targeting DYRK1A with off-target activity on DYRK2. DYRKs are canonical CMGC kinases and broad-spectrum ATP-competitive kinase inhibitors such as staurosporine and its derivatives inhibit DYRK2 at low nanomolar concentrations (https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors). Although there is a high degree of conservation between the kinase domains of class I and class II DYRKs, structural studies indicated that subtle amino acid substitutions in the hydrophobic inhibitor–docking pocket between DYRK1A and DYRK2 could confer significant degrees of inhibitor specificity (2). Interestingly, these amino acid substitutions contributed to the development and identification of various class-specific and often isoform-specific inhibitors for the DYRKs. Indeed, compound 5j that exhibited more than 100-fold sensitivity for DYRK1A over DYRK1B has no activity for class 2 DYRKs (115). Cocrystallization studies revealed that specific isoleucine to valine replacements in the docking site of curcumin resulted in a larger pocket in the class I DYRKs and thus reduced the shape complementarity to the inhibitor (25). Similarly, ID-8 an indole derivative exhibits an IC50 of <100 nM for class 1 DYRKs but >10 μM for class 2 DYRKs, suggesting significant room for developing specific inhibitors for the kinases (116). Similarly, β-carboline derivatives such as harmine or AnnH75 exhibit more in vivo and in vitro potency for class I than class II DYRKs (Table 2). However, the benzimidazole derivatives such as INDY, TG003, and DYR219 exhibit a pan-DYRK activity in vitro and in vivo (Table 2) and have been reported to trigger degradation of DYRK proteins when treated in cells (117, 118). This might explain some of the pronounced in vivo efficacy compared with in vitro observations for DYRK inhibitors wherein prolonged treatment leads to inhibition + degradation of the DYRK target, leading to a significant phenotype. Some promiscuous casein kinase inhibitors derived from benzimidazole potently inhibited DYRK1A and DYRK2 in vitro (119, 120). Silmitasertib (CX-4945), a potent and selective inhibitor of CK2 (with IC50 of 1 nM in vitro), is an orally bioavailable drug currently in phase 1/2 of clinical trials for cancer (121). Intriguingly, silmitasertib potently inhibits both class I and II DYRKs (122). The group did not report the IC50 for DYRK2; however, DYRK3 IC50 was reported to be 18 nM (122). Because the kinase domains of DYRK2 and DYRK3 are >90% similar at the amino acid level, there is a good chance that silmitasertib could indeed be a potent DYRK2 inhibitor as well. Silmitasertib exhibits blood–brain barrier penetrance similar to brain-penetrant DYR219 (118) and SM07883 (123) and could therefore potentially pharmacologically target the DYRKs in the brain.
Table 2.
Compound | Structure | %Inhibition/IC50 | Other kinase targets | References |
---|---|---|---|---|
1. Established potent and selective cell-permeable DYRK2 inhibitors | ||||
LDN192960 | 13 nM | Haspin (10 nM); PIM3 (10 nM) PIM1/2; DYRK1A (122 nM) DYRK1B; DYRK3 (<3 nM) | (26, 124, 125) | |
GSK626616 | <1 nM | DYRK3 (0.7 nM); DYRK1A/B; CK | (131) | |
Leucettine L41 | 35 nM |
CLK1 (15 nM) CLK 2 DYRK1A (40 nM) GSK3 (410 nM) |
(132) | |
EHT 5372 | 10.8 nM |
DYRK1A (0.22 nM) DYRK1B (0.28 nM); CLK1 (22.8 nM) CLK2 (88.8 nM); DYRK3 (93.2 nM); GSK3alpha (7.44 nM) GSK3Beta (221 nM) |
(133) | |
EHT 1610 | 3.16 nM |
DYRK1A (0.36 nM) DYRK1B (0.59 nM) CLK1 (11 nM) CLK2 (32.5 nM) CLK3 (1420 nM) DYRK3 (21.1 nM) GSK3 (9.11 nM) |
(134) | |
TG003 | 80% inhibition at 1000 nM | DYRK1A (24.01 nM) DYRK1B (34.39 nM) CLK1/2/3 DYRK3 |
https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors (133) | |
INDY | 27.7 nM |
DYRK1A (139 nM) DYRK1B (69.2 nM) CLK1/2/3; DYRK3; CK |
(135, 136) | |
2. Brain-penetrant DYRK2 inhibitors | ||||
DYR219 | benzimidazole analog | 90% inhibition at 10,000 nM |
DYRK1A (34 nM) DYRK1B (29 nM) GSK3; CDK5 |
(118) |
SM07883 | 3-Acylamino-isoquinoline analog | 16 nM |
DYRK1A (1.6 nM) DYRK1B (8 nM) GSK3Beta (10.8 nM) CLK4 (3 nM) |
(123) |
3. Potent DYRK2 lead compounds/analogs reported | ||||
RD0392 | 45.6 nM |
DYRK1A (60.2 nM) DYRK1B (53.3 nM) |
(136) | |
Abbott COT/TPL2 compound 41 | 70 nM | COT/TPL2; DYRK3; DYRK1A | https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors | |
Fluoro-DANDY analog 5 g | <50 nM |
DYRK1A (9.34 nM) DYRK1B; DYRK3 CDK2/5 GSK3 ERK2 |
(137) | |
SC97202 | 13.8 nM |
DYRK1A (<0.1 nM) DYRK1B (0.31 nM) DYRK3 (23.38 nM) |
(138) | |
LY2835219 | 61 nM |
CDKs PIM1 (50 nM) HIPK2 (31 nM) CK2 (117 nM) GSK3β (192 nM) JNK (389 nM) |
(139) | |
Milciclib | 48 nM |
CDK2 (45 nM) CLK1 (3 nM) CLK2 (1 nM) CLK4 (13 nM) DYRK1A (17 nM) DYRK1B (5 nM) |
(140) | |
PAC1 | 140 nM |
CLK1 (430 nM) CLK4 (400 nM) DYRK1A (230 nM) DYRK1B (400 nM) DYRK3 (280 nM) |
(141) | |
Compound 6i | 130 nM | DYRK1A (210 nM) | (142) | |
4. Other identified DYRK2 inhibitors | ||||
Curcumin | 5 nM | Highly nonspecific at >15-μM concentrations | (25) | |
Harmine | 900 nM |
CLKs; DYRK1A (80 nM) DYRK3 (800 nM); monoamine oxidase |
(127) | |
Ro3306 | 80% inhibition at 1000 nM | CDKs; PKC; SGK | https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors | |
BX-795 | 72% inhibition at 1000 nM | PDK1; AMPK; NUAK1 | https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors (143) | |
A-443654 | 74% inhibition at 100 nM | AKT; PKA; DYRK1A; DYRK3 | https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors | |
AnnH75 | 58% inhibition at 1000 nM |
CLK1; DYRK1A (181 nM) Haspin |
(144) | |
WNK-463 | 80% inhibition at 1000 nM | WNKs | https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors | |
GNF2133 | 95% inhibition at 2000 nM |
DYRK1A (6.2 nM) DYRK1B; DYRK3; CLK1/2; RIPK2; FLT3/4 |
(145) |
The kinase in bold format indicates the original target reported for the compound. IC50 values are provided where available.
A recently well-characterized DYRK2 inhibitor is the acridine analog, LDN192960 (26). Initially developed as a Haspin inhibitor (124, 125), LDN192960 inhibited DYRK2 with 13 nM IC50 and exhibited antitumor activity in mouse models of MM and TNBC (26). Interestingly, LDN192960 exhibited a ‘mixed’ mode of DYRK2 inhibition and cocrystal structure of LDN192960 with DYRK2 revealed that two water molecules mediated multiple hydrogen bonds between LDN192960 and DYRK2 active pocket (26). Similar water molecule–mediated interaction was also observed in the cocrystal structure of DYRK1A with inhibitor DJM2005 wherein a water molecule facilitated hydrogen bonding to further stabilize the inhibitor bound structure (2). LDN192960 reduced tumor burden of syngrafted and patient-derived xenografted mouse models of TNBC and delayed bone degeneration of allografted MM mouse models (26). In cells, LDN192960 exhibited potent cytotoxicity toward cancer lines with minimal impact on noncancerous cells (26). In fact, LDN192960 induced cytotoxicity to CD138+ primary myeloma cells of patients with significantly less impact on matched peripheral mononuclear cells (26). LDN192960 also exhibited additive effects in combination with FDA-approved proteasome inhibitor carfilzomib in inducing cytotoxicity in myeloma cells (26). LDN192960 was bioavailable in vivo and a dose of 50 mg/Kg body weight was sufficient to target neoplasia (26). Thus, LDN192960 inhibited DYRK2 in vivo and reduced 26S proteasome activity and thereby targeted proteasome-dependent neoplastic diseases such as TNBC and MM (25, 26).
Interestingly, curcumin inhibits DYRK2 with an IC50 of 5 nM and cocrystal structure revealed that curcumin binds to the active site pocket of DYRK2 via hydrophobic interactions (25). Curcumin is highly promiscuous, nonbioavailable, and unstable in the serum-free solution and has been labeled as both a pan-assay interference compound and an invalid metabolic panacea (126). Curcumin aggregates at concentrations greater than 15 μM, and most studies reporting controversial biological targets for curcumin used high 20- to 100μM concentrations leading to possible false positives (126). However, at lower 1- to 3μM concentrations, curcumin ablates DYRK2-mediated 26S proteasome phosphorylation in cells, reduces proteasome activity, and impairs cell proliferation in TNBC and MM cell lines in vitro and in vivo (25). DYRK2 KO cells exhibit no further off-target effects on proteasome activity with curcumin (25). Although neither a viable drug scaffold nor a highly potent kinase inhibitor, curcumin could serve as a decent compound for DYRK2 inhibition when used with proper controls.
In Table 2, we have listed those published inhibitors which have been tested directly on DYRK2 activity in vivo or in vitro along with a few interesting scaffolds identified in medicinal chemistry publications (https://www.kinase-screen.mrc.ac.uk/kinase-inhibitors) (127, 128). An important observation is that not a single one of these DYRK2 inhibitors exhibit protumorigenic/pro–cell-proliferation properties.
Conclusion and future perspectives
The controversial role of DYRK2 in cancer is evident (Fig. 3). Recent review articles mentioned the conflicting literature on DYRK2 (8, 28) and reported that mRNA expression data show that DYRK2 levels are higher in invasive breast carcinoma and lung adenocarcinoma than normal/adjacent tissue control (8, 26). However, others maintain that DYRK2 is a major tumor suppressor across all breast cancer subtypes and that DYRK2 depletion promotes proliferation (27). Indeed, being a CMGC kinase with a +1Pro active site, there are expected redundancies between DYRK2, other DYRKs, and possibly other CMGC kinases. A recent review article provides a list of substrates exhibiting overlapping DYRK kinases phosphorylating the same residues (8). Furthermore, as a member of the EDVP complex, cancer-specific mutations can alter DYRK2 substrate signature in specific cancer cells, which could be either tumor suppressive (3) or oncogenic (89). Although immunohistochemistry of tumors in patients with glioma shows lower levels of DYRK2 correlates with poorer survival (65), a receptor tyrosine kinase–transduced p53 null glioma mouse model exhibits higher DYRK2 activity and potentially altered signaling with diverse substrates (34). It is a definite possibility that we have just seen the tip of the iceberg when it comes to deconvoluting the redundancy and substrate overlap between the DYRKs and other CMGC kinases. As already shown, meticulous inhibitor screens and development of phospho-specific antibodies to the substrates could pave the way to dissect specificities and potential redundancies between the CMGC kinases (80). With the establishment of genetic engineering CRISPR KO/knock-in strategies and advanced quantitative phosphoproteomics, we might be able to dig deeper into identifying novel substrates and mechanisms in various cancer types. One major way forward would be to generate a conditional KO mouse model of DYRK2 in various cancer GEMM backgrounds to study tumor development and thereby ascertain the specific role of DYRK2 in each cancer subtype. A latest study reports that mouse embryonic fibroblasts with DYRK2 deletion exhibits significant downregulation of major cell cycle and proliferation drivers and markers such as Ki67, Aurora kinase A, PLK1, Bub1, and Bub1b (19). Although not in a cancer model, these observations are consistent with those observed in TNBC tumors where depletion of DYRK2 leads to reduced cell proliferation with greatly reduced Ki67 (10, 25, 26). However, DYRK2 could drive tumor-suppressor functions pairing with substrates reported in Figure 3 and Table 1 or other yet-undiscovered substrates in cancer type–specific mechanisms (27). In future, the roles of DYRK2 in cancer need to be approached in a more holistic way using multiple controlled models in each study including but not limited to genetic depletions and biochemical analyses, using specific inhibitors, in vivo animal models, and in vitro cell-based assays. Although overexpression of mRNA in cancers does indicate a potential oncogenic role, correlating that to corresponding increase of protein levels is important because mRNA and protein level often do not correlate in tumor samples. Before immunohistochemistry analysis on patient samples, proper antibody optimization steps are necessary while data analyses and sample size determinations need to be supported by proper statistical principles. Furthermore, ectopic overexpression of DYRK2 often results in false-positive substrate phosphorylation/binding and such experiments should always be accompanied with controls to ascertain the physiological/bona fide roles of the kinase. Like its paralogue DYRK1A (111, 112) and many other kinases, DYRK2 may indeed play both protumorigenic and antitumorigenic roles in different cancer types and subtypes, which is often determined by spatiotemporal interactions between kinases and specific substrates. Novel cancer therapeutic targets are a need of the hour, and hence, controversies delaying the establishment of a potential target or a tumor suppressor need to be objectively and quickly addressed. Deconvolution of the enigmatic roles of DYRK2 in various cancer types and subtypes should be prioritized among those in the field making our tools and expertise available for the greater scientific community in this endeavor.
Conflict of interest
The authors declare that they have no conflicts of interest with the contents of this article.
Acknowledgments
The authors thank Prof. Albena Dinkova-Kostova and Dr Carolyn Worby for valuable comments and Mr Vibhav Nadkarni for illustrations.
Author contributions
V. T., L. d. l. V., and S. B. wrote the article.
Funding and additional information
This work was supported by grants from the Ninewells Cancer Campaign PhD studentship (to V. T.), Cancer Research UK C52419/A22869 (to L. d. l. V.), and the Ninewells Cancer Campaign Cancer Research Award and Tayside Health Funds (to S. B.).
Edited by Alex Toker
References
- 1.Manning G., Whyte D.B., Martinez R., Hunter T., Sudarsanam S. The protein kinase complement of the human genome. Science. 2002;298:1912–1934. doi: 10.1126/science.1075762. [DOI] [PubMed] [Google Scholar]
- 2.Soundararajan M., Roos A.K., Savitsky P., Filippakopoulos P., Kettenbach A.N., Olsen J.V., Gerber S.A., Eswaran J., Knapp S., Elkins J.M. Structures of down syndrome kinases, DYRKs, reveal mechanisms of kinase activation and substrate recognition. Structure. 2013;21:986–996. doi: 10.1016/j.str.2013.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Mehnert M., Ciuffa R., Frommelt F., Uliana F., van Drogen A., Ruminski K., Gstaiger M., Aebersold R. Multi-layered proteomic analyses decode compositional and functional effects of cancer mutations on kinase complexes. Nat. Commun. 2020;11:3563. doi: 10.1038/s41467-020-17387-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Lochhead P.A., Sibbet G., Morrice N., Cleghon V. Activation-loop autophosphorylation is mediated by a novel transitional intermediate form of DYRKs. Cell. 2005;121:925–936. doi: 10.1016/j.cell.2005.03.034. [DOI] [PubMed] [Google Scholar]
- 5.Lee S.B., Ko A., Oh Y.T., Shi P., D'Angelo F., Frangaj B., Koller A., Chen E.I., Cardozo T., Iavarone A., Lasorella A. Proline hydroxylation primes protein kinases for autophosphorylation and activation. Mol. Cel. 2020;79:376–389.e378. doi: 10.1016/j.molcel.2020.06.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Kannan N., Neuwald A.F. Evolutionary constraints associated with functional specificity of the CMGC protein kinases MAPK, CDK, GSK, SRPK, DYRK, and CK2alpha. Protein Sci. 2004;13:2059–2077. doi: 10.1110/ps.04637904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Campbell L.E., Proud C.G. Differing substrate specificities of members of the DYRK family of arginine-directed protein kinases. FEBS Lett. 2002;510:31–36. doi: 10.1016/s0014-5793(01)03221-5. [DOI] [PubMed] [Google Scholar]
- 8.Boni J., Rubio-Perez C., López-Bigas N., Fillat C., de la Luna S. The DYRK family of kinases in cancer: Molecular functions and therapeutic opportunities. Cancers (Basel) 2020;12 doi: 10.3390/cancers12082106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Himpel S., Tegge W., Frank R., Leder S., Joost H.G., Becker W. Specificity determinants of substrate recognition by the protein kinase DYRK1A. J. Biol. Chem. 2000;275:2431–2438. doi: 10.1074/jbc.275.4.2431. [DOI] [PubMed] [Google Scholar]
- 10.Guo X., Wang X., Wang Z., Banerjee S., Yang J., Huang L., Dixon J.E. Site-specific proteasome phosphorylation controls cell proliferation and tumorigenesis. Nat. Cel. Biol. 2016;18:202–212. doi: 10.1038/ncb3289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Moreno Dorta R., Banerjee S., Jackson A., Quinn J., Baillie G., Dixon J.E., Dinkova-Kostova A., Edwards J., de la Vega L. The stress-responsive kinase DYRK2 activates heat shock factor 1 promoting resistance to proteotoxic stress. Cell Death Differ. 2020 doi: 10.1038/s41418-020-00686-8. Dec 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Taira N., Nihira K., Yamaguchi T., Miki Y., Yoshida K. DYRK2 is targeted to the nucleus and controls p53 via Ser46 phosphorylation in the apoptotic response to DNA damage. Mol. Cel. 2007;25:725–738. doi: 10.1016/j.molcel.2007.02.007. [DOI] [PubMed] [Google Scholar]
- 13.Taira N., Yamamoto H., Yamaguchi T., Miki Y., Yoshida K. ATM augments nuclear stabilization of DYRK2 by inhibiting MDM2 in the apoptotic response to DNA damage. J. Biol. Chem. 2010;285:4909–4919. doi: 10.1074/jbc.M109.042341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Taira N., Mimoto R., Kurata M., Yamaguchi T., Kitagawa M., Miki Y., Yoshida K. DYRK2 priming phosphorylation of c-Jun and c-Myc modulates cell cycle progression in human cancer cells. J. Clin. Invest. 2012;122:859–872. doi: 10.1172/JCI60818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Pérez M., García-Limones C., Zapico I., Marina A., Schmitz M.L., Muñoz E., Calzado M.A. Mutual regulation between SIAH2 and DYRK2 controls hypoxic and genotoxic signaling pathways. J. Mol. Cel. Biol. 2012;4:316–330. doi: 10.1093/jmcb/mjs047. [DOI] [PubMed] [Google Scholar]
- 16.Woods Y.L., Cohen P., Becker W., Jakes R., Goedert M., Wang X., Proud C.G. The kinase DYRK phosphorylates protein-synthesis initiation factor eIF2Bepsilon at Ser539 and the microtubule-associated protein tau at Thr212: Potential role for DYRK as a glycogen synthase kinase 3-priming kinase. Biochem. J. 2001;355:609–615. doi: 10.1042/bj3550609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Cole A.R., Causeret F., Yadirgi G., Hastie C.J., McLauchlan H., McManus E.J., Hernandez F., Eickholt B.J., Nikolic M., Sutherland C. Distinct priming kinases contribute to differential regulation of collapsin response mediator proteins by glycogen synthase kinase-3 in vivo. J. Biol. Chem. 2006;281:16591–16598. doi: 10.1074/jbc.M513344200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Aranda S., Laguna A., Luna S.d. l. DYRK family of protein kinases: Evolutionary relationships, biochemical properties, and functional roles. FASEB J. 2011;25:449–462. doi: 10.1096/fj.10-165837. [DOI] [PubMed] [Google Scholar]
- 19.Yoshida S., Aoki K., Fujiwara K., Nakakura T., Kawamura A., Yamada K., Ono M., Yogosawa S., Yoshida K. The novel ciliogenesis regulator DYRK2 governs Hedgehog signaling during mouse embryogenesis. Elife. 2020;9 doi: 10.7554/eLife.57381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Hossain D., Ferreira Barbosa J.A., Cohen E.A., Tsang W.Y. HIV-1 Vpr hijacks EDD-DYRK2-DDB1(DCAF1) to disrupt centrosome homeostasis. J. Biol. Chem. 2018;293:9448–9460. doi: 10.1074/jbc.RA117.001444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Hossain D., Javadi Esfehani Y., Das A., Tsang W.Y. Cep78 controls centrosome homeostasis by inhibiting EDD-DYRK2-DDB1(Vpr)(BP) EMBO Rep. 2017;18:632–644. doi: 10.15252/embr.201642377. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Jung H.Y., Wang X., Jun S., Park J.I. Dyrk2-associated EDD-DDB1-VprBP E3 ligase inhibits telomerase by TERT degradation. J. Biol. Chem. 2013;288:7252–7262. doi: 10.1074/jbc.M112.416792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Maddika S., Chen J. Protein kinase DYRK2 is a scaffold that facilitates assembly of an E3 ligase. Nat. Cel. Biol. 2009;11:409–419. doi: 10.1038/ncb1848. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Wang X., Singh S., Jung H.Y., Yang G., Jun S., Sastry K.J., Park J.I. HIV-1 Vpr protein inhibits telomerase activity via the EDD-DDB1-VPRBP E3 ligase complex. J. Biol. Chem. 2013;288:15474–15480. doi: 10.1074/jbc.M112.416735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Banerjee S., Ji C., Mayfield J.E., Goel A., Xiao J., Dixon J.E., Guo X. Ancient drug curcumin impedes 26S proteasome activity by direct inhibition of dual-specificity tyrosine-regulated kinase 2. Proc. Natl. Acad. Sci. 2018;115:8155. doi: 10.1073/pnas.1806797115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Banerjee S., Wei T., Wang J., Lee J.J., Gutierrez H.L., Chapman O., Wiley S.E., Mayfield J.E., Tandon V., Juarez E.F., Chavez L., Liang R., Sah R.L., Costello C., Mesirov J.P. Inhibition of dual-specificity tyrosine phosphorylation-regulated kinase 2 perturbs 26S proteasome-addicted neoplastic progression. Proc. Natl. Acad. Sci. U. S. A. 2019;116:24881–24891. doi: 10.1073/pnas.1912033116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Yoshida S., Yoshida K. Multiple functions of DYRK2 in cancer and tissue development. FEBS Lett. 2019;593:2953–2965. doi: 10.1002/1873-3468.13601. [DOI] [PubMed] [Google Scholar]
- 28.Correa-Sáez A., Jiménez-Izquierdo R., Garrido-Rodríguez M., Morrugares R., Muñoz E., Calzado M.A. Updating dual-specificity tyrosine-phosphorylation-regulated kinase 2 (DYRK2): Molecular basis, functions and role in diseases. Cell Mol. Life Sci. 2020;77:4747–4763. doi: 10.1007/s00018-020-03556-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Wippich F., Bodenmiller B., Trajkovska M.G., Wanka S., Aebersold R., Pelkmans L. Dual specificity kinase DYRK3 couples stress granule condensation/dissolution to mTORC1 signaling. Cell. 2013;152:791–805. doi: 10.1016/j.cell.2013.01.033. [DOI] [PubMed] [Google Scholar]
- 30.Owusu M., Bannauer P., Ferreira da Silva J., Mourikis T.P., Jones A., Májek P., Caldera M., Wiedner M., Lardeau C.-H., Mueller A.C., Menche J., Kubicek S., Ciccarelli F.D., Loizou J.I. Mapping the human kinome in response to DNA damage. Cell Rep. 2019;26:555–563.e556. doi: 10.1016/j.celrep.2018.12.087. [DOI] [PubMed] [Google Scholar]
- 31.Morrugares R., Correa-Sáez A., Moreno R., Garrido-Rodríguez M., Muñoz E., de la Vega L., Calzado M.A. Phosphorylation-dependent regulation of the NOTCH1 intracellular domain by dual-specificity tyrosine-regulated kinase 2. Cell Mol. Life Sci. 2020;77:2621–2639. doi: 10.1007/s00018-019-03309-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Gorringe K.L., Boussioutas A., Bowtell D.D. Novel regions of chromosomal amplification at 6p21, 5p13, and 12q14 in gastric cancer identified by array comparative genomic hybridization. Genes Chromosomes Cancer. 2005;42:247–259. doi: 10.1002/gcc.20136. [DOI] [PubMed] [Google Scholar]
- 33.Miller C.T., Aggarwal S., Lin T.K., Dagenais S.L., Contreras J.I., Orringer M.B., Glover T.W., Beer D.G., Lin L. Amplification and overexpression of the dual-specificity tyrosine-(Y)-phosphorylation regulated kinase 2 (DYRK2) gene in esophageal and lung adenocarcinomas. Cancer Res. 2003;63:4136–4143. [PubMed] [Google Scholar]
- 34.Wang H., Diaz A.K., Shaw T.I., Li Y., Niu M., Cho J.-H., Paugh B.S., Zhang Y., Sifford J., Bai B., Wu Z., Tan H., Zhou S., Hover L.D., Tillman H.S. Deep multiomics profiling of brain tumors identifies signaling networks downstream of cancer driver genes. Nat. Commun. 2019;10:3718. doi: 10.1038/s41467-019-11661-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Weaver B.A., Cleveland D.W. Does aneuploidy cause cancer? Curr. Opin. Cell Biol. 2006;18:658–667. doi: 10.1016/j.ceb.2006.10.002. [DOI] [PubMed] [Google Scholar]
- 36.Donnelly N., Storchová Z. Causes and consequences of protein folding stress in aneuploid cells. Cell Cycle. 2015;14:495–501. doi: 10.1080/15384101.2015.1006043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Ohashi A., Ohori M., Iwai K., Nakayama Y., Nambu T., Morishita D., Kawamoto T., Miyamoto M., Hirayama T., Okaniwa M., Banno H., Ishikawa T., Kandori H., Iwata K. Aneuploidy generates proteotoxic stress and DNA damage concurrently with p53-mediated post-mitotic apoptosis in SAC-impaired cells. Nat. Commun. 2015;6:7668. doi: 10.1038/ncomms8668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Collins G.A., Goldberg A.L. The logic of the 26S proteasome. Cell. 2017;169:792–806. doi: 10.1016/j.cell.2017.04.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Coux O., Tanaka K., Goldberg A.L. Structure and functions of the 20S and 26S proteasomes. Annu. Rev. Biochem. 1996;65:801–847. doi: 10.1146/annurev.bi.65.070196.004101. [DOI] [PubMed] [Google Scholar]
- 40.Besche H.C., Peth A., Goldberg A.L. Getting to first base in proteasome assembly. Cell. 2009;138:25–28. doi: 10.1016/j.cell.2009.06.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Dephoure N., Zhou C., Villen J., Beausoleil S.A., Bakalarski C.E., Elledge S.J., Gygi S.P. A quantitative atlas of mitotic phosphorylation. Proc. Natl. Acad. Sci. U. S. A. 2008;105:10762–10767. doi: 10.1073/pnas.0805139105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Dai C., Whitesell L., Rogers A.B., Lindquist S. Heat shock factor 1 is a powerful multifaceted modifier of carcinogenesis. Cell. 2007;130:1005–1018. doi: 10.1016/j.cell.2007.07.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Meng L., Gabai V.L., Sherman M.Y. Heat-shock transcription factor HSF1 has a critical role in human epidermal growth factor receptor-2-induced cellular transformation and tumorigenesis. Oncogene. 2010;29:5204–5213. doi: 10.1038/onc.2010.277. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Mendillo M.L., Santagata S., Koeva M., Bell G.W., Hu R., Tamimi R.M., Fraenkel E., Ince T.A., Whitesell L., Lindquist S. HSF1 drives a transcriptional program distinct from heat shock to support highly malignant human cancers. Cell. 2012;150:549–562. doi: 10.1016/j.cell.2012.06.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Baler R., Dahl G., Voellmy R. Activation of human heat shock genes is accompanied by oligomerization, modification, and rapid translocation of heat shock transcription factor HSF1. Mol. Cell Biol. 1993;13:2486–2496. doi: 10.1128/mcb.13.4.2486. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Jolly C., Morimoto R.I. Role of the heat shock response and molecular chaperones in oncogenesis and cell death. J. Natl. Cancer Inst. 2000;92:1564–1572. doi: 10.1093/jnci/92.19.1564. [DOI] [PubMed] [Google Scholar]
- 47.Dayalan Naidu S., Dinkova-Kostova A.T. Regulation of the mammalian heat shock factor 1. FEBS J. 2017;284:1606–1627. doi: 10.1111/febs.13999. [DOI] [PubMed] [Google Scholar]
- 48.Guettouche T., Boellmann F., Lane W.S., Voellmy R. Analysis of phosphorylation of human heat shock factor 1 in cells experiencing a stress. BMC Biochem. 2005;6:4. doi: 10.1186/1471-2091-6-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Murshid A., Chou S.D., Prince T., Zhang Y., Bharti A., Calderwood S.K. Protein kinase A binds and activates heat shock factor 1. PLoS One. 2010;5 doi: 10.1371/journal.pone.0013830. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Tang Z., Dai S., He Y., Doty R.A., Shultz L.D., Sampson S.B., Dai C. MEK guards proteome stability and inhibits tumor-suppressive amyloidogenesis via HSF1. Cell. 2015;160:729–744. doi: 10.1016/j.cell.2015.01.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Pirkkala L., Alastalo T.-P., Zuo X., Benjamin I.J., Sistonen L. Disruption of heat shock factor 1 reveals an essential role in the ubiquitin proteolytic pathway. Mol. Cell Biol. 2000;20:2670. doi: 10.1128/mcb.20.8.2670-2675.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Shah S.P., Lonial S., Boise L.H. When cancer fights back: Multiple myeloma, proteasome inhibition, and the heat-shock response. Mol. Cancer Res. 2015;13:1163. doi: 10.1158/1541-7786.MCR-15-0135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Sharma C., Seo Y.H. Small molecule inhibitors of HSF1-activated pathways as potential next-generation anticancer therapeutics. Molecules (Basel, Switzerland) 2018;23 doi: 10.3390/molecules23112757. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Sha Z., Goldberg A.L. Multiple myeloma cells are exceptionally sensitive to heat shock, which overwhelms their proteostasis network and induces apoptosis. Proc. Natl. Acad. Sci. U. S. A. 2020;117:21588–21597. doi: 10.1073/pnas.2001323117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Liebl M.C., Hofmann T.G. Cell fate regulation upon DNA damage: p53 serine 46 kinases pave the cell death road. BioEssays. 2019;41:1900127. doi: 10.1002/bies.201900127. [DOI] [PubMed] [Google Scholar]
- 56.Donehower L.A., Soussi T., Korkut A., Liu Y., Schultz A., Cardenas M., Li X., Babur O., Hsu T.-K., Lichtarge O., Weinstein J.N., Akbani R., Wheeler D.A. Integrated analysis of TP53 gene and pathway alterations in the cancer genome atlas. Cell Rep. 2019;28:1370–1384.e1375. doi: 10.1016/j.celrep.2019.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Kandoth C., McLellan M.D., Vandin F., Ye K., Niu B., Lu C., Xie M., Zhang Q., McMichael J.F., Wyczalkowski M.A., Leiserson M.D.M., Miller C.A., Welch J.S., Walter M.J., Wendl M.C. Mutational landscape and significance across 12 major cancer types. Nature. 2013;502:333–339. doi: 10.1038/nature12634. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Fanucchi S., Veale R.B. Role of p53/FAK association and p53Ser46 phosphorylation in staurosporine-mediated apoptosis: Wild type versus mutant p53-R175H. FEBS Lett. 2009;583:3557–3562. doi: 10.1016/j.febslet.2009.10.059. [DOI] [PubMed] [Google Scholar]
- 59.Muller P.A.J., Vousden K.H. Mutant p53 in cancer: New functions and therapeutic opportunities. Cancer Cell. 2014;25:304–317. doi: 10.1016/j.ccr.2014.01.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Leiszter K., Galamb O., Sipos F., Krenács T., Veres G., Wichmann B., Kalmár A., Patai Á V., Tóth K., Valcz G., Molnár B., Tulassay Z. Sporadic colorectal cancer development shows rejuvenescence regarding epithelial proliferation and apoptosis. PLoS One. 2013;8 doi: 10.1371/journal.pone.0074140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Yan H., Hu K., Wu W., Li Y., Tian H., Chu Z., Koeffler H.P., Yin D. Low expression of DYRK2 (dual specificity tyrosine phosphorylation regulated kinase 2) correlates with poor prognosis in colorectal cancer. PLoS One. 2016;11 doi: 10.1371/journal.pone.0159954. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Ito D., Yogosawa S., Mimoto R., Hirooka S., Horiuchi T., Eto K., Yanaga K., Yoshida K. Dual-specificity tyrosine-regulated kinase 2 is a suppressor and potential prognostic marker for liver metastasis of colorectal cancer. Cancer Sci. 2017;108:1565–1573. doi: 10.1111/cas.13280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Wang Y., Sun J., Wei X., Luan L., Zeng X., Wang C., Zhao W. Decrease of miR-622 expression suppresses migration and invasion by targeting regulation of DYRK2 in colorectal cancer cells. OncoTargets Ther. 2017;10:1091–1100. doi: 10.2147/OTT.S125724. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Zhang X., Xu P., Ni W., Fan H., Xu J., Chen Y., Huang W., Lu S., Liang L., Liu J., Chen B., Shi W. Downregulated DYRK2 expression is associated with poor prognosis and Oxaliplatin resistance in hepatocellular carcinoma. Pathol. Res. Pract. 2016;212:162–170. doi: 10.1016/j.prp.2016.01.002. [DOI] [PubMed] [Google Scholar]
- 65.Shen Y., Zhang L., Wang D., Bao Y., Liu C., Xu Z., Huang W., Cheng C. Regulation of glioma cells migration by DYRK2. Neurochem. Res. 2017;42:3093–3102. doi: 10.1007/s11064-017-2345-2. [DOI] [PubMed] [Google Scholar]
- 66.Yamashita S., Chujo M., Tokuishi K., Anami K., Miyawaki M., Yamamoto S., Kawahara K. Expression of dual-specificity tyrosine-(Y)-phosphorylation-regulated kinase 2 (DYRK2) can be a favorable prognostic marker in pulmonary adenocarcinoma. J. Thorac. Cardiovasc. Surg. 2009;138:1303–1308. doi: 10.1016/j.jtcvs.2009.08.003. [DOI] [PubMed] [Google Scholar]
- 67.Yamashita S., Chujo M., Moroga T., Anami K., Tokuishi K., Miyawaki M., Kawano Y., Takeno S., Yamamoto S., Kawahara K. DYRK2 expression may be a predictive marker for chemotherapy in non-small cell lung cancer. Anticancer Res. 2009;29:2753–2757. [PubMed] [Google Scholar]
- 68.Yamaguchi N., Mimoto R., Yanaihara N., Imawari Y., Hirooka S., Okamoto A., Yoshida K. DYRK2 regulates epithelial-mesenchymal-transition and chemosensitivity through Snail degradation in ovarian serous adenocarcinoma. Tumour Biol. 2015;36:5913–5923. doi: 10.1007/s13277-015-3264-y. [DOI] [PubMed] [Google Scholar]
- 69.Hussain S.P., Schwank J., Staib F., Wang X.W., Harris C.C. TP53 mutations and hepatocellular carcinoma: Insights into the etiology and pathogenesis of liver cancer. Oncogene. 2007;26:2166–2176. doi: 10.1038/sj.onc.1210279. [DOI] [PubMed] [Google Scholar]
- 70.Dai F., Chen Y., Song Y., Huang L., Zhai D., Dong Y., Lai L., Zhang T., Li D., Pang X., Liu M., Yi Z. A natural small molecule harmine inhibits angiogenesis and suppresses tumour growth through activation of p53 in endothelial cells. PLoS One. 2012;7 doi: 10.1371/journal.pone.0052162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Chehab N.H., Malikzay A., Stavridi E.S., Halazonetis T.D. Phosphorylation of Ser-20 mediates stabilization of human p53 in response to DNA damage. Proc. Natl. Acad. Sci. U. S. A. 1999;96:13777–13782. doi: 10.1073/pnas.96.24.13777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Park C.S., Lewis A.H., Chen T.J., Bridges C.S., Shen Y., Suppipat K., Puppi M., Tomolonis J.A., Pang P.D., Mistretta T.A., Ma L.Y., Green M.R., Rau R., Lacorazza H.D. A KLF4-DYRK2-mediated pathway regulating self-renewal in CML stem cells. Blood. 2019;134:1960–1972. doi: 10.1182/blood.2018875922. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Wang Y., Wu Y., Miao X., Zhu X., Miao X., He Y., Zhong F., Ding L., Liu J., Tang J., Huang Y., Xu X., He S. Silencing of DYRK2 increases cell proliferation but reverses CAM-DR in Non-Hodgkin's Lymphoma. Int. J. Biol. Macromol. 2015;81:809–817. doi: 10.1016/j.ijbiomac.2015.08.067. [DOI] [PubMed] [Google Scholar]
- 74.Wolf E., Eilers M. Targeting MYC proteins for tumor therapy. Annu. Rev. Cancer Biol. 2020;4:61–75. [Google Scholar]
- 75.Hann S.R. Role of post-translational modifications in regulating c-Myc proteolysis, transcriptional activity and biological function. Semin. Cancer Biol. 2006;16:288–302. doi: 10.1016/j.semcancer.2006.08.004. [DOI] [PubMed] [Google Scholar]
- 76.Lutterbach B., Hann S.R. Overexpression of c-Myc and cell immortalization alters c-Myc phosphorylation. Oncogene. 1997;14:967–975. doi: 10.1038/sj.onc.1200920. [DOI] [PubMed] [Google Scholar]
- 77.Bhatia K., Huppi K., Spangler G., Siwarski D., Iyer R., Magrath I. Point mutations in the c–Myc transactivation domain are common in Burkitt's lymphoma and mouse plasmacytomas. Nat. Genet. 1993;5:56–61. doi: 10.1038/ng0993-56. [DOI] [PubMed] [Google Scholar]
- 78.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. doi: 10.1101/gad.836800. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Seo H.R., Kim J., Bae S., Soh J.-W., Lee Y.-S. Cdk5-mediated phosphorylation of c-Myc on Ser-62 is essential in transcriptional activation of cyclin B1 by cyclin G1. J. Biol. Chem. 2008;283:15601–15610. doi: 10.1074/jbc.M800987200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Morton S., Davis R.J., McLaren A., Cohen P. A reinvestigation of the multisite phosphorylation of the transcription factor c-Jun. EMBO J. 2003;22:3876–3886. doi: 10.1093/emboj/cdg388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Wei W., Jin J., Schlisio S., Harper J.W., Kaelin W.G. The v-Jun point mutation allows c-Jun to escape GSK3-dependent recognition and destruction by the Fbw7 ubiquitin ligase. Cancer Cell. 2005;8:25–33. doi: 10.1016/j.ccr.2005.06.005. [DOI] [PubMed] [Google Scholar]
- 82.Huang C.C., Wang J.M., Kikkawa U., Mukai H., Shen M.R., Morita I., Chen B.K., Chang W.C. Calcineurin-mediated dephosphorylation of c-Jun Ser-243 is required for c-Jun protein stability and cell transformation. Oncogene. 2008;27:2422–2429. doi: 10.1038/sj.onc.1210888. [DOI] [PubMed] [Google Scholar]
- 83.Mimoto R., Taira N., Takahashi H., Yamaguchi T., Okabe M., Uchida K., Miki Y., Yoshida K. DYRK2 controls the epithelial-mesenchymal transition in breast cancer by degrading Snail. Cancer Lett. 2013;339:214–225. doi: 10.1016/j.canlet.2013.06.005. [DOI] [PubMed] [Google Scholar]
- 84.Ryu K.J., Park S.M., Park S.H., Kim I.K., Han H., Kim H.J., Kim S.H., Hong K.S., Kim H., Kim M., Yoon S.J., Asaithambi K., Lee K.H., Park J.Y., Hah Y.S. p38 stabilizes snail by suppressing DYRK2-mediated phosphorylation that is required for GSK3β-βTrCP-induced snail degradation. Cancer Res. 2019;79:4135–4148. doi: 10.1158/0008-5472.CAN-19-0049. [DOI] [PubMed] [Google Scholar]
- 85.Ma B., Chen Y., Chen L., Cheng H., Mu C., Li J., Gao R., Zhou C., Cao L., Liu J., Zhu Y., Chen Q., Wu S. Hypoxia regulates Hippo signalling through the SIAH2 ubiquitin E3 ligase. Nat. Cel. Biol. 2015;17:95–103. doi: 10.1038/ncb3073. [DOI] [PubMed] [Google Scholar]
- 86.Moreno P., Lara-Chica M., Soler-Torronteras R., Caro T., Medina M., Álvarez A., Salvatierra Á., Muñoz E., Calzado M.A. The expression of the ubiquitin ligase SIAH2 (seven in Absentia homolog 2) is increased in human lung cancer. PLoS One. 2015;10 doi: 10.1371/journal.pone.0143376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Qi J., Tripathi M., Mishra R., Sahgal N., Fazli L., Ettinger S., Placzek W.J., Claps G., Chung L.W., Bowtell D., Gleave M., Bhowmick N., Ronai Z.A. The E3 ubiquitin ligase Siah2 contributes to castration-resistant prostate cancer by regulation of androgen receptor transcriptional activity. Cancer Cell. 2013;23:332–346. doi: 10.1016/j.ccr.2013.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Khurana A., Nakayama K., Williams S., Davis R.J., Mustelin T., Ronai Z.e. Regulation of the ring finger E3 ligase Siah2 by p38 MAPK. J. Biol. Chem. 2006;281:35316–35326. doi: 10.1074/jbc.M606568200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Matsuura K., Huang N.J., Cocce K., Zhang L., Kornbluth S. Downregulation of the proapoptotic protein MOAP-1 by the UBR5 ubiquitin ligase and its role in ovarian cancer resistance to cisplatin. Oncogene. 2017;36:1698–1706. doi: 10.1038/onc.2016.336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Schiewek J., Schumacher U., Lange T., Joosse S.A., Wikman H., Pantel K., Mikhaylova M., Kneussel M., Linder S., Schmalfeldt B., Oliveira-Ferrer L., Windhorst S. Clinical relevance of cytoskeleton associated proteins for ovarian cancer. J. Cancer Res. Clin. Oncol. 2018;144:2195–2205. doi: 10.1007/s00432-018-2710-9. [DOI] [PubMed] [Google Scholar]
- 91.Bijnsdorp I.V., Hodzic J., Lagerweij T., Westerman B., Krijgsman O., Broeke J., Verweij F., Nilsson R.J., Rozendaal L., van Beusechem V.W., van Moorselaar J.A., Wurdinger T., Geldof A.A. miR-129-3p controls centrosome number in metastatic prostate cancer cells by repressing CP110. Oncotarget. 2016;7:16676–16687. doi: 10.18632/oncotarget.7572. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Hu S., Danilov A.V., Godek K., Orr B., Tafe L.J., Rodriguez-Canales J., Behrens C., Mino B., Moran C.A., Memoli V.A., Mustachio L.M., Galimberti F., Ravi S., DeCastro A., Lu Y. CDK2 inhibition causes anaphase catastrophe in lung cancer through the centrosomal protein CP110. Cancer Res. 2015;75:2029–2038. doi: 10.1158/0008-5472.CAN-14-1494. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Colebatch A.J., Dobrovic A., Cooper W.A. TERT gene: Its function and dysregulation in cancer. J. Clin. Pathol. 2019;72:281. doi: 10.1136/jclinpath-2018-205653. [DOI] [PubMed] [Google Scholar]
- 94.Matsuo R., Ochiai W., Nakashima K., Taga T. A new expression cloning strategy for isolation of substrate-specific kinases by using phosphorylation site-specific antibody. J. Immunol. Methods. 2001;247:141–151. doi: 10.1016/s0022-1759(00)00313-6. [DOI] [PubMed] [Google Scholar]
- 95.Galoczova M., Coates P., Vojtesek B. STAT3, stem cells, cancer stem cells and p63. Cell Mol. Biol. Lett. 2018;23:12. doi: 10.1186/s11658-018-0078-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Yeh Y.T., Ou-Yang F., Chen I.F., Yang S.F., Wang Y.Y., Chuang H.Y., Su J.H., Hou M.F., Yuan S.S. STAT3 ser727 phosphorylation and its association with negative estrogen receptor status in breast infiltrating ductal carcinoma. Int. J. Cancer. 2006;118:2943–2947. doi: 10.1002/ijc.21771. [DOI] [PubMed] [Google Scholar]
- 97.An T., Li S., Pan W., Tien P., Zhong B., Shu H.-B., Wu S. DYRK2 negatively regulates type I interferon induction by promoting TBK1 degradation via Ser527 phosphorylation. PLoS Pathog. 2015;11 doi: 10.1371/journal.ppat.1005179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Katlinski K.V., Gui J., Katlinskaya Y.V., Ortiz A., Chakraborty R., Bhattacharya S., Carbone C.J., Beiting D.P., Girondo M.A., Peck A.R., Puré E., Chatterji P., Rustgi A.K., Diehl J.A., Koumenis C. Inactivation of interferon receptor promotes the establishment of immune privileged tumor microenvironment. Cancer Cell. 2017;31:194–207. doi: 10.1016/j.ccell.2017.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Zitvogel L., Galluzzi L., Kepp O., Smyth M.J., Kroemer G. Type I interferons in anticancer immunity. Nat. Rev. Immunol. 2015;15:405–414. doi: 10.1038/nri3845. [DOI] [PubMed] [Google Scholar]
- 100.Dotto G.P. Notch tumor suppressor function. Oncogene. 2008;27:5115–5123. doi: 10.1038/onc.2008.225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Lee H.-J., Kim M.-Y., Park H.-S. Phosphorylation-dependent regulation of Notch1 signaling: The fulcrum of Notch1 signaling. BMB Rep. 2015;48:431–437. doi: 10.5483/BMBRep.2015.48.8.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Mimoto R., Imawari Y., Hirooka S., Takeyama H., Yoshida K. Impairment of DYRK2 augments stem-like traits by promoting KLF4 expression in breast cancer. Oncogene. 2017;36:1862–1872. doi: 10.1038/onc.2016.349. [DOI] [PubMed] [Google Scholar]
- 103.Kumamoto T., Yamada K., Yoshida S., Aoki K., Hirooka S., Eto K., Yanaga K., Yoshida K. Impairment of DYRK2 by DNMT1-mediated transcription augments carcinogenesis in human colorectal cancer. Int. J. Oncol. 2020;56:1529–1539. doi: 10.3892/ijo.2020.5020. [DOI] [PubMed] [Google Scholar]
- 104.Lan X., Wu N., Wu L., Qu K., Osoro E.K., Guan D., Du X., Wang B., Chen S., Miao J., Ren J., Liu L., Li H., Ning Q., Li D. The human novel gene LNC-HC inhibits hepatocellular carcinoma cell proliferation by sequestering hsa-miR-183-5p. Mol. Ther. Nucleic Acids. 2020;20:468–479. doi: 10.1016/j.omtn.2020.03.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Wu X., Zahari M.S., Ma B., Liu R., Renuse S., Sahasrabuddhe N.A., Chen L., Chaerkady R., Kim M.S., Zhong J., Jelinek C., Barbhuiya M.A., Leal-Rojas P., Yang Y., Kashyap M.K. Global phosphotyrosine survey in triple-negative breast cancer reveals activation of multiple tyrosine kinase signaling pathways. Oncotarget. 2015;6:29143–29160. doi: 10.18632/oncotarget.5020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Kumamoto T., Yamada K., Yoshida S., Aoki K., Yanaga K., Yoshida K. DNA methylation of dual-specificity tyrosine-regulated kinase 2 (DYRK2) promoter regulates proliferation of human colorectal cancer. J. Am. Coll. Surgeons. 2019;229 E91. [Google Scholar]
- 107.Sakamoto K., Schmidt J.W., Wagner K.-U. Mouse models of breast cancer. Methods Mol. Biol. 2015;1267:47–71. doi: 10.1007/978-1-4939-2297-0_3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Kwon M.-c., Berns A. Mouse models for lung cancer. Mol. Oncol. 2013;7:165–177. doi: 10.1016/j.molonc.2013.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Järvinen A.K., Autio R., Haapa-Paananen S., Wolf M., Saarela M., Grénman R., Leivo I., Kallioniemi O., Mäkitie A.A., Monni O. Identification of target genes in laryngeal squamous cell carcinoma by high-resolution copy number and gene expression microarray analyses. Oncogene. 2006;25:6997–7008. doi: 10.1038/sj.onc.1209690. [DOI] [PubMed] [Google Scholar]
- 110.Zou C., Lyu Y., Jiang J., Cao Y., Wang M., Sang C., Zhang R., Li H., Liew C.-C., Cheng C., Zhao S. Use of peripheral blood transcriptomic biomarkers to distinguish high-grade cervical squamous intraepithelial lesions from low-grade lesions. Oncol. Lett. 2020;20:2280–2290. doi: 10.3892/ol.2020.11779. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Fernández-Martínez P., Zahonero C., Sánchez-Gómez P. DYRK1A: The double-edged kinase as a protagonist in cell growth and tumorigenesis. Mol. Cell Oncol. 2015;2 doi: 10.4161/23723548.2014.970048. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Laham A.J., Saber-Ayad M., El-Awady R. DYRK1A: A down syndrome-related dual protein kinase with a versatile role in tumorigenesis. Cell Mol. Life Sci. 2020 doi: 10.1007/s00018-020-03626-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Cohen P. Protein kinases — the major drug targets of the twenty-first century? Nat. Rev. Drug Discov. 2002;1:309–315. doi: 10.1038/nrd773. [DOI] [PubMed] [Google Scholar]
- 114.Cohen P., Alessi D.R. Kinase drug discovery--what's next in the field? ACS Chem. Biol. 2013;8:96–104. doi: 10.1021/cb300610s. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Falke H., Chaikuad A., Becker A., Loaëc N., Lozach O., Abu Jhaisha S., Becker W., Jones P.G., Preu L., Baumann K., Knapp S., Meijer L., Kunick C. 10-Iodo-11H-indolo[3,2-c]quinoline-6-carboxylic acids are selective inhibitors of DYRK1A. J. Med. Chem. 2015;58:3131–3143. doi: 10.1021/jm501994d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Bellmaine S.F., Ovchinnikov D.A., Manallack D.T., Cuddy C.E., Elefanty A.G., Stanley E.G., Wolvetang E.J., Williams S.J., Pera M. Inhibition of DYRK1A disrupts neural lineage specificationin human pluripotent stem cells. Elife. 2017;6 doi: 10.7554/eLife.24502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Sonamoto R., Kii I., Koike Y., Sumida Y., Kato-Sumida T., Okuno Y., Hosoya T., Hagiwara M. Identification of a DYRK1A inhibitor that induces degradation of the target kinase using Co-chaperone CDC37 fused with luciferase nanoKAZ. Sci. Rep. 2015;5:12728. doi: 10.1038/srep12728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Velazquez R., Meechoovet B., Ow A., Foley C., Shaw A., Smith B., Oddo S., Hulme C., Dunckley T. Chronic Dyrk1 inhibition delays the onset of AD-like pathology in 3xTg-AD mice. Mol. Neurobiol. 2019;56:8364–8375. doi: 10.1007/s12035-019-01684-9. [DOI] [PubMed] [Google Scholar]
- 119.Cozza G., Sarno S., Ruzzene M., Girardi C., Orzeszko A., Kazimierczuk Z., Zagotto G., Bonaiuto E., Di Paolo M.L., Pinna L.A. Exploiting the repertoire of CK2 inhibitors to target DYRK and PIM kinases. Biochim. Biophys. Acta. 2013;1834:1402–1409. doi: 10.1016/j.bbapap.2013.01.018. [DOI] [PubMed] [Google Scholar]
- 120.Pagano M.A., Bain J., Kazimierczuk Z., Sarno S., Ruzzene M., Di Maira G., Elliott M., Orzeszko A., Cozza G., Meggio F., Pinna L.A. The selectivity of inhibitors of protein kinase CK2: An update. Biochem. J. 2008;415:353–365. doi: 10.1042/BJ20080309. [DOI] [PubMed] [Google Scholar]
- 121.Siddiqui-Jain A., Drygin D., Streiner N., Chua P., Pierre F., O'Brien S.E., Bliesath J., Omori M., Huser N., Ho C., Proffitt C., Schwaebe M.K., Ryckman D.M., Rice W.G., Anderes K. CX-4945, an orally bioavailable selective inhibitor of protein kinase CK2, inhibits prosurvival and angiogenic signaling and exhibits antitumor efficacy. Cancer Res. 2010;70:10288–10298. doi: 10.1158/0008-5472.CAN-10-1893. [DOI] [PubMed] [Google Scholar]
- 122.Kim H., Lee K.S., Kim A.K., Choi M., Choi K., Kang M., Chi S.W., Lee M.S., Lee J.S., Lee S.Y., Song W.J., Yu K., Cho S. A chemical with proven clinical safety rescues Down-syndrome-related phenotypes in through DYRK1A inhibition. Dis. Model Mech. 2016;9:839–848. doi: 10.1242/dmm.025668. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Melchior B., Mittapalli G.K., Lai C., Duong-Polk K., Stewart J., Guner B., Hofilena B., Tjitro A., Anderson S.D., Herman D.S., Dellamary L., Swearingen C.J., Sunil K.C., Yazici Y. Tau pathology reduction with SM07883, a novel, potent, and selective oral DYRK1A inhibitor: A potential therapeutic for Alzheimer's disease. Aging Cell. 2019;18 doi: 10.1111/acel.13000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Cuny G.D., Robin M., Ulyanova N.P., Patnaik D., Pique V., Casano G., Liu J.F., Lin X., Xian J., Glicksman M.A., Stein R.L., Higgins J.M. Structure-activity relationship study of acridine analogs as haspin and DYRK2 kinase inhibitors. Bioorg. Med. Chem. Lett. 2010;20:3491–3494. [PMC free article] [PubMed] [Google Scholar]
- 125.Cuny G.D., Ulyanova N.P., Patnaik D., Liu J.F., Lin X., Auerbach K., Ray S.S., Xian J., Glicksman M.A., Stein R.L., Higgins J.M. Structure-activity relationship study of beta-carboline derivatives as haspin kinase inhibitors. Bioorg. Med. Chem. Lett. 2012;22:2015–2019. doi: 10.1016/j.bmcl.2012.01.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Nelson K.M., Dahlin J.L., Bisson J., Graham J., Pauli G.F., Walters M.A. The essential medicinal chemistry of curcumin. J. Med. Chem. 2017;60:1620–1637. doi: 10.1021/acs.jmedchem.6b00975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.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. doi: 10.1042/BJ20070797. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Banerjee S., Buhrlage S.J., Huang H.-T., Deng X., Zhou W., Wang J., Traynor R., Prescott A.R., Alessi D.R., Gray N.S. Characterization of WZ4003 and HTH-01-015 as selective inhibitors of the LKB1-tumour-suppressor-activated NUAK kinases. Biochem. J. 2014;457:215–225. doi: 10.1042/BJ20131152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Imawari Y., Mimoto R., Hirooka S., Morikawa T., Takeyama H., Yoshida K. Downregulation of dual-specificity tyrosine-regulated kinase 2 promotes tumor cell proliferation and invasion by enhancing cyclin-dependent kinase 14 expression in breast cancer. Cancer Sci. 2018;109:363–372. doi: 10.1111/cas.13459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Yokoyama-Mashima S., Yogosawa S., Kanegae Y., Hirooka S., Yoshida S., Horiuchi T., Ohashi T., Yanaga K., Saruta M., Oikawa T., Yoshida K. Forced expression of DYRK2 exerts anti-tumor effects via apoptotic induction in liver cancer. Cancer Lett. 2019;451:100–109. doi: 10.1016/j.canlet.2019.02.046. [DOI] [PubMed] [Google Scholar]
- 131.Erickson-Miller C.L., Creasy C., Chadderton A., Hopson C.B., Valoret E.I., Gorczyca M., Elefante L., Wojchowski D.M., Chomo M., Fitch D.M., Duffy K.J. GSK626616: A DYRK3 inhibitor as a potential new therapy for the treatment of anemia. Blood. 2007;110 [Google Scholar]
- 132.Debdab M., Carreaux F., Renault S., Soundararajan M., Fedorov O., Filippakopoulos P., Lozach O., Babault L., Tahtouh T., Baratte B., Ogawa Y., Hagiwara M., Eisenreich A., Rauch U., Knapp S. Leucettines, a class of potent inhibitors of cdc2-like kinases and dual specificity, tyrosine phosphorylation regulated kinases derived from the marine sponge leucettamine B: Modulation of alternative pre-RNA splicing. J. Med. Chem. 2011;54:4172–4186. doi: 10.1021/jm200274d. [DOI] [PubMed] [Google Scholar]
- 133.Coutadeur S., Benyamine H., Delalonde L., de Oliveira C., Leblond B., Foucourt A., Besson T., Casagrande A.-S., Taverne T., Girard A., Pando M.P., Désiré L. A novel DYRK1A (Dual specificity tyrosine phosphorylation-regulated kinase 1A) inhibitor for the treatment of Alzheimer's disease: Effect on Tau and amyloid pathologies in vitro. J. Neurochem. 2015;133:440–451. doi: 10.1111/jnc.13018. [DOI] [PubMed] [Google Scholar]
- 134.Chaikuad A., Diharce J., Schröder M., Foucourt A., Leblond B., Casagrande A.-S., Désiré L., Bonnet P., Knapp S., Besson T. An unusual binding model of the methyl 9-Anilinothiazolo[5,4-f] quinazoline-2-carbimidates (EHT 1610 and EHT 5372) confers high selectivity for dual-specificity tyrosine phosphorylation-regulated kinases. J. Med. Chem. 2016;59:10315–10321. doi: 10.1021/acs.jmedchem.6b01083. [DOI] [PubMed] [Google Scholar]
- 135.Ogawa Y., Nonaka Y., Goto T., Ohnishi E., Hiramatsu T., Kii I., Yoshida M., Ikura T., Onogi H., Shibuya H., Hosoya T., Ito N., Hagiwara M. Development of a novel selective inhibitor of the Down syndrome-related kinase Dyrk1A. Nat. Commun. 2010;1:86. doi: 10.1038/ncomms1090. [DOI] [PubMed] [Google Scholar]
- 136.Kii I., Sumida Y., Goto T., Sonamoto R., Okuno Y., Yoshida S., Kato-Sumida T., Koike Y., Abe M., Nonaka Y., Ikura T., Ito N., Shibuya H., Hosoya T., Hagiwara M. Selective inhibition of the kinase DYRK1A by targeting its folding process. Nat. Commun. 2016;7:11391. doi: 10.1038/ncomms11391. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Neumann F., Gourdain S., Albac C., Dekker A.D., Bui L.C., Dairou J., Schmitz-Afonso I., Hue N., Rodrigues-Lima F., Delabar J.M., Potier M.C., Le Caër J.P., Touboul D., Delatour B., Cariou K. DYRK1A inhibition and cognitive rescue in a Down syndrome mouse model are induced by new fluoro-DANDY derivatives. Sci. Rep. 2018;8:2859. doi: 10.1038/s41598-018-20984-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Hutterer C., Milbradt J., Hamilton S., Zaja M., Leban J., Henry C., Vitt D., Steingruber M., Sonntag E., Zeitträger I., Bahsi H., Stamminger T., Rawlinson W., Strobl S., Marschall M. Inhibitors of dual-specificity tyrosine phosphorylation-regulated kinases (DYRK) exert a strong anti-herpesviral activity. Antiviral Res. 2017;143:113–121. doi: 10.1016/j.antiviral.2017.04.003. [DOI] [PubMed] [Google Scholar]
- 139.Gelbert L.M., Cai S., Lin X., Sanchez-Martinez C., Del Prado M., Lallena M.J., Torres R., Ajamie R.T., Wishart G.N., Flack R.S., Neubauer B.L., Young J., Chan E.M., Iversen P., Cronier D. Preclinical characterization of the CDK4/6 inhibitor LY2835219: In-vivo cell cycle-dependent/independent anti-tumor activities alone/in combination with gemcitabine. Invest. New Drugs. 2014;32:825–837. doi: 10.1007/s10637-014-0120-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Shi Y., Bray W., Smith A.J., Zhou W., Calaoagan J., Lagisetti C., Sambucetti L., Crews P., Lokey R.S., Webb T.R. An exon skipping screen identifies antitumor drugs that are potent modulators of pre-mRNA splicing, suggesting new therapeutic applications. PLoS One. 2020;15 doi: 10.1371/journal.pone.0233672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Loaëc N., Attanasio E., Villiers B., Durieu E., Tahtouh T., Cam M., Davis R.A., Alencar A., Roué M., Bourguet-Kondracki M.L., Proksch P., Limanton E., Guiheneuf S., Carreaux F., Bazureau J.P. Marine-derived 2-aminoimidazolone alkaloids. Leucettamine B-related polyandrocarpamines inhibit mammalian and Protozoan DYRK & CLK kinases. Mar. Drugs. 2017;15 doi: 10.3390/md15100316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Myrianthopoulos V., Kritsanida M., Gaboriaud-Kolar N., Magiatis P., Ferandin Y., Durieu E., Lozach O., Cappel D., Soundararajan M., Filippakopoulos P., Sherman W., Knapp S., Meijer L., Mikros E., Skaltsounis A.-L. Novel inverse binding mode of indirubin derivatives yields improved selectivity for DYRK kinases. ACS Med. Chem. Lett. 2013;4:22–26. doi: 10.1021/ml300207a. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Zagorska A., Deak M., Campbell D.G., Banerjee S., Hirano M., Aizawa S., Prescott A.R., Alessi D.R. New roles for the LKB1-NUAK pathway in controlling myosin phosphatase complexes and cell adhesion. Sci. Signaling. 2010;3 doi: 10.1126/scisignal.2000616. [DOI] [PubMed] [Google Scholar]
- 144.Rüben K., Wurzlbauer A., Walte A., Sippl W., Bracher F., Becker W. Selectivity profiling and biological activity of novel β-carbolines as potent and selective DYRK1 kinase inhibitors. PLoS One. 2015;10 doi: 10.1371/journal.pone.0132453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Liu Y.A., Jin Q., Zou Y., Ding Q., Yan S., Wang Z., Hao X., Nguyen B., Zhang X., Pan J., Mo T., Jacobsen K., Lam T., Wu T.Y.H., Petrassi H.M. Selective DYRK1A inhibitor for the treatment of type 1 diabetes: Discovery of 6-azaindole derivative GNF2133. J. Med. Chem. 2020;63:2958–2973. doi: 10.1021/acs.jmedchem.9b01624. [DOI] [PubMed] [Google Scholar]