Next Article in Journal
Genetics and Epigenetics in Asthma
Previous Article in Journal
Applications of Mesenchymal Stem Cells in Skin Regeneration and Rejuvenation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in the Development of Eukaryotic Elongation Factor 2 Kinase (eEF2K) Natural Product and Synthetic Small Molecule Inhibitors for Cancer Chemotherapy

1
Li Dak Sum Yip Yio Chin Kenneth Li Marine Biopharmaceutical Research Center, College of Food and Pharmaceutical Sciences, Ningbo University, Ningbo 315800, China
2
Institute of Drug Discovery Technology, Ningbo University, Ningbo 315211, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(5), 2408; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052408
Submission received: 12 February 2021 / Revised: 24 February 2021 / Accepted: 24 February 2021 / Published: 27 February 2021

Abstract

:
Eukaryotic elongation factor 2 kinase (eEF2K or Ca2+/calmodulin-dependent protein kinase, CAMKIII) is a new member of an atypical α-kinase family different from conventional protein kinases that is now considered as a potential target for the treatment of cancer. This protein regulates the phosphorylation of eukaryotic elongation factor 2 (eEF2) to restrain activity and inhibit the elongation stage of protein synthesis. Mounting evidence shows that eEF2K regulates the cell cycle, autophagy, apoptosis, angiogenesis, invasion, and metastasis in several types of cancers. The expression of eEF2K promotes survival of cancer cells, and the level of this protein is increased in many cancer cells to adapt them to the microenvironment conditions including hypoxia, nutrient depletion, and acidosis. The physiological function of eEF2K and its role in the development and progression of cancer are here reviewed in detail. In addition, a summary of progress for in vitro eEF2K inhibitors from anti-cancer drug discovery research in recent years, along with their structure–activity relationships (SARs) and synthetic routes or natural sources, is also described. Special attention is given to those inhibitors that have been already validated in vivo, with the overall aim to provide reference context for the further development of new first-in-class anti-cancer drugs that target eEF2K.

1. Introduction

Targeted therapy is an important strategy for cancer treatment, and this has been well applied in actual clinical applications [1]. At present, most clinically used targeted cancer drugs are inhibitors of tyrosine kinases [1,2]. However, even when these drugs have exceptional efficacy initially, the later emergence of drug resistance limits their usefulness [2]. Finding new drug targets and developing new targeted anticancer agents have accordingly become important aspects of drug discovery and development.
Eukaryotic elongation factor 2 kinase (eEF2K) is the first α-kinase to have been discovered [3]. The activity of eEF2K depends on calcium ions and calmodulin (CaM) [4]. Therefore, eEF2K has also been called CaM-dependent protein kinase III or CAMKIII, and it appears this way in some literature reports [5]. The “α-kinase” is an atypical protein kinase family, and while these often have similar ATP binding pockets to typical protein kinase family members, they are differentiated by having a translocated conserved region with an alternate sequence, GXGXXG [6]. Because of this, drugs that target eEF2K are less likely to modulate typical kinases and should ideally not have or create cross-resistance with traditional kinase drugs. Increasing evidence shows that eEF2K is highly expressed in a variety of tumor tissues and that it is related to the development and prognosis of several kinds of malignancies such as breast cancer, ovarian cancer, colon cancer, glioma, medulloblastoma, hepatocellular carcinoma, and prostate cancer [7,8,9,10,11,12]. In addition, eEF2K can also participate in the regulation of the tumor cell cycle, proliferation, autophagy, apoptosis, angiogenesis, invasion, and metastasis, among other processes [13,14,15]. For all of these reasons, eEF2K is a potential therapeutic target for anticancer drug development.
Recently, with the deepening research on the function of eEF2K, increased attention has turned to the development of eEF2K inhibitors. However, relatively fewer reports have disclosed the development of eEF2K inhibitors compared with studies on the physiological functions and pathological effects of eEF2K. While the molecules summarized in this review have been reported as eEF2K inhibitors, the chemical structures are very different, and no predominant pharmacophore or active scaffold has emerged. Herein, the physiological functions of eEF2K are introduced and its various regulatory effects on tumors are discussed. Furthermore, a summary of recent research progress on eEF2K inhibitors for cancer chemotherapy is presented along with the detailed structure–activity relationships (SARs) and synthetic routes or natural sources of existing eEF2K inhibitors.

2. Physiological Function of eEF2K

Cellular and organismal survival requires the continuous maintenance of protein synthesis, categorizing proteins as essential for function. During protein synthesis, the control of elongation regulates mRNA translation and adapts cells to changes in nutrients, energy, and oxygen. The protein eukaryotic elongation factor 2 (eEF2) plays a key role in protein elongation and is the only substrate of eEF2K known to date [16]. As a member of the GTP-binding translation elongation factor family, eEF2 promotes GTP-dependent ribosome translocation [17]. During elongation, eEF2 promotes the movement of peptidyl-tRNAs from the A site to the P site of the ribosome. In eukaryotic cells, eEF2 can be completely inactivated via phosphorylation by eEF2K [18]. Currently, eEF2K is the only α-kinase with activity dependent on Ca2+ ions [19]. In vivo, Ca2+ ions bind to calmodulin with high affinity and activate the eEF2K kinase domain, and this rapidly triggers the autophosphorylation of eEF2K at Thr384, leading to kinase conformational changes that form a binding pocket [13]. Active eEF2K will phosphorylate eEF2 at Thr56, thereby inhibiting the activity of eEF2 and ultimately preventing the extension of polypeptide chains in the process of protein synthesis [20]. What is known about the structure of eEF2K and the specific regulation of the extension process has been reviewed in detail by Proud [5], as well as more recently by Karakas and Ozpolat [13], and will not be elaborated here.

3. The Role of eEF2K in Cancer

It has been shown that eEF2K is overexpressed and regulates tumor progression in several types of malignancies, including breast cancer, glioma cancer, pancreatic cancer, lung cancer, neuroblastoma, and colorectal cancer [8,21,22]. Previous studies have found that eEF2K is associated with tumor proliferation and survival, tumorigenesis, invasion, drug resistance, and poor prognosis [23,24]. For instance, microRNA 603 inhibits tumor formation in triple-negative breast cancer by the targeted inhibition of eEF2K [25]. It was also reported that eEF2K promotes the proliferation of ovarian cancer cells and that its expression is positively correlated with poor prognosis [24]. In hepatocellular carcinoma, eEF2K promotes angiogenesis through PI3K/Akt and STAT3 signaling [11]. Similarly, eEF2K is positively correlated with lung cancer proliferation, invasion and metastasis, and poor prognosis [26]. These early data indicate the importance of eEF2K in cancers, and suggest that it is a potential new target for cancer chemotherapeutic treatments.

3.1. eEF2K Helps Tumor Cells to Cope with Harsh Environments

The rapid proliferation of cancer cells in tumors requires a high amount of energy, in part due to greatly upregulated protein synthesis. The process leads to tumors creating their own harsh microenvironments that have, for example, reduced availability of nutrients, low pH, and insufficient oxygen. Further adaptation of the tumor cells, including by increasing expression of eEF2K, helps to regulate the synthesis of proteins in the harsh environment and be protected for continued proliferation. Under conditions of nutritional deprivation, tumor cells with high eEF2K expression can continue to survive, while those with low eEF2K expression have been shown to die [27]. When intracellular nutrition is insufficient, the content of ATP decreases, and AMP or ADP increases to activate AMP-activated protein kinase (AMPK). After activation, AMPK can induce the phosphorylation of eEF2K at Ser398 or Ser491 to thereby inhibit the function of eEF2 (Figure 1). This ultimately reduces the rate and energy consumption of intracellular proliferation and protein synthesis and promotes energy production processes such as glucose metabolism and fatty acid oxidation. The mammalian target of rapamycin complex 1 (mTORC1) is another energy-related protein that has been found to downregulate eEF2K [28]. Activated mTORC1 inhibits the activation of eEF2K by inducing its phosphorylation at a variety of residues, including Ser70, Ser78, Ser359, Ser366, Ser392, Ser396, and Ser470 [29,30]. Regarding its regulation, mTORC1 is stimulated by amino acids, hormones, growth factors, and cellular nutrients [4,28]. Under nutritional deprivation, the activity of mTORC1 is inhibited, partially alleviating its phosphorylation and inhibitory effect on eEF2K. At the same time, activated AMPK has been shown to further indirectly inhibit the activity of mTORC1 [31]. In addition, mTORC1-mediated inhibition of eEF2K is essential for proliferation of adenomatous polyposis coli (APC)-deficient cells. Rapamycin targets eEF2 indirectly through the mTORC1-S6K-eEF2K pathway, and treatment of APC-deficient adenomas with rapamycin induces tumor cell growth arrest and differentiation [32].
The rapid proliferation of tumor cells also requires a large amount of oxygen, which frequently is overdrawn enough to induce a hypoxic environment. It has been found that hypoxia inhibits protein synthesis in breast cancer cells in part through 4E-BP1 and the eEF2K pathway, controlled by mTOR [33]. Additionally, eEF2K is activated and induces eEF2 phosphorylation during hypoxia independent of AMPK and mTORC1 signaling [34]. The eEF2K residue Pro98 is a generally conserved linker between the calmodulin binding domain and the catalytic domain, and when it is hydroxylated, the binding of calmodulin to eEF2K is reduced and the activity of eEF2K is significantly limited. Under normoxia, proline hydroxylase catalyzes the hydroxylation of Pro98 of eEF2K, thus inhibiting eEF2K activity. However, when the cells are hypoxic, the activity of proline hydroxylase is inhibited, thereby releasing the normal inhibition of eEF2K (Figure 1) [34]. Normal cells rely on mitochondrial oxidative phosphorylation to produce energy, while tumor cells mainly generate energy through glycolysis under hypoxic/normoxic conditions (Warburg effect) [35]. The upregulation of eEF2K accelerates glycolysis to promote human breast cancer cells in development and progression. eEF2K inhibits protein phosphatase 2A-A (PP2A-A) synthesis, thereby interfering with its promotion of c-Myc ubiquitin-proteasome degradation, and finally activates the transcription of pyruvate kinase M2 subtype (PKM2) to promote glycolysis [36].
Low pH is frequently a major feature of tumor microenvironments. In tumor cells, unrestricted glycolysis leads to a large accumulation of lactic acid, thereby acidifying the local environment [37,38]. Under acidic conditions, or low pH, overexpression of eEF2K inhibits protein synthesis. At neutral pH, by contrast, overexpression of eEF2K does not affect protein synthesis, indicating the activation of eEF2K in acidic conditions (Figure 1) [39]. However, the activation of eEF2K is independent of the activity of mTORC1. It is understood that the affinity of eEF2K to CaM is enhanced at acidic pH, and the histidine residue (H108) in CaM is essential for the activation of eEF2K [40].
Overall, the activity of eEF2K is known to change with the conditions of the tumor microenvironment (e.g., energy deficiency, hypoxia, low pH), thereby regulating the process of tumor protein synthesis and ultimately protecting the survival of tumor cells in otherwise harsh conditions.

3.2. eEF2K Inhibits Cell Apoptosis

Cell apoptosis is typically dysregulated in cancers, which leads to rampant proliferation. Accordingly, inducing apoptosis in cancer cells is an important mechanism of anti-tumor drugs. The expression of eEF2K can inhibit apoptosis and promote cancer cell survival [4]. The suppression of this mechanism would enable eEF2K to be used mechanistically as a new drug target for single or multi-agent cancer chemotherapeutics. The caspases represent an important family of proteins that regulate cell apoptosis. Among them, caspase 8 and caspase 9 are the key regulatory proteins in extrinsic and intrinsic apoptotic pathways, respectively [41]. The cleavage of these caspases will eventually lead to the cleavage of caspase 3 and ultimately apoptosis of the cell [42]. Some eEF2K inhibitors have been shown to induce tumor cell apoptosis by these mediated extrinsic and/or intrinsic apoptosis pathways. The cleavage of caspase 8 that is induced by tumor necrosis factor (TNF) family proteins is an important aspect of the extrinsic apoptosis pathway [43]. The TNF-related apoptosis-inducing ligand (TRAIL) belongs to the TNF family. TRAIL can bind to the death receptors DR4 and DR5 to form the death-inducing signaling complex (DISC) and to upregulate Fas-associated protein with death domain (FADD), thereby inducing caspase-8-dependent apoptosis (Figure 2) [44]. Treatment of glioma cells with the eEF2K inhibitor, NH125 (1), showed the enhancement of TRAIL-induced apoptosis, and, with the increase of dosed NH125, the cleaved PARP and caspase 8 levels increased significantly [45]. Bcl-2 is another important family of proteins that regulate endogenous apoptosis [46]. The founding member protein Bcl-2 and Bcl-xL are anti-apoptotic proteins in the Bcl-2 family, and NH125 down-regulates the expression of Bcl-xL in glioma cells [45]. It was additionally shown that silencing eEF2K induces caspase-9 cleavage and Bcl-2 downregulation in breast cancer cells (Figure 2) [47]. Meanwhile, inhibiting eEF2K enhances the effect of doxorubicin in an orthotopic model of breast cancer [47]. Furthermore, eEF2K is highly expressed in pancreatic cancer (PaCa) and acts to inhibit apoptosis [14]. Treatment of PaCa cells with the natural product inhibitor of eEF2K, rottlerin (29), not only induces the collapse of mitochondrial potential causing intrinsic apoptosis, but also causes extrinsic apoptosis regulated by TRAIL and caspase 8 [14]. At the same time, rottlerin also effects the expression of TG2 (PKC-δ/tissue transglutaminase), which in turn activates apoptosis-inducing factor (AIF), and ultimately causes caspase-dependent apoptosis (Figure 2) [14].

3.3. eEF2K Regulates the Cell Cycle

The cell cycle is inextricably tied to protein synthesis, and thus the impact of eEF2K in elongation can be understandably expanded. For one thing, phosphorylation of eEF2K residues Ser359 and Ser366 leads to its inactivation, which accordingly regulates cell cycle progression [31]. The inhibition of eEF2K has been shown to arrest breast cancer cells at the G0/G1-S phase [48]. Conversely, eEF2K is inactivated during G1 cell growth, so eEF2 becomes active and promotes protein synthesis [48]. The Ser366 residue of eEF2K is a point of regulation by MKK and mTOR pathways, and when cells enter the G1 phase, Ser366 is phosphorylated rapidly in an MKK-dependent manner (Figure 3) [49]. During S phase DNA replication, eEF2K is slowly dephosphorylated, and eEF2 activity is eventually inhibited completely during G2 and mitosis (Figure 3) [48]. It is well-known that eEF2K is a calcium/calmodulin-dependent protein kinase, and calcium binding to CaM has an important influence on the process of mitosis [5,50]. In the G1/S transition, the intracellular Ca2+ concentration increases. Under environmental conditions of high calcium concentration, calmodulin can bind to eEF2K and activate it so that the cells enter the S phase (Figure 3) [51]. Another function of cellular calcium is to upregulate cAMP levels [52]. The cAMP in turn activates PKA, which can activate eEF2K by phosphorylation of the Ser500 residue [52,53]. The regulation of eEF2K on the G2/M phase is related to the phosphorylation at the Ser359 residue, and this leads to inhibition of eEF2K activity independent of Ca2+ concentration (Figure 3) [54]. It has been shown that eEF2K can also be regulated by cell cycle-related proteins to affect cell cycle progression. For example, human cyclin-dependent kinase 1 (CDC2) is regulated by mTORC1 and gets activated in the early stage of mitosis; it then inactivates eEF2K, and thus protein synthesis is carried out in mitotic cells (Figure 3) [55]. Cell cycle progression is closely related to proliferation in tumor cells, and chemical interference can accordingly inhibit proliferation and eventually lead to cell death [56]. The successful launch of the CDK4 and CDK6 targeting cell cycle inhibitor palbociclib, which was approved by the U.S. FDA in 2015, has encouraged further research on cell cycle-regulating anticancer agents for use as drugs [57]. Since eEF2K has a regulatory effect on multiple links in the cell cycle process, this represents an attractive new target for future cancer treatments.

3.4. eEF2K Regulates Cell Autophagy

Autophagy is a “self-consuming” program of cells that is used to remove damage and dysfunction or unnecessary proteins, and it is closely associated with human diseases such as cancer [58,59]. Under conditions of starvation or stress, excess cells may undergo autophagy so that the remaining cells can better survive [60,61]. Autophagy mechanistically works by lysosomes in cells degrading their own organelles and other macromolecules, and it is an important process for eukaryotes to carry out the turnover of intracellular substances [62,63]. As previously discussed, tumor cell proliferation often results in microenvironment nutritional starvation or stress conditions. Autophagy can protect cells from apoptosis and promotes tumor progression [62,64]. mTOR is an important regulator of autophagy and the upstream protein of eEF2K. Previous studies have demonstrated that eEF2K induces autophagy to protect cancer cells survival [65,66]. For example, it was found that during amino acid starvation and endoplasmic reticulum (ER) stress, eEF2K is activated to induce autophagy [67,68]. Inhibiting eEF2K-mediated autophagy has been reported to enhance the antitumor effect of MK-2206, an AKT inhibitor, on human nasopharyngeal carcinoma and human glioma cells [65,66]. The induction of autophagy is mediated via the TSC2/mTOR/S6 kinase/eEF-2 kinase pathway [65]. It was also found that silencing of eEF2K can inhibit autophagy through the mTORC1/p70S6K signaling pathway and increase the sensitivity of human glioma cells to 2-deoxy-d-glucose (2-DG) [69].
However, in contrast to the above outcome in human nasopharyngeal carcinoma and human glioma cells, it was found that silencing of eEF2K activity can induce autophagy to promote the proliferation of colon cancer cells but not enhance the anticancer effect of MK-2206 in human colon cancer cells [70]. The negative regulation of eEF2K on autophagy in colon cancer cells is dependent on the activation of the AMPK-ULK1 pathway [70]. Another experiment on human lung cancer cells showed that eEF2K protects cell survival under nutrient deprivation, but this effect is due to its inhibition of protein synthesis rather than regulation of autophagy [15]. Thus, the particular cancer cell type and specific mechanism of action being observed is important to consider.

3.5. eEF2K Promotes Tumor Angiogenesis, Metastasis, and Invasion

Angiogenesis, the growth of new blood vessels, provides tumors with more nutrients to promote their growth, and plays a key role in tumor proliferation, metastasis, and invasion [71,72,73]. Overexpression of eEF2K has been shown to promote angiogenesis, invasion, and metastasis in multiples types of tumors [74,75]. Inhibition of eEF2K expression can likewise prevent these tumor processes. For example, knockdown of eEF2K was found to prevent tumor progression and angiogenesis of hepatocellular carcinoma via the PI3K/Akt and STAT3 signaling pathway [11]. In triple-negative breast cancer (TNBC) cells, the proto-oncogene transcription factor forkhead box M1 (FOXM1) can regulate eEF2K and affect breast cancer cell migration and invasion, progression, and tumorigenesis [23]. The dual inhibitory effect of microRNA-34a on the FOXM1/eEF2K axis can regulate the growth and invasion of TNBC [76]. In addition, TNBC with mutations in PTEN and p53 is more sensitive to eEF2K inhibitors, and this effect is related to the AKT signaling pathway [8]. Similarly, it was found after knocking out eEF2K that the invasion and metastasis of lung cancer cells was inhibited [26]. Proud et al. found that this inhibitory effect may be related to integrin signaling proteins to control cell–cell/cell–extracellular matrix interactions and cell mobility [77].

4. Natural Product and Synthetic Small Molecule Inhibitors of eEF2K

Since eEF2K does not belong to a typical large family of kinases, and it is the only α-kinase that depends on Ca2+ and calmodulin, the structural dynamics of this protein have only recently been characterized [78]. Earlier studies, however, did find that blocking the function of eEF2K can effectively kill cancer cells without affecting normal cells [79]. Since the function of eEF2K is related to CaM, ATP, and eEF2, the corresponding binding sites are the key targets know for competitive inhibitors. Specific eEF2K inhibitors will ideally only affect the activity of the eEF2 protein to modulate elongation and cell cycle regulation, thus offering a new strategy for anticancer drug discovery and development. Many structurally distinct natural product and synthetic small molecule inhibitors of eEF2K have been reported to date (Table 1). However, due to the limited amount of existing research on eEF2K, no optimal pharmacophore or structure scaffold has been found for inhibiting this molecular target. Accordingly, the development of eEF2K inhibitors is still in the early preclinical stage. The previously reported eEF2K inhibitors are delineated according to whether they were discovered in single-target studies or as multi-target inhibitors, and each is described in the following sections.

4.1. Discovery and Development of Single Target eEF2K Inhibitors

4.1.1. NH125

NH125, or 1-benzyl-3-cetyl-2-methylimidazolium iodide (1, Figure 4), was first discovered to be an imidazolium histidine kinase inhibitor [80]. Later in an eEF2K enzyme activity test on 1 and a series of analogue molecules, this compound was found to significantly inhibit the enzyme activity of eEF2K in vitro (IC50 = 60 nM) [81]. The in vitro phenotypic effect of 1 was further tested, and it was found that this molecule inhibited proliferation of several different types of cancer cells [82]. Interestingly, in tumor cells with eEF2K at knocked down levels, compound 1 was still anti-proliferative. In-depth studies have found that the observed effects of 1 in vitro may correlate to induced phosphorylation of eEF2 [82]. Re-examination of the experiment of the inhibition of eEF2K enzyme activity by 1 in vitro led to the verification of this conclusion [83]. In addition, in vivo studies have shown that 1 combined with radiotherapy was more effective than 1 as a single agent or radiotherapy alone in delaying the growth of esophageal squamous cell carcinoma [73].

4.1.2. TX-1918

TX-1918, 2-((3,5-dimethyl-4-hydroxyphenyl)methylene)-4-cyclopentene-1,3-dione (2, Figure 4), was discovered as an eEF2K inhibitor with IC50 of 0.44 μM in vitro [84]. A series of 2-hydroxyarylidene-4-cyclopentene-1,3-diones was synthesized for testing as protein tyrosine kinase (PTK) inhibitors using a modified Knoevenagel reaction of m,m′-disubstituted p-hydroxybenzaldehydes (3, Scheme 1) with 4-cyclopentene-1,3-dione (4, Scheme 1) under acidic conditions. It was found that 2 can inhibit various tyrosine kinases, for instance, protein kinase A (PKA): IC50 = 44 μM; protein kinase C (PKC): IC50 = 44 μM; Src-K: IC50 = 4.4 μM; and EGFR-K: IC50 = 440 μM, but the necessary concentration is orders of magnitude higher than for eEF2K (IC50 = 0.44 μM) [84]. Cytotoxicity testing showed that TX-1918 inhibited the proliferation of HepG2 human liver cancer cells (IC50 = 2.7 μM) two orders of magnitude more potently than HCT116 human colon cancer cells (IC50 = 230 μM) [84]. This finding emphasizes the important role of the cancer cell type being studied and the related expression level of eEF2K when considering the testing and advancement of eEF2K inhibitors through preclinical development from proteins to cells, and animals. However, the 2-substituted 4-cyclopentene-1,3-dione moiety present in 2 and all analogues prepared by the synthesis shown in Scheme 1 should caution concern of reactivity or toxicity that likely precludes the further development of this series.
While the non-ATP-competitive compounds 1 and 2 have strong inhibitory effects on eEF2K in vitro, their demonstrated anti-proliferative effects at the cellular level are limited, and no in vivo validation has been reported to date. A much greater relative abundance of discovery and development research has focused on ATP-competitive inhibitors of eEF2K.

4.1.3. A-484954 and Its Derivatives

A-484954, systematically 7-amino-1-cyclopropyl-3-ethyl-2,4-dioxo-1,2,3,4-tetrahydropyrido[2,3-d]pyrimidine-6-carboxamide (5, Figure 5), is a reportedly highly selective eEF2K inhibitor with the in vitro IC50 value of 0.28 μM that was identified from a chemical library high-throughput screening (HTS) effort using a wide panel of serine/threonine and tyrosine kinases [82]. However, the effect that 5 has on tumor cell proliferation is not obvious until higher concentrations between 10–100 μM and an absence of serum in the bioassay [82,85]. This suggests that the compound may not reach the molecular target in the cell for any of a number of possible reasons. The further design and synthesis of a series of pyrido[2,3-d]pyrimidine-2,4-dione analogues related to 5 have been reported [86]. First, direct alkylation of uracil derivatives (6) with alkyl iodides in the presence of aqueous sodium hydroxide afforded the corresponding 1,3-disubstituted-6-aminouracils (7) [86]. Next, the hydrochloride (9) was prepared by the reaction of the Vilsmeier reagent (8) with intermediate 7, then treated with triethylamine and cyanoacetamide to obtain 5 and a series of analogues in 47–81% yield [86]. As shown in shown in Scheme 2, this synthesis allowed for the rapid generation of analogues but constrained the positions of differentiation available for evaluating a structure–activity relationship (SAR).
The SAR study of pyrido[2,3-d]pyrimidine-2,4-diones indicated that the main pyridine ring and the amido substituent at R2 were the key features of this class of molecules that had important influence on the improvement of eEF2K inhibitory activity. In addition, the molecules having the R1 group on the scaffold be ethyl (5, A-484954) or n-propyl (10, Figure 5, IC50 = 930 nM) had the most potent inhibitory activity [86]. However, when the R1 group was replaced by hydrogen, methyl, or benzyl derivatives, the activity was significantly decreased more than ten-fold [86]. Additionally, moderate loss of activity was observed after changing the cyclopropyl substituent at R3 group to ethyl group [86]. Interestingly, when R3 was a methyl group, the inhibitory activity was completely abolished [86]. Further study found that analogues 5 and 10 were able to inhibit AMPK-mediated activation of eEF2K [86]. The preliminary SAR study on this scaffold offers insights for the design of new eEF2K inhibitors. In 2020, Liang and coworkers incorporated the “PROteolysis TArget Chimeric” (PROTAC) strategy in combination with lead molecule 5 [87]. The PROTAC strategy involves the recruitment of E3 ubiquitin ligases with a common binding motif that can be tethered across a linker unit to a protein target ligand, here starting with 5 as a moiety designed to target eEF2K. Among a series of synthesized analogues with linkers of various chain length and constitution, compound 11 (“11l”, Figure 5) had the maximum eEF2K degradation rate (Dr) value of 56.7% and could induce apoptosis in human breast carcinoma MDA-MB-231 cells in vitro [87].

4.1.4. TS-2 and TS-4

Ishihara et al. described the synthesis of a series of novel 5,6-dihydro-4H-1,3-selenazine analogues that were prepared by the reaction of selenamides (12) with α,β-unsaturated ketones (13) under mild conditions in good yields (Scheme 3) [88]. Shortly thereafter, these compounds were screened for inhibitory activity in multiple protein kinases, including eEF2K, PKA, PKC, and protein tyrosine kinase (PTK), CaMK-I/II, and v-src [89]. Through this SAR study, 4-ethyl-4-hydroxy-2-p-tolyl-5,6-dihydro-4H-1,3-selenazine (TS-2, 14, Figure 6) and 4-hydroxy-6-isopropyl-4-methyl-2-p-tolyl-5,6-dihydro-4H-1,3-selenazine (TS-4, 15, Figure 6) were found to have the most selective inhibitory activity against eEF2K over other protein kinases [89]. The inhibition of eEF2K by 14 and 15 were quantified in a separate purified kinase in vitro, resulting in respective IC50 values of 0.36 and 0.31 μM [89]. Compound 14 was further studied and suggested to be an ATP-binding site inhibitor, since it competitively prevented ATP binding to eEF2K and non-competitively blocked eEF2 binding with eEF2K [89]. Further SAR data indicated that when the selenium atom in 14 and 15 was replaced with sulfur, bulky phenyl or styryl substituents were added meta to the toluyl group, or ring size was reduced to 1,3-selenazole, the inhibitory effect for all protein kinases studied was abolished [89]. Although most selenium-containing compounds are relatively unstable, the intramolecular conjugated system comprising the toluyl group and dihydro-1,3-selenazine ring appears to greatly improve stability in this series and offer increased value for these compounds as leads. However, due to these demands and the already explored SAR presented in Scheme 3, there may be limited avenues for further optimization of this series of compounds.

4.1.5. Thieno[2,3-b]pyridines

Thieno[2,3-b]pyridines were early considered as anticancer agents that target phospholipase C-γ (PLC-γ), and some of these inhibit the proliferation of several breast cancer cell types (MCF7, TamR7, SKBr3, MDA-MB-231/468, etc.) and arrest cells at the G2/M phase [90,91]. Two thieno[2,3-b]pyridine-containing heterocycles compounds (16 and 17, Figure 7) have been identified as ATP-competitive inhibitors of eEF2K via high-throughput screening, and their IC50 values were 0.22 and 2.5 μM, respectively [92]. In a follow-up study, it was determined that even small changes made to the structure of 16 significantly reduced the inhibitory activity against eEF2K, so another structural analogue (17) was selected for further optimization [93]. This led to the design and synthesis of a series of thieno[2,3-b]pyridine analogues 18 (Figure 7) to be tested as eEF2K inhibitors. The synthetic route (Scheme 4) is briefly described as follows: ketones 19 were reacted with aldehydes in the presence of KOH to generate α-β-unsaturated ketones, which were then cyclized with 2-cyanothioacetamide to obtain the 2-mercaptopyridine intermediates 20 that yielded thieno[2,3-b]pyridine analogues 18 after reaction with chloroacetonitrile followed by condensation with formamide [93]. These compounds were subsequently tested for eEF2K inhibition in vitro.
The SAR study of related compounds indicated that position R3 on 18 (the B-region indicated on 21) is open to modification with a wide range of alkyl and aryl substituents while maintaining or moderately enhancing the in vitro eEF2K inhibitory activity; but heterocyclic substitution was preferred [93]. In addition, a six-membered to 10-membered ring size study incorporating the R1 and R2 positions on 18 (the A-region indicated on 21) showed the impact on inhibitory activity. For example, incorporating the furan-2-yl group at R3, and a nine-membered ring including R1 and R2, compound 21 (Figure 7) yielded the best activity (IC50 = 0.17 μM), followed by the eight-membered (IC50 = 0.64 μM) and seven-membered (IC50 = 1.1 μM) ring analogues [93]. However, compounds containing six- and 10-membered rings lost activity (IC50 > 20 μM) [93]. Among these compounds and the 43 total analogues tested in the same study, 21 demonstrated the highest anti-proliferative activity with EC50 of 17 μM against human colon cancer HCT-116 cells [93].

4.1.6. β-Phenylalanines

In 2018, Liu et al. predicted by in silico high-throughput screening a simple β-phenylalanine derivative (22, Figure 8) that was later confirmed to have moderate in vitro inhibitory activity on eEF2K (IC50 = 35.1 μM) [94]. To evaluate the potential of β-phenylalanine derivatives on eEF2K, a series of 46 such compounds were designed and synthesized (23, Figure 8) for evaluation of inhibitory activity on eEF2K and MDA-MB-231/436 breast cancer cells in vitro [94]. The synthetic route to obtain the desired β-phenylalanine derivatives (23) is shown in Scheme 5. This uses affordable commercial benzaldehydes 24 and malonic acid as the starting materials to connect by the Knoevenagel reaction to generate substituted β-phenylalanines (25). These molecules are simply esterified to afford carboxyl-protected intermediates (26) for regiospecific acylation (27), and then hydrolysis of the protecting group to yield the desired products (23). Overall, this scheme allows for the efficient and affordable generation of a large library of molecules for further testing.
The SAR study on 46 synthesized β-phenylalanine derivatives and analogues indicated that compounds containing the sulfamide group as a linker to a substituted phenyl group with para electron-withdrawing groups, such as CN and CF3, showed more potent activity [94]. Introduction of o,p-dichloro substitution on the on the β-phenylalanine also increased activity, and 3-((4-cyanophenyl)sulfonamido)-3-(2,4-dichlorophenyl) propanoic acid (“21l”, 28, Figure 8) yielded the best eEF2k enzymatic activity of compounds tested in vitro (IC50 = 5.5 μM) [94]. Molecular docking and molecular dynamic simulations suggest that this molecule is a potential ATP-competing inhibitor. Compound 28 was further found to be weakly anti-proliferative in vitro against MDA-MB-231 cells and MDA-MB-436 cells with IC50 values of 12.6 and 19.8 μM, respectively, and most notably demonstrated in vivo efficacy inhibiting tumor growth by inducing apoptosis via eEF2K inhibition in the xenograft mouse model of TNBC with the same two cell lines [94]. The exciting validation of 28 in vivo is promising for the development of this as a lead compound for new anticancer drugs, as well as encouraging for studies of other new ATP-competitive eEF2K inhibitors.

4.1.7. Fluoxetine

Fluoxetine (29, Figure 9) is a selective serotonin reuptake inhibitor (SSRI) drug in widespread clinical use as an anti-depressant under the common trade name “Prozac”. Following a growing trend of drug repurposing and repositioning, 29 has been studied in other systems such as triple negative breast cancer (TNBC). Early evidence has shown that eEF2K is overexpressed in several types of malignancies, especially including TNBC. It was determined that 29 exhibits low-to-sub μM anti-proliferative IC50 activities against various TNBC cells in vitro, including MDA-MB-231 and MDA-MB-436 cells [95]. Additionally, 29 induced apoptosis and autophagic cell death in these TNBC cells [95]. The associated anti-TNBC mechanism of action studies has shown that fluoxetine induced autophagic cell death by inhibiting eEF2K and activating the AMPK-mTOR-ULK signaling pathway [95]. These results do well to suggest that inhibition of eEF2K may be a promising treatment strategy for TNBC, and also that other existing drugs may be able to be repurposed or repositioned for eEF2K inhibition. Although the activity of 29 has yet to be proven in vivo for TNBC, the information available about absorption, distribution, metabolism, excretion, and toxicity (ADMETox) of this drug in humans make it an interesting lead for further investigation.

4.1.8. Rottlerin

The natural product rottlerin, 5,7-dihydroxy-2,2-dimethyl-6-(2,4,6-trihydroxy-3-methyl-5-acetylbenzyl)-8-cinnamoyl-1,2-chromene (30, Figure 10), which is isolated from the pericarps of the red kamala tree Mallotus philippinensis, has been used as a protein kinase Cδ (PKCδ) selective inhibitor [96]. This molecule is associated with impacting a variety of cellular processes, such as proliferation, survival, apoptosis, and autophagy [97]. Early studies found that 30 can induce apoptosis on a variety of cancer cell types, such as breast cancer, colon cancer, lung cancer, leukemia, and myeloma [98,99,100,101]. It was initially speculated that this cytotoxic activity of 30 is related to its inhibitory effect on PKC, but it was later found that 30 induces tumor cell apoptosis through inhibition of eEF2K by using reverse phase-protein array (RPPA), cell viability (MTS) assay, and eEF2K knockdown technology [14,75,102]. Several studies have demonstrated that 30 also activates calcium channels and affects the function of mitochondria, further enhancing its use potential as an inhibitor of eEF2K [103,104]. It was reported that 30 significantly inhibits eEF2K at concentrations that are relevant both in vitro and in vivo, and this could explain its antiproliferative activity on human glioma cells and blockage of gliomal cells at the G1–S interface [105]. Furthermore, 30 downregulates the mRNA expression of eEF2K through the ubiquitin-proteasome pathway [105]. Although inhibition of eEF2K is not the only mechanism of action for 30, this does appear to be one important pathway for its activity profile, which should encourage further studies of structurally related natural products and analogues. Finally, the reported efficacy of 30 for reducing pancreatic tumors in vivo without overt effects on normal tissue offers more validation for both this compound and eEF2K as a molecular target for future drug development [106].

4.1.9. Thymoquinone

Thymoquinone (TQ, 31, Figure 10) is an active principal natural product found in seeds of the black cumin or Roman coriander plant, Nigella sativa [107]. As a potential anti-tumor agent, 31 has been shown to inhibit breast, lung, ovarian, liver, prostate, colorectal, and leukemic cancers in vitro and in vivo [108,109,110,111,112]. Also sometimes considered as a pan-assay interference nuisance compound (PAIN), 31 can interact with a variety of tumor-related targets and pathways, such as nuclear factor-kappa B (NF-κB), tumor necrosis factor-α (TNF-α), STAT3, PTEN, Bcl-2, p53, and PPAR-γ, etc., to exert anti-proliferation, induction of apoptosis and oxidative stress, cell cycle arrest, anti-angiogenesis, and cellular metastasis [109,112,113,114]. After more in-depth research on 31, it was found that this molecule can also inhibit TNBC cell proliferation and migration/invasion by inhibiting the NF-κB/miR-603/eEF2K pathway [115]. However, there are many obstacles to consider for the development of para-quinones and other PAINs like 31 as drugs and lead compounds, particularly including toxicity.

4.1.10. 6-hydroxystaurosporinone

6-hydroxystaurosporinone (32, Figure 10), a bisindole alkaloid, was first obtained from the fruit bodies of Groening’s slime mould, Lycogala epidendrum [116]. This compound exhibited in vitro antiproliferative activity against HeLa and Jurkat cells with IC50 values of 5.4 and 1.34 μg/mL, respectively [116]. Further in vitro mechanistic studies showed that 32 decreased phosphorylation of protein targets of PKC at 1 μg/mL, along with each eEF2K, PKA, and VEGFR-1 kinase at 10 μg/mL [116]. Other staurosporine analogues have been reported as kinase inhibitors and cytotoxic agents in the past, but these have yet to be tested for inhibitory activity against eEF2K [117]. Further studies could thus provide preliminary SAR data for this series of molecules.

4.1.11. Myriaporone 3/4

The polyketide myriaporone 3/4 (33, Figure 10), which was first isolated from the bryozoan false coral, Myriapora truncata, has been reported as an inhibitor of eukaryotic protein synthesis [118]. Furthermore, a mechanistic investigation was able to determine that 33 blocks protein synthesis in the elongation stage by directly binding to eEF2K to induce phosphorylation of eEF2 [119]. Compound 33 displays low nM anti-proliferative activity against several cancer cell lines (L-929, PtK2, KB-3-1, PC-3, and A-549) in vitro, inhibits angiogenesis-like tube formation by endothelial cells in vitro in nanomolar concentrations, and is significantly selective (≥ 300x) for acting on cancer cells over normal ones [119]. The potency and selectivity of 33 could make this compound an attractive lead molecule or drug candidate, but the activity apparently has yet to be validated in vivo.

4.1.12. Leptosin M

Leptosin M (34, Figure 10) was isolated from a Leptosphaeria sp. fungus strain originally separated from growth of the marine alga Sargassum tortile [120]. In vitro assays showed that 34 is broadly cytotoxic against a panel array of 39 different human cancer cell lines in low μM concentrations [120]. In addition, 34 was found to inhibit PTK and eEF2K function by 40–70% at 10 μg/mL (~13 μM) but showed no activity against PKA, PKC, or EGFR at 100 μg/mL (~130 μM) [120]. It would thus be interesting to see future studies evaluate the eEF2K inhibitory potential of other natural leptosins or analogues thereof.

4.1.13. Calyxin Y

Calyxin Y (35, Figure 10), isolated from Alpinia katsumadai, a ginger or galangal used in Traditional Chinese medicine (TCM), exhibits potent antiproliferative activity in several cancer cell lines. One of the many cancer types that eEF2K is overexpressed in is hepatocellular carcinoma (HCC). In 2017, it was reported that a combination of 35 and cisplatin could synergistically inhibit cell viability and induce cell death in both the wild type HepG2 and cisplatin-resistant HepG2 cancer cells [121]. Mechanistic studies revealed that 35 down-regulates eEF2K by promoting SCF βTrCP-mediated protein degradation and enhances the anticancer activity of cisplatin in HCC cells via apoptosis and autophagy [121]. It is exciting that the drug–lead combination also showed in vivo efficacy in a cisplatin-resistant HepG2 xenograft trial without overt toxicity [121], offering some promise of this mixture or other imaginable combinations having use in future therapy.

4.2. Discovery and Development of Multi-Target Inhibitors of eEF2K

4.2.1. Inhibitors Targeting PLK1/eEF2K

Multi-target inhibitors are becoming more purposefully developed with the intention of improving efficacy through synergism, safety through reduced dosage, and general prevention of single-target mutation-based drug resistance [127]. A series of 1-(4-(2-substituted-pyridin-4-yl)-3-substituted-phenyl)-3-phenylurea derivatives (36, Figure 11) were predicted to be PLK1/eEF2K dual-targeting inhibitors by in silico methods [122]. These compounds were synthesized (Scheme 6), and several were validated as having in vitro inhibition of PLK1 and eEF2K. In brief, 4-bromopyridin-2-amine (37) was alkylated to afford 38, which was then coupled to a series of 2-substituted 1-bromo-4-nitrobenzenes (39) via the canonical Suzuki reaction to form intermediates (40). After amino protection, these compounds (41) were reduced from nitros to amines (42), and the final products (36) were obtained by N-substitution with the aromatic isocyanates (43). The resulting SAR study indicated that the best lead compound was 44 (“18i”; Figure 11), which contains an ethyl group at the R1 position, a trifluoromethyl group with the R2 position and the 3-chlorine atom at the R3 position. This compound showed low micromolar in vitro cytotoxicity against a variety of cancer cell lines and strongly inhibited both target kinases (PLK1 IC50 = 0.085 μM and eEF2K IC50 = 0.762 μM) [122]. Accordingly, the multi-target strategy appears to be effective with this kinase set and could lead to further exciting developments in the future.

4.2.2. Inhibitors Targeting GLP-1R/eEF2K

In 2019, two new sorbicillinoids, 13-hydroxy-dihydrotrichodermolide (45, Figure 12) and 10,11,27,28-tetrahydrotrisorbicillinone C (46, Figure 12), were discovered from the sponge-derived fungus Penicillium chrysogenum [123]. These natural products strongly inhibit eEF2K and GLP-1R, with Kd values of 118 and 28.5 nM, respectively, for 45, and 74.6 and 16.2  nM, respectively, for 46 [123]. GLP-1R is more typically related to diabetes than cancers, but it also has implications for various cancers that especially include those effecting the pancreas [128]. The further evaluation of these complex natural products or related analogues in vitro and in vivo could prove to be helpful for the development of eEF2K inhibiting anticancer agents, perhaps especially for pancreatic cancers.

4.2.3. Inhibitors Targeting the Protein/Protein Interaction of Hsp90 and eEF2K

In 2001, Hait et al. first revealed that the natural product geldanamycin (47, Figure 13) and its synthetic derivative 17-allylamino-17-demethoxygeldanamycin (17-AAG or tanespimycin; 48, Figure 13) had nanomolar inhibitory activity against human glioma in vitro, and significantly inhibited glioma xenografts in nude mice in vivo studies [124]. Further studies showed that these molecules disrupt the interaction of eEF2K with Hsp90 and that this is an important mechanism of cancer cell cytotoxicity for 47 and 48 [124]. At the end of 2020, the Takahashi group reported, after high-throughput screening of a compound library of natural products, that 47, 48, and related derivatives may inhibit receptor tyrosine kinase-like orphan receptor 1 (ROR1) [125]. This indicates that the inhibition of HSP90 by compounds with the core scaffold of 47 may be promising for treatment of ROR1-positive lung adenocarcinoma, and further expands the known mechanism of action for the molecules [125]. Compound 48 has been the subject of dozens of clinical trials for treating various forms of cancers [126], and the core natural product scaffold of geldanamycin is a promising lead for the development of further generations of inhibitors.

5. The Role of Natural Product and Synthetic Small Molecule Activators of Eef2k

A continually increasing number of studies have shown that eEF2K is related to the occurrence and development of a variety of tumors, and is considered a potential target for tumor chemotherapy. Subsequently, the development of related inhibitors has also made certain progress. Interestingly, the regulation of tumor autophagy by protein synthesis is bidirectional [129]. Some studies have found that eEF2K activators can also have certain anti-tumor effects. For instance, ritonavir (RTV, 49; Figure 14) and lopinavir (LPV, 50; Figure 14) led to an associated increase in eEF2 phosphorylation via the AMPK/eEF2K pathway along with impaired protein synthesis [130,131,132]. Resveratrol (51, Figure 14) can inhibit the proliferation and migration of vascular endothelial cells by the activation of AMPK and induced phosphorylation of residue serine 398 of eEF2K, thus leading to inhibited eEF2 activity [133]. The activation of eEF2K by this molecule is illuminating for the multiple directions available for its use as a drug candidate [133]. Huanglian Jiedu decoction (HLJDD), a traditional Chinese Medicinal (TCM), could activate AMPK signaling and further inhibit the mTOR pathway, thus reducing the phosphorylation of eEF2K in hepatocellular carcinoma (HCC) cells [134]. The activated eEF2K leads to the loss of eEF2 activity, and therefore the elongation of new peptides is blocked [134].

6. Conclusions and Future Outlook

The compounds here reviewed come from many aspects of medicinal chemistry: computer-assisted drug design, drug repurposing and repositioning, high throughput screening, multi-targeted drug design, natural products discovery and derivatization, and synthetic structure optimization. More than a third of the molecules are natural products that offer complex chemical scaffolds for further optimization, especially after in vivo anticancer activity has been validated (i.e., for compounds 30, 31, 35, 47, and 48). However, there are still certain shortcomings to be resolved: (1) Existing inhibitors, and especially the natural products, are not often as selective to eEF2K as would be desired. Therefore, the full mechanism of action of such compounds is not clear. (2) Many current ATP non-competitive eEF2K inhibitors have potent inhibitory effects in vitro, but their anti-tumor effects in vivo have not been studied, are not pronounced, or may be absent for any number of reasons that arise when advancing candidates from protein inhibition assays to whole cells in vitro and onward to animals. (3) While some ATP competitive eEF2K inhibitors show anticancer efficacy in vivo, there remains an opportunity to further optimize their drug-like properties or even potency. (4) No single best core pharmacophore nucleus that specifically inhibits eEF2K has been reported, and the structures of the existing inhibitors are very diverse. (5) The crystal structure of eEF2K remains unknown, and although MHCK A can be used as a template, the similarity between the two is only about 40%. This may contribute to the fact that many designed inhibitors are not yet very specific.
Although important progress has been made in the research and development of eEF2K inhibitors, the complete mechanism of anti-cancer drugs is usually more complicated than single-target action. Further understanding the mechanism of action for the compounds presented here, or the molecular characterization of eEF2K, will help in the optimization of the structure for new eEF2K inhibitors. Additionally, the design and synthesis of purposeful multi-target inhibitors of eEF2K with other relevant anti-cancer targets is also an important research direction for eEF2K inhibitors. Considering the structural diversity and complexity of eEF2K inhibitors, especially natural product inhibitors, how to produce them at scale at low cost remains a problem to be solved. Therefore, further synthetic organic and medicinal chemistry efforts are expected to expand the preclinical and clinical development of eEF2K inhibitors.
Overcoming all of these challenges remains an area of great opportunity for more scientific research efforts to be devoted to the discovery and development of new eEF2K inhibitors. It is suggested that eEF2K inhibitors can be developed as sensitizing agents to improve the outcomes of radiotherapy, as was shown by the in vivo efficacy in an esophageal cancer xenograft of such treatment in combination with compound 1. Furthermore, since compounds 28 and 30 are effective individual anticancer agents in vivo, 35 in combination with cisplatin can overcome cisplatin-resistance in a HepG2 xenograft, and eEF2K has emerged as one important mechanism for the known single agent anticancer activity of 47 and 48, this molecular target seems to hold great promise for the development of new first-in-class anticancer drugs.

Author Contributions

Conceptualization: B.Z., N.W. and C.B.N.; literature review and analysis: J.Z., Q.Z., Z.W. and C.B.N.; writing—original draft preparation: B.Z. and N.W.; review and editing: S.H., Y.Z. and C.B.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work and the associated article publishing charge was funded by the National Natural Science Foundation of China (91856126 and 82050410451), the Natural Science Foundation of Ningbo City (2018A610410), Foundation of Ningbo University for Grant (XYL20023), the National 111 Project of China (D16013), the Li Dak Sum Yip Yio Chin Kenneth Li Marine Biopharmaceutical Development Fund, and the K.C. Wong Magna Fund in Ningbo University.

Institutional Review Board Statement

Not applicable. The animal studies discussed in this review article were not conducted by the authors of the manuscript.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this review.

References

  1. Lee, Y.T.; Tan, Y.J.; Oon, C.E. Molecular targeted therapy: Treating cancer with specificity. Eur. J. Clin. Pharmacol. 2018, 834, 188–196. [Google Scholar] [CrossRef] [PubMed]
  2. Rosenzweig, S.A. Chapter Three—Acquired Resistance to Drugs Targeting Tyrosine Kinases. Adv. Cancer. Res. 2018, 138, 71–98. [Google Scholar] [PubMed]
  3. Cote, G.P.; Luo, X.; Murphy, M.B.; Egelhoff, T.T. Mapping of the novel protein kinase catalytic domain of dictyostelium myosin II heavy chain kinase A. J. Biol. Chem. 1997, 272, 6846–6849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kenney, J.W.; Moore, C.E.; Wang, X.; Proud, C.G. Eukaryotic elongation factor 2 kinase, an unusual enzyme with multiple roles. Adv. Biol. Regul. 2014, 55, 15–27. [Google Scholar] [CrossRef]
  5. Proud, C.G. Regulation and roles of elongation factor 2 kinase. Biochem. Soc. Trans. 2015, 43, 328–332. [Google Scholar] [CrossRef] [PubMed]
  6. Yamaguchi, H.; Matsushita, M.; Nairn, A.C.; Kuriyan, J. Crystal structure of the atypical protein kinase domain of a TRP channel with phosphotransferase activity. Mol. Cell 2001, 7, 1047–1057. [Google Scholar] [CrossRef]
  7. Delaidelli, A.; Khan, D.; Leprivier, G.; Pfister, S.M.; Taylor, M.D.; Maris, J.M.; Sorensen, P. OS5-173 Inhibition of eEF2K as a novel therapeutic strategy in neuroblastoma and medulloblastoma. Can. J. Neurol. Sci. 2016, 43, S3. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, J.C.; Voisin, V.; Wang, S.; Wang, D.Y.; Jones, R.A.; Datti, A.; Uehling, D.; Al-Awar, R.; Egan, S.E.; Bader, G.D. Combined deletion of Pten and p53 in mammary epithelium accelerates triple-negative breast cancer with dependency on eEF2K. EMBO Mol. Med. 2015, 6, 1542–1560. [Google Scholar] [CrossRef]
  9. Gassart, A.D.; Martinon, F. Translating the anticancer properties of eEF2K. Cell Cycle 2017, 16, 299–300. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, H.; Bialkowska, A.; Rusovici, R.; Chanchevalap, S.; Shim, H.; Katz, J.P.; Yang, V.W.; Yun, C.C. Lysophosphatidic acid facilitates proliferation of colon cancer cells via induction of Krüppel-like factor 5. J. Biol. Chem. 2007, 282, 15541–15549. [Google Scholar] [CrossRef] [Green Version]
  11. Zhou, Y.; Li, Y.; Xu, S.; Lu, J.; Zhu, Z.; Chen, S.; Tan, Y.; He, P.; Xu, J.; Proud, C.G.; et al. Eukaryotic elongation factor 2 kinase promotes angiogenesis in hepatocellular carcinoma via PI3K/Akt and STAT3. Int. J. Cancer. 2020, 146, 1383–1395. [Google Scholar] [CrossRef] [PubMed]
  12. Horman, S.; Beauloye, C.; Vertommen, D.; Vanoverschelde, J.L.; Hue, L.; Rider, M.H. Myocardial ischemia and increased heart work modulate the phosphorylation state of eukaryotic elongation factor-2. J. Biol. Chem. 2003, 278, 41970–41976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Karakas, D.; Ozpolat, B. Eukaryotic elongation factor-2 kinase (eEF2K) signaling in tumor and microenvironment as a novel molecular target. J. Mol. Med. 2020, 98, 775–787. [Google Scholar] [CrossRef] [PubMed]
  14. Ashour, A.A.; Abdel-Aziz, A.A.H.; Mansour, A.M.; Alpay, S.N.; Huo, L.F.; Ozpolat, B. Targeting elongation factor-2 kinase (eEF-2K) induces apoptosis in human pancreatic cancer cells. Apoptosis 2014, 19, 241–258. [Google Scholar] [CrossRef]
  15. Moore, C.E.; Wang, X.; Xie, J.; Pickford, J.; Barron, J.; Regufe da Mota, S.; Versele, M.; Proud, C.G. Elongation factor 2 kinase promotes cell survival by inhibiting protein synthesis without inducing autophagy. Cell Signal. 2016, 28, 284–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Wang, X.; Li, W.; Williams, M.; Terada, N.; Alessi, D.R.; Proud, C.G. Regulation of elongation factor 2 kinase by p90(RSK1) and p70 S6 kinase. EMBO J. 2001, 20, 4370–4379. [Google Scholar] [CrossRef] [PubMed]
  17. Grant, C.M. Regulation of translation by hydrogen peroxide. Antioxid. Redox. Signal. 2011, 15, 191–203. [Google Scholar] [CrossRef]
  18. Spahn, C.M.; Gomez-Lorenzo, M.G.; Grassucci, R.A.; Jørgensen, R.; Andersen, G.R.; Beckmann, R.; Penczek, P.A.; Ballesta, J.P.; Frank, J. Domain movements of elongation factor eEF2 and the eukaryotic 80S ribosome facilitate tRNA translocation. EMBO J. 2004, 23, 1008–1019. [Google Scholar] [CrossRef] [Green Version]
  19. Hizli, A.A.; Chi, Y.; Swanger, J.; Carter, J.H.; Liao, Y.; Welcker, M.; Ryazanov, A.G.; Clurman, B.E. Phosphorylation of eukaryotic elongation factor 2 (eEF2) by cyclin A-cyclin-dependent kinase 2 regulates its inhibition by eEF2 kinase. Mol. Cell. Biol. 2013, 33, 596–604. [Google Scholar] [CrossRef] [Green Version]
  20. Kruiswijk, F.; Yuniati, L.; Magliozzi, R.; Low, T.Y.; Lim, R.; Bolder, R.; Mohammed, S.; Proud, C.G.; Heck, A.J.; Pagano, M.; et al. Coupled activation and degradation of eEF2K regulates protein synthesis in response to genotoxic stress. Sci. Signal. 2012, 5, ra40. [Google Scholar] [CrossRef] [Green Version]
  21. Leprivier, G.; Rotblat, B.; Khan, D.; Jan, E.; Sorensen, P.H. Stress-mediated translational control in cancer cells. Biochim. Biophys. Acta 2015, 1849, 845–860. [Google Scholar] [CrossRef]
  22. Russnes, H.G.; Caldas, C. eEF2K-a new target in breast cancers with combined inactivation of p53 and PTEN. EMBO Mol. Med. 2014, 6, 1512–1514. [Google Scholar] [CrossRef] [PubMed]
  23. Hamurcu, Z.; Ashour, A.; Kahraman, N.; Ozpolat, B. FOXM1 regulates expression of eukaryotic elongation factor 2 kinase and promotes proliferation, invasion and tumorgenesis of human triple negative breast cancer cells. Oncotarget 2016, 7, 16619–16635. [Google Scholar] [CrossRef] [Green Version]
  24. Shi, N.; Chen, X.; Liu, R.; Wang, D.; Su, M.; Wang, Q.; He, A.; Gu, H. Eukaryotic elongation factors 2 promotes tumor cell proliferation and correlates with poor prognosis in ovarian cancer. Tissue Cell 2018, 53, 53–60. [Google Scholar] [CrossRef] [PubMed]
  25. Bayraktar, R.; Pichler, M.; Kanlikilicer, P.; Ivan, C.; Bayraktar, E.; Kahraman, N.; Aslan, B.; Oguztuzun, S.; Ulasli, M.; Arslan, A.; et al. MicroRNA 603 acts as a tumor suppressor and inhibits triple-negative breast cancer tumorigenesis by targeting elongation factor 2 kinase. Oncotarget 2017, 8, 11641–11658. [Google Scholar] [CrossRef] [Green Version]
  26. Bircan, H.A.; Gurbuz, N.; Pataer, A.; Caner, A.; Kahraman, N.; Bayraktar, E.; Bayraktar, R.; Erdogan, M.A.; Kabil, N.; Ozpolat, B. Elongation factor-2 kinase (eEF-2K) expression is associated with poor patient survival and promotes proliferation, invasion and tumor growth of lung cancer. Lung Cancer 2018, 124, 31–39. [Google Scholar] [CrossRef] [PubMed]
  27. Leprivier, G.; Remke, M.; Rotblat, B.; Dubuc, A.; Mateo, A.R.; Kool, M.; Agnihotri, S.; El-Naggar, A.; Yu, B.; Somasekharan, S.P.; et al. The eEF2 kinase confers resistance to nutrient deprivation by blocking translation elongation. Cell 2013, 153, 1064–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Jewell, J.L.; Guan, K.L. Nutrient signaling to mTOR and cell growth. Trends Biochem. Sci. 2013, 38, 233–242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Ryazanov, A.G.; Shestakova, E.A.; Natapov, P.G. Phosphorylation of elongation factor 2 by EF-2 kinase affects rate of translation. Nature 1988, 334, 170–173. [Google Scholar] [CrossRef] [PubMed]
  30. Proud, C.G. mTORC1 regulates the efficiency and cellular capacity for protein synthesis. Biochem. Soc. Trans. 2013, 41, 923–926. [Google Scholar] [CrossRef] [Green Version]
  31. Proud, C.G. Signalling to translation: How signal transduction pathways control the protein synthetic machinery. Biochem. J. 2007, 403, 217–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Faller, W.J.; Jackson, T.J.; Knight, J.R.; Ridgway, R.A.; Jamieson, T.; Karim, S.A.; Jones, C.; Radulescu, S.; Huels, D.J.; Myant, K.B.; et al. mTORC1-mediated translational elongation limits intestinal tumour initiation and growth. Nature 2015, 517, 497–500. [Google Scholar] [CrossRef] [Green Version]
  33. Connolly, E.; Braunstein, S.; Formenti, S.; Schneider, R.J. Hypoxia inhibits protein synthesis through a 4E-BP1 and elongation factor 2 kinase pathway controlled by mTOR and uncoupled in breast cancer cells. Mol. Cell. Biol. 2006, 26, 3955–3965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Moore, C.E.J.; Mikolajek, H.; Sergio, R.D.M.; Wang, X.; Kenney, J.W.; Werner, J.R.M.; Proud, C.G.J.M.; Biology, C. Elongation Factor 2 Kinase Is Regulated by Proline Hydroxylation and Protects Cells during Hypoxia. Mol. Cell. Biol. 2015, 35, 1788–1804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [Green Version]
  36. Cheng, Y.; Ren, X.; Yuan, Y.; Shan, Y.; Li, L.; Chen, X.; Zhang, L.; Takahashi, Y.; Yang, J.W.; Han, B.; et al. eEF-2 kinase is a critical regulator of Warburg effect through controlling PP2A-A synthesis. Oncogene 2016, 35, 6293–6308. [Google Scholar] [CrossRef]
  37. Helmlinger, G.; Schell, A.; Dellian, M.; Forbes, N.S.; Jain, R.K. Acid production in glycolysis-impaired tumors provides new insights into tumor metabolism. Clin. Cancer Res. 2002, 8, 1284–1291. [Google Scholar]
  38. Shime, H.; Yabu, M.; Akazawa, T.; Kodama, K.; Matsumoto, M.; Seya, T.; Inoue, N. Tumor-secreted lactic acid promotes IL-23/IL-17 proinflammatory pathway. J. Immunol. 2008, 180, 7175–7183. [Google Scholar] [CrossRef]
  39. Dorovkov, M.V.; Pavur, K.S.; Petrov, A.G. Regulation of elongation factor-2 kinase by pH. Biochemistry 2002, 41, 13444–13450. [Google Scholar] [CrossRef] [PubMed]
  40. Xie, J.; Mikolajek, H.; Pigott, C.R.; Hooper, K.J.; Mellows, T.; Moore, C.E.; Mohammed, H.; Werner, J.M.; Thomas, G.J.; Proud, C.G. Molecular mechanism for the control of eukaryotic elongation factor 2 kinase by pH: Role in cancer cell survival. Mol. Cell. Biol. 2015, 35, 1805–1824. [Google Scholar] [CrossRef] [Green Version]
  41. Theodoropoulos, G.E.; Gazouli, M.; Vaiopoulou, A.; Leandrou, M.; Nikouli, S.; Vassou, E.; Kouraklis, G.; Nikiteas, N. Polymorphisms of Caspase 8 and Caspase 9 gene and colorectal cancer susceptibility and prognosis. Int. J. Colorectal. Dis. 2011, 26, 1113–1118. [Google Scholar] [CrossRef] [PubMed]
  42. Park, H.S.; Jun, D.Y.; Han, C.R.; Kim, Y.H.J.B.B.R.C. Protein tyrosine kinase p56lck-deficiency confers hypersusceptibility to rho-fluorophenylalanine (pFPhe)-induced apoptosis by augmenting mitochondrial apoptotic pathway in human Jurkat T cells. Biochem. Biophys. Res. Commun. 2008, 377, 280–285. [Google Scholar] [CrossRef]
  43. Wang, L.; Du, F.; Wang, X. TNF-alpha induces two distinct caspase-8 activation pathways. Cell 2008, 133, 693–703. [Google Scholar] [CrossRef] [Green Version]
  44. Bellail, A.C.; Tse, M.C.L.; Song, J.H.; Phuphanich, S.; Olson, J.J.; Sun, S.Y.; Hao, C.H. DR5-mediated DISC controls caspase-8 cleavage and initiation of apoptosis in human glioblastomas. J. Cell. Mol. Med. 2010, 14, 1303–1317. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, Y.; Cheng, Y.; Zhang, L.; Ren, X.C.; Huber-Keener, K.J.; Lee, S.; Yun, J.; Wang, H.G.; Yang, J.M. Inhibition of eEF-2 kinase sensitizes human glioma cells to TRAIL and down-regulates Bcl-xL expression. Biochem. Biophys. Res. Commun. 2011, 414, 129–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Van Loo, G.; Saelens, X.; van Gurp, M.; MacFarlane, M.; Martin, S.J.; Vandenabeele, P. The role of mitochondrial factors in apoptosis: A Russian roulette with more than one bullet. Cell Death Differ. 2002, 9, 1031–1042. [Google Scholar] [CrossRef] [PubMed]
  47. Tekedereli, I.; Alpay, S.N.; Tavares, C.D.J.; Cobanoglu, Z.E.; Kaoud, T.S.; Sahin, I.; Sood, A.K.; Lopez-Berestein, G.; Dalby, K.N.; Ozpolat, B. Targeted Silencing of Elongation Factor 2 Kinase Suppresses Growth and Sensitizes Tumors to Doxorubicin in an Orthotopic Model of Breast Cancer. PLoS ONE 2012, 7, e41171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Parmer, T.G.; Ward, M.D.; Yurkow, E.J.; Vyas, V.H.; Kearney, T.J.; Hait, W.N. Activity and regulation by growth factors of calmodulin-dependent protein kinase III (elongation factor 2-kinase) in human breast cancer. Br. J. Cancer 1999, 79, 59–64. [Google Scholar] [CrossRef] [Green Version]
  49. Roberts, E.C.; Hammond, K.; Traish, A.M.; Resing, K.A.; Ahn, N.G. Identification of G2/M targets for the MAP kinase pathway by functional proteomics. Proteomics 2010, 6, 4541–4553. [Google Scholar] [CrossRef] [PubMed]
  50. Ratan, R.R.; Maxfield, F.R.; Shelanski, M.L. Long-lasting and rapid calcium changes during mitosis. J. Cell Biol. 1988, 107, 993–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Santella, L. The role of calcium in the cell cycle: Facts and hypotheses. Biochem. Biophys. Res. Commun. 1998, 244, 317–324. [Google Scholar] [CrossRef] [PubMed]
  52. Berridge, M.J.; Lipp, P.; Bootman, M.D. The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 2000, 1, 11–21. [Google Scholar] [CrossRef]
  53. Gutzkow, K.B.; Låhne, H.U.; Naderi, S.; Torgersen, K.M.; Skålhegg, B.; Koketsu, M.; Uehara, Y.; Blomhoff, H.K. Cyclic AMP inhibits translation of cyclin D3 in T lymphocytes at the level of elongation by inducing eEF2-phosphorylation. Cell. Signal. 2003, 15, 871–881. [Google Scholar] [CrossRef]
  54. Pyr Dit Ruys, S.; Wang, X.; Smith, E.M.; Herinckx, G.; Hussain, N.; Rider, M.H.; Vertommen, D.; Proud, C.G. Identification of autophosphorylation sites in eukaryotic elongation factor-2 kinase. Biochem. J. 2012, 442, 681–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Smith, E.M.; Proud, C.G. cdc2–cyclin B regulates eEF2 kinase activity in a cell cycle- and amino acid-dependent manner. EMBO J. 2008, 27, 1005–1016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Lee, B.; Sandhu, S.; Mcarthur, G. Cell cycle control as a promising target in melanoma. Curr. Opin. Oncol. 2015, 27, 141–150. [Google Scholar] [CrossRef]
  57. Petroni, G.; Formenti, S.C.; Chen-Kiang, S.; Galluzzi, L. Immunomodulation by anticancer cell cycle inhibitors. Nat. Rev. Immunol. 2020, 20, 669–679. [Google Scholar] [CrossRef]
  58. Levine, B.; Kroemer, G. Autophagy in the Pathogenesis of Disease. Cell 2008, 132, 27–42. [Google Scholar] [CrossRef] [Green Version]
  59. Mizushima, N.; Levine, B.; Cuervo, A.M.; Klionsky, D.J. Autophagy fights disease through cellular self-digestion. Nature 2008, 451, 1069–1075. [Google Scholar] [CrossRef] [Green Version]
  60. Kondo, Y.; Kanzawa, T.; Sawaya, R.; Kondo, S. The role of autophagy in cancer development and response to therapy. Nat. Rev. Cancer 2005, 5, 726–734. [Google Scholar] [CrossRef]
  61. Cheng, Y.; Li, H.; Ren, X.; Niu, T.; Hait, W.N.; Yang, J. Cytoprotective Effect of the Elongation Factor-2 Kinase-Mediated Autophagy in Breast Cancer Cells Subjected to Growth Factor Inhibition. PLoS ONE 2010, 5, e9715. [Google Scholar] [CrossRef]
  62. Jung, S.; Jeong, H.; Yu, S.W. Autophagy as a decisive process for cell death. Exp. Mol. Med. 2020, 52, 921–930. [Google Scholar] [CrossRef] [PubMed]
  63. White, E.; Mehnert, J.M.; Chan, C.S. Autophagy, Metabolism, and Cancer. Clin. Cancer Res. 2015, 21, 5037–5046. [Google Scholar] [CrossRef] [Green Version]
  64. White, E. Deconvoluting the context-dependent role for autophagy in cancer. Nat. Rev. Cancer 2012, 12, 401–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Cheng, Y.; Ren, X.; Zhang, Y.; Patel, R.; Sharma, A.; Wu, H.; Robertson, G.P.; Yan, L.; Rubin, E.; Yang, J.M. eEF-2 kinase dictates cross-talk between autophagy and apoptosis induced by Akt Inhibition, thereby modulating cytotoxicity of novel Akt inhibitor MK-2206. Cancer Res. 2011, 71, 2654–2663. [Google Scholar] [CrossRef] [Green Version]
  66. Zhao, Y.Y.; Tian, Y.; Liu, L.; Zhan, J.H.; Hou, X.; Chen, X.; Zhou, T.; Huang, Y.; Zhang, L. Inhibiting eEF-2 kinase-mediated autophagy enhanced the cytocidal effect of AKT inhibitor on human nasopharyngeal carcinoma. Drug Des. Dev. Ther. 2018, 12, 2655–2663. [Google Scholar] [CrossRef] [Green Version]
  67. Py, B.F.; Boyce, M.; Yuan, J. A critical role of eEF-2K in mediating autophagy in response to multiple cellular stresses. Autophagy 2009, 5, 393–396. [Google Scholar] [CrossRef] [Green Version]
  68. Boyce, M.; Py, B.F.; Ryazanov, A.G.; Minden, J.S.; Long, K.; Ma, D.; Yuan, J. A pharmacoproteomic approach implicates eukaryotic elongation factor 2 kinase in ER stress-induced cell death. Cell Death Differ. 2008, 15, 589–599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Wu, H.; Zhu, H.; Liu, D.X.; Niu, T.K.; Ren, X.; Patel, R.; Hait, W.N.; Yang, J.M. Silencing of elongation factor-2 kinase potentiates the effect of 2-deoxy-D-glucose against human glioma cells through blunting of autophagy. Cancer Res. 2009, 69, 2453–2460. [Google Scholar] [CrossRef] [Green Version]
  70. Xie, C.M.; Liu, X.Y.; Sham, K.W.; Lai, J.M.; Cheng, C.H. Silencing of EEF2K (eukaryotic elongation factor-2 kinase) reveals AMPK-ULK1-dependent autophagy in colon cancer cells. Autophagy 2014, 10, 1495–1508. [Google Scholar] [CrossRef] [Green Version]
  71. Chen, C.; Xu, Z.Q.; Zong, Y.P.; Ou, B.C.; Shen, X.H.; Feng, H.; Zheng, M.H.; Zhao, J.K.; Lu, A.G. CXCL5 induces tumor angiogenesis via enhancing the expression of FOXD1 mediated by the AKT/NF-κB pathway in colorectal cancer. Cell Death Dis. 2019, 10, 178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Hoshi, T.; Watanabe Miyano, S.; Watanabe, H.; Sonobe, R.M.K.; Seki, Y.; Ohta, E.; Nomoto, K.; Matsui, J.; Funahashi, Y. Lenvatinib induces death of human hepatocellular carcinoma cells harboring an activated FGF signaling pathway through inhibition of FGFR–MAPK cascades. Biochem. Biophys. Res. Commun. 2019, 513, 1–7. [Google Scholar] [CrossRef]
  73. Zhu, H.; Song, H.; Chen, G.; Yang, X.; Liu, J.; Ge, Y.; Lu, J.; Qin, Q.; Zhang, C.; Xu, L.; et al. eEF2K promotes progression and radioresistance of esophageal squamous cell carcinoma. Radiother. Oncol. 2017, 124, 439–447. [Google Scholar] [CrossRef] [PubMed]
  74. Shi, Q.; Xu, X.; Liu, Q.; Luo, F.; Shi, J.; He, X. MicroRNA-877 acts as a tumor suppressor by directly targeting eEF2K in renal cell carcinoma. Oncol. Lett. 2016, 11, 1474–1480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ashour, A.A.; Gurbuz, N.; Alpay, S.N.; Abdel-Aziz, A.A.H.; Mansour, A.M.; Huo, L.; Ozpolat, B. Elongation factor-2 kinase regulates TG2/β1 integrin/Src/uPAR pathway and epithelial-mesenchymal transition mediating pancreatic cancer cells invasion. J. Cell. Mol. Med. 2014, 18, 2235–2251. [Google Scholar] [CrossRef]
  76. Bayraktar, R.; Ivan, C.; Bayraktar, E.; Kanlikilicer, P.; Kabil, N.N.; Kahraman, N.; Mokhlis, H.A.; Karakas, D.; Rodriguez-Aguayo, C.; Arslan, A.; et al. Dual Suppressive Effect of miR-34a on the FOXM1/eEF2-Kinase Axis Regulates Triple-Negative Breast Cancer Growth and Invasion. Clin. Cancer Res. 2018, 24, 4225–4241. [Google Scholar] [CrossRef] [Green Version]
  77. Xie, J.; Shen, K.; Lenchine, R.V.; Gethings, L.A.; Trim, P.J.; Snel, M.F.; Zhou, Y.; Kenney, J.W.; Kamei, M.; Kochetkova, M.; et al. Eukaryotic elongation factor 2 kinase upregulates the expression of proteins implicated in cell migration and cancer cell metastasis. Int. J. Cancer 2018, 142, 1865–1877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Will, N.; Lee, K.; Hajredini, F.; Giles, D.H.; Abzalimov, R.R.; Clarkson, M.; Dalby, K.N.; Ghose, R. Structural Dynamics of the Activation of Elongation Factor 2 Kinase by Ca(2+)-Calmodulin. J. Mol. Biol. 2018, 430, 2802–2821. [Google Scholar] [CrossRef]
  79. Alexey, R.G.; Michael, D.W. Identification of a new class of protein kinases represented by eukaryotic elongation factor-2 kinase. Proc. Natl. Acad. Sci. USA 1997, 94, 4884–4889. [Google Scholar]
  80. Yamamoto, K.; Kitayama, T.; Ishida, N.; Watanabe, T.; Tanabe, H.; Takatani, M.; Okamoto, T.; Utsumi, R. Identification and Characterization of a Potent Antibacterial Agent, NH125 against Drug-resistant Bacteria. Biosci. Biotech. Biochem. 2000, 64, 919–923. [Google Scholar] [CrossRef] [PubMed]
  81. Arora, S.; Yang, J.M.; Kinzy, T.; Utsumi, R.; Okamoto, T.; Kitayama, T.; Ortiz, P.; Hait, W. Identification and Characterization of an Inhibitor of Eukaryotic Elongation Factor 2 Kinase against Human Cancer Cell Lines. Cancer Res. 2003, 63, 6894–6899. [Google Scholar]
  82. Chen, Z.; Gopalakrishnan, S.M.; Bui, M.H.; Soni, N.B.; Warrior, U.; Johnson, E.F.; Donnelly, J.B.; Glaser, K.B. 1-Benzyl-3-cetyl-2-methylimidazolium iodide (NH125) induces phosphorylation of eukaryotic elongation factor-2 (eEF2). J. Biol. Chem. 2011, 286, 43951–43958. [Google Scholar] [CrossRef] [Green Version]
  83. Devkota, A.K.; Tavares, C.D.; Warthaka, M.; Abramczyk, O.; Marshall, K.D.; Kaoud, T.S.; Gorgulu, K.; Ozpolat, B.; Dalby, K.N. Investigating the kinetic mechanism of inhibition of elongation factor 2 kinase by NH125: Evidence of a common in vitro artifact. Biochemistry 2012, 51, 2100–2112. [Google Scholar] [CrossRef] [Green Version]
  84. Hori, H.; Nagasawa, H.; Ishibashi, M.; Uto, Y.; Hirata, A.; Saijo, K.; Ohkura, K.; Kirk, K.L.; Uehara, Y. TX-1123: An antitumor 2-hydroxyarylidene-4-cyclopentene-1,3-dione as a protein tyrosine kinase inhibitor having low mitochondrial toxicity. Bioorg. Med. Chem. 2002, 10, 3257–3265. [Google Scholar] [CrossRef]
  85. Tomoko, K.; Muneyoshi, O.; Hideyuki, Y. Mechanisms underlying the relaxation by A484954, a eukaryotic elongation factor 2 kinase inhibitor, in rat isolated mesenteric artery. J. Pharmacol. Sci. 2018, 137, 86–92. [Google Scholar]
  86. Edupuganti, R.; Wang, Q.; Tavares, C.D.J.; Chitjian, C.A.; Bachman, J.L.; Ren, P.; Anslyn, E.V.; Dalby, K.N. Synthesis and biological evaluation of pyrido[2,3-d]pyrimidine-2,4-dione derivatives as eEF-2K inhibitors. Bioorg. Med. Chem. 2014, 22, 4910–4916. [Google Scholar] [CrossRef] [PubMed]
  87. Liu, Y.; Zhen, Y.; Wang, G.; Yang, G.; Fu, L.; Liu, B.; Ouyang, L. Designing an eEF2K-Targeting PROTAC small molecule that induces apoptosis in MDA-MB-231 cells. Eur. J. Med. Chem. 2020, 204, 112505. [Google Scholar] [CrossRef]
  88. Koketsu, M.; Senda, T.; Yoshimura, K.; Ishihara, H. Synthesis and characterization of novel 1,3-selenazine derivatives. BF3·Et2O-assisted reaction of primary selenoamides with α,β-unsaturated ketones. J. Chem. Soc. Perkin Trans. 1999, 1, 453–456. [Google Scholar] [CrossRef]
  89. Cho, S.I.; Koketsu, M.; Ishihara, H.; Matsushita, M.; Nairn, A.C.; Fukazawa, H.; Uehara, Y. Novel compounds, ‘1,3-selenazine derivatives’ as specific inhibitors of eukaryotic elongation factor-2 kinase. Biochim. Biophys. Acta (BBA) Gen. Subj. 2000, 1475, 207–215. [Google Scholar] [CrossRef]
  90. Reynisson, J.; Court, W.; O’Neill, C.; Day, J.; Patterson, L.; McDonald, E.; Workman, P.; Katan, M.; Eccles, S.A. PLC, Phospholipase C, PLC-gamma, The identification of novel PLC-γ inhibitors using virtual high throughput screening. Bioorg. Med. Chem. 2009, 17, 3169–3176. [Google Scholar] [CrossRef]
  91. Leung, E.; Hung, J.M.; Barker, D.; Reynisson, J. The effect of a thieno[2,3-b]pyridine PLC-γ inhibitor on the proliferation, morphology, migration and cell cycle of breast cancer cells. MedChemComm 2014, 5, 99–106. [Google Scholar] [CrossRef]
  92. Ostanin, K.; Hunsaker, T. Enzyme Assay and Use Thereof. US Patent 7338775B1, 4 March 2008. Current Assignee: Myrexis Inc.. [Google Scholar]
  93. Lockman, J.W.; Reeder, M.D.; Suzuki, K.; Ostanin, K.; Willardsen, J.A. Inhibition of eEF2-K by thieno[2,3-b]pyridine analogues. Bioorg. Med. Chem. Lett. 2010, 20, 2283–2286. [Google Scholar] [CrossRef]
  94. Guo, Y.; Zhao, Y.; Wang, G.; Chen, Y.; Jiang, Y.; Ouyang, L.; Liu, B. Design, synthesis and structure–activity relationship of a focused library of β-phenylalanine derivatives as novel eEF2K inhibitors with apoptosis-inducing mechanisms in breast cancer. Eur. J. Med. Chem. 2018, 143, 402–418. [Google Scholar] [CrossRef]
  95. Sun, D.; Zhu, L.; Zhao, Y.; Jiang, Y.; Chen, L.; Yu, Y.; Ouyang, L. Fluoxetine induces autophagic cell death via eEF2K-AMPK-mTOR-ULK complex axis in triple negative breast cancer. Cell Prolif. 2018, 51, e12402. [Google Scholar] [CrossRef] [Green Version]
  96. Soltoff, S.P. Rottlerin: An inappropriate and ineffective inhibitor of PKCδ. Trends Pharmacol. Sci. 2007, 28, 453–458. [Google Scholar] [CrossRef]
  97. Akar, U.; Ozpolat, B.; Mehta, K.; Fok, J.; Kondo, Y.; Lopez-Berestein, G. Tissue transglutaminase inhibits autophagy in pancreatic cancer cells. Mol. Cancer Res. 2007, 5, 241–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. McCracken, M.A.; Miraglia, L.J.; McKay, R.A.; Strobl, J.S. Protein kinase C delta is a prosurvival factor in human breast tumor cell lines. Mol. Cancer Ther. 2003, 2, 273–281. [Google Scholar] [PubMed]
  99. Lim, J.H.; Woo, S.M.; Min, K.J.; Park, E.J. Rottlerin induces apoptosis of HT29 colon carcinoma cells through NAG-1 upregulation via an ERK and p38 MAPK-dependent and PKC δ-independent mechanism. Chem. Biol. Interact. 2012, 197, 1–7. [Google Scholar] [CrossRef]
  100. Clark, A.S.; West, K.A.; Blumberg, P.M.; Dennis, P.A. Altered protein kinase C (PKC) isoforms in non-small cell lung cancer cells: PKCdelta promotes cellular survival and chemotherapeutic resistance. Cancer Res. 2003, 63, 780–786. [Google Scholar]
  101. Ni, H.; Ergin, M.; Tibudan, S.S.; Denning, M.F.; Izban, K.F.; Alkan, S. Protein kinase C-delta is commonly expressed in multiple myeloma cells and its downregulation by rottlerin causes apoptosis. Br. J. Haematol. 2003, 121, 849–856. [Google Scholar] [CrossRef]
  102. Gschwendt, M.; Kittstein, W.; Marks, F. Elongation factor-2 kinase: Effective inhibition by the novel protein kinase inhibitor rottlerin and relative insensitivity towards staurosporine. FEBS Lett. 1994, 338, 85–88. [Google Scholar] [CrossRef] [Green Version]
  103. Soltoff, S.P. Rottlerin Is a Mitochondrial Uncoupler That Decreases Cellular ATP Levels and Indirectly Blocks Protein Kinase Cδ Tyrosine Phosphorylation. J. Biol. Chem. 2001, 276, 37986–37992. [Google Scholar] [CrossRef] [PubMed]
  104. Xu, S.Z. Rottlerin induces calcium influx and protein degradation in cultured lenses independent of effects on protein kinase C delta. Basic Clin. Pharmacol. Toxicol. 2007, 101, 459–464. [Google Scholar] [CrossRef]
  105. Parmer, T.G.; Ward, M.D.; Hait, W.N. Effects of rottlerin, an inhibitor of calmodulin-dependent protein kinase III, on cellular proliferation, viability, and cell cycle distribution in malignant glioma cells. Cell Growth Differ. 1997, 8, 327–334. [Google Scholar] [PubMed]
  106. Ohno, I.; Eibl, G.; Odinokova, I.; Edderkaoui, M.; Damoiseaux, R.D.; Yazbec, M.; Abrol, R.; Goddard, W.A.; Yokosuka, O.; Pandol, S.J.; et al. Rottlerin stimulates apoptosis in pancreatic cancer cells through interactions with proteins of the Bcl-2 family. Am. J. Physiol. Gastrointest. Liver Physiol. 2009, 298, G63–G73. [Google Scholar] [CrossRef] [Green Version]
  107. Mansour, M.A.; Nagi, M.N.; El-Khatib, A.S.; Al-Bekairi, A.M. Effects of thymoquinone on antioxidant enzyme activities, lipid peroxidation and DT-diaphorase in different tissues of mice: A possible mechanism of action. Cell Biochem. Funct. 2002, 20, 143–151. [Google Scholar] [CrossRef]
  108. Banerjee, S.; Padhye, S.; Azmi, A.; Wang, Z.; Philip, P.A.; Kucuk, O.; Sarkar, F.H.; Mohammad, R.M. Review on molecular and therapeutic potential of thymoquinone in cancer. Nutr. Cancer 2010, 62, 938–946. [Google Scholar] [CrossRef] [PubMed]
  109. Asaduzzaman Khan, M.; Tania, M.; Fu, S.; Fu, J. Thymoquinone, as an anticancer molecule: From basic research to clinical investigation. Oncotarget 2017, 8, 51907–51919. [Google Scholar] [CrossRef] [Green Version]
  110. Ballout, F.; Monzer, A.; Fatfat, M.; Ouweini, H.E.; Jaffa, M.A.; Abdel-Samad, R.; Darwiche, N.; Abou-Kheir, W.; Gali-Muhtasib, H. Thymoquinone induces apoptosis and DNA damage in 5-Fluorouracil-resistant colorectal cancer stem/progenitor cells. Oncotarget 2020, 11, 2959–2972. [Google Scholar] [CrossRef] [PubMed]
  111. AlGhamdi, A.A.; Mohammed, M.R.S.; Zamzami, M.A.; Al-Malki, A.L.; Qari, M.H.; Khan, M.I.; Choudhry, H. Untargeted metabolomics identifies key metabolic pathways altered by thymoquinone in leukemic cancer cells. Nutrients 2020, 12, 1792. [Google Scholar] [CrossRef] [PubMed]
  112. Arafa el, S.A.; Zhu, Q.; Shah, Z.I.; Wani, G.; Barakat, B.M.; Racoma, I.; El-Mahdy, M.A.; Wani, A.A. Thymoquinone up-regulates PTEN expression and induces apoptosis in doxorubicin-resistant human breast cancer cells. Mutat. Res. 2011, 706, 28–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Akter, Z.; Ahmed, F.R.; Tania, M.; Khan, M.A. Targeting inflammatory mediators: An anticancer mechanism of thymoquinone action. Curr. Med. Chem. 2021, 28, 80–92. [Google Scholar] [CrossRef]
  114. Ahmad, A.; Mishra, R.K.; Vyawahare, A.; Kumar, A.; Rehman, M.U.; Qamar, W.; Khan, A.Q.; Khan, R. Thymoquinone (2-Isoprpyl-5-methyl-1, 4-benzoquinone) as a chemopreventive/anticancer agent: Chemistry and biological effects. Saudi Pharm. J. 2019, 27, 1113–1126. [Google Scholar] [CrossRef]
  115. Kabil, N.; Bayraktar, R.; Kahraman, N.; Mokhlis, H.A.; Calin, G.A.; Lopez-Berestein, G.; Ozpolat, B. Thymoquinone inhibits cell proliferation, migration, and invasion by regulating the elongation factor 2 kinase (eEF-2K) signaling axis in triple-negative breast cancer. Breast Cancer Res. Ther. 2018, 171, 593–605. [Google Scholar] [CrossRef]
  116. Ishibashi, M. Isolation of bioactive natural products from myxomycetes. Med. Chem. 2005, 1, 575–590. [Google Scholar] [CrossRef] [PubMed]
  117. Li, T.; Wang, N.; Zhang, T.; Zhang, B.; Sajeevan, T.P.; Joseph, V.; Armstrong, L.; He, S.; Yan, X.; Naman, C.B. A systematic review of recently reported marine derived natural product kinase inhibitors. Mar. Drugs 2019, 17, 493. [Google Scholar] [CrossRef] [Green Version]
  118. Bronstrup, M.; Sasse, F. Natural products targeting the elongation phase of eukaryotic protein biosynthesis. Nat. Prod. Rep. 2020, 37, 752–762. [Google Scholar] [CrossRef] [PubMed]
  119. Muthukumar, Y.; Roy, M.; Raja, A.; Taylor, R.E.; Sasse, F. The marine polyketide myriaporone 3/4 stalls translation by targeting the elongation phase. Chembiochem 2013, 14, 260–264. [Google Scholar] [CrossRef]
  120. Yamada, T.; Iwamoto, C.; Yamagaki, N.; Yamanouchi, T.; Minoura, K.; Yamori, T.; Uehara, Y.; Andoh, T.; Umemura, K.; Numata, A. Leptosins M-N1, cytotoxic metabolites from a Leptosphaeria species separated from a marine alga. Structure determination and biological activities. Tetrahedron 2002, 58, 479–487. [Google Scholar] [CrossRef]
  121. Zhang, C.; Lei, J.L.; Zhang, H. Calyxin Y sensitizes cisplatin-sensitive and resistant hepatocellular carcinoma cells to cisplatin through apoptotic and autophagic cell death via SCF βTrCP-mediated eEF2K degradation. Oncotarget 2017, 8, 70595–70616. [Google Scholar] [CrossRef]
  122. Pan, Z.; Chen, Y.; Liu, J.; Jiang, Q.; Yang, S.; Guo, L.; He, G. Design, synthesis, and biological evaluation of polo-like kinase 1/eukaryotic elongation factor 2 kinase (PLK1/EEF2K) dual inhibitors for regulating breast cancer cells apoptosis and autophagy. Eur. J. Med. Chem. 2018, 144, 517–528. [Google Scholar] [CrossRef]
  123. Cao, M.J.; Zhu, T.; Liu, J.T.; Ouyang, L.; Lin, H.W. New sorbicillinoid derivatives with GLP-1R and eEF2K affinities from a sponge-derived fungus Penicillium chrysogenum 581F1. Nat. Prod. Res. 2020, 34, 2880–2886. [Google Scholar] [CrossRef] [PubMed]
  124. Yang, J.; Yang, J.M.; Iannone, M.; Shih, W.J.; Lin, Y.; Hait, W.N. Disruption of the EF-2 kinase/Hsp90 protein complex: A possible mechanism to inhibit glioblastoma by geldanamycin. Cancer Res. 2001, 61, 4010–4016. [Google Scholar] [PubMed]
  125. Khaledian, B.; Taguchi, A.; Shin-Ya, K.; Kondo-Ida, L.; Kagaya, N.; Suzuki, M.; Kajino, T.; Yamaguchi, T.; Shimada, Y.; Takahashi, T. Inhibition of heat shock protein 90 destabilizes receptor tyrosine kinase ROR1 in lung adenocarcinoma. Cancer Sci. 2021, in press. [Google Scholar] [CrossRef] [PubMed]
  126. Talaei, S.; Mellatyar, H.; Asadi, A.; Akbarzadeh, A.; Sheervalilou, R.; Zarghami, N. Spotlight on 17-AAG as an Hsp90 inhibitor for molecular targeted cancer treatment. Chem. Biol. Drug Des. 2019, 93, 760–786. [Google Scholar] [CrossRef] [PubMed]
  127. Sivakumar, K.C.; Haixiao, J.; Naman, C.B.; Sajeevan, T.P. Prospects of multitarget drug designing strategies by linking molecular docking and molecular dynamics to explore the protein–Ligand recognition process. Drug Dev. Res. 2020, 81, 685–699. [Google Scholar] [CrossRef] [PubMed]
  128. Korner, M.; Christ, E.; Wild, D.; Reubi, J.C. Glucagon-like peptide-1 receptor overexpression in cancer and its impact on clinical applications. Front. Endocrinol. 2012, 3, 158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Apel, A.; Zentgraf, H.; Büchler, M.W.; Herr, I. Autophagy-A double-edged sword in oncology. Int. J. Cancer 2009, 125, 991–995. [Google Scholar] [CrossRef] [PubMed]
  130. Hong-Brown, L.Q.; Kazi, A.A.; Lang, C.H. Mechanisms mediating the effects of alcohol and HIV anti-retroviral agents on mTORC1, mTORC2 and protein synthesis in myocytes. World J. Biol. Chem. 2012, 3, 110–120. [Google Scholar] [CrossRef]
  131. Hong-Brown, L.Q.; Brown, C.R.; Huber, D.S.; Lang, C.H. Lopinavir impairs protein synthesis and induces eEF2 phosphorylation via the activation of AMP-activated protein kinase. J. Cell. Biochem. 2008, 105, 814–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Xu, X.; Ong, Y.K.; Wang, Y. Role of adjunctive treatment strategies in COVID-19 and a review of international and national clinical guidelines. Mil. Med. Res. 2020, 7, 22. [Google Scholar] [CrossRef] [PubMed]
  133. Khan, A.A.; Dace, D.S.; Ryazanov, A.G.; Kelly, J.; Apte, R.S. Resveratrol regulates pathologic angiogenesis by a eukaryotic elongation factor-2 kinase-regulated pathway. Am. J. Pathol. 2010, 177, 481–492. [Google Scholar] [CrossRef] [PubMed]
  134. Wang, N.; Feng, Y.; Tan, H.Y.; Cheung, F.; Hong, M.; Lao, L.; Nagamatsu, T. Inhibition of eukaryotic elongation factor-2 confers to tumor suppression by a herbal formulation Huanglian-Jiedu decoction in human hepatocellular carcinoma. J. Ethnopharmacol. 2015, 164, 309–318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The effects that some typical tumor microenvironmental conditions have on eEF2K.
Figure 1. The effects that some typical tumor microenvironmental conditions have on eEF2K.
Ijms 22 02408 g001
Figure 2. Some pathway effects of eEF2K on apoptosis in tumor cells.
Figure 2. Some pathway effects of eEF2K on apoptosis in tumor cells.
Ijms 22 02408 g002
Figure 3. The regulatory effects of eEF2K on the cell cycle, especially prevalent in cancers.
Figure 3. The regulatory effects of eEF2K on the cell cycle, especially prevalent in cancers.
Ijms 22 02408 g003
Figure 4. The structures of eEF2K inhibitors 1 and 2.
Figure 4. The structures of eEF2K inhibitors 1 and 2.
Ijms 22 02408 g004
Scheme 1. General synthesis of 2 and related analogues.
Scheme 1. General synthesis of 2 and related analogues.
Ijms 22 02408 sch001
Scheme 2. General synthesis of pyrido[2,3-d]pyrimidine-2,4-dione analogues.
Scheme 2. General synthesis of pyrido[2,3-d]pyrimidine-2,4-dione analogues.
Ijms 22 02408 sch002
Figure 5. The structures of active pyrido[2,3-d]pyrimidine-2,4-diones 5, 10, and the related PROTAC lead, 11.
Figure 5. The structures of active pyrido[2,3-d]pyrimidine-2,4-diones 5, 10, and the related PROTAC lead, 11.
Ijms 22 02408 g005
Scheme 3. General synthesis of 5,6-dihydro-4H-1,3-selenazine analogues (TS series).
Scheme 3. General synthesis of 5,6-dihydro-4H-1,3-selenazine analogues (TS series).
Ijms 22 02408 sch003
Figure 6. Structures of the active and relatively stable 5,6-dihydro-4H-1,3-selenazines, 14 and 15.
Figure 6. Structures of the active and relatively stable 5,6-dihydro-4H-1,3-selenazines, 14 and 15.
Ijms 22 02408 g006
Figure 7. The structures of thieno[2,3-b]pyridine analogues.
Figure 7. The structures of thieno[2,3-b]pyridine analogues.
Ijms 22 02408 g007
Scheme 4. General synthesis of thieno[2,3-b]pyridine analogues.
Scheme 4. General synthesis of thieno[2,3-b]pyridine analogues.
Ijms 22 02408 sch004
Figure 8. The structures of active β-phenylalanine derivatives and analogues, including 22 and 28.
Figure 8. The structures of active β-phenylalanine derivatives and analogues, including 22 and 28.
Ijms 22 02408 g008
Scheme 5. General synthesis of β-phenylalanine analogues.
Scheme 5. General synthesis of β-phenylalanine analogues.
Ijms 22 02408 sch005
Figure 9. The structure of fluoxetine, an approved SSRI drug later evaluated as an eEF2K inhibitor.
Figure 9. The structure of fluoxetine, an approved SSRI drug later evaluated as an eEF2K inhibitor.
Ijms 22 02408 g009
Figure 10. Structures of some natural products that are eEF2K inhibitors (3035).
Figure 10. Structures of some natural products that are eEF2K inhibitors (3035).
Ijms 22 02408 g010
Figure 11. The structures of the PLK1/eEF2K dual-targeting inhibitor, 44.
Figure 11. The structures of the PLK1/eEF2K dual-targeting inhibitor, 44.
Ijms 22 02408 g011
Scheme 6. Synthesis route to 1-(4-(2-substituted-pyridin-4-yl)-3-substituted-phenyl)-3-phenylurea PLK1/eEF2K dual-targeting inhibitors.
Scheme 6. Synthesis route to 1-(4-(2-substituted-pyridin-4-yl)-3-substituted-phenyl)-3-phenylurea PLK1/eEF2K dual-targeting inhibitors.
Ijms 22 02408 sch006
Figure 12. Structures of two active eEF2K inhibiting sorbicillinoid natural products (45 and 46) from the sponge-derived fungus Penicillium chrysogenum.
Figure 12. Structures of two active eEF2K inhibiting sorbicillinoid natural products (45 and 46) from the sponge-derived fungus Penicillium chrysogenum.
Ijms 22 02408 g012
Figure 13. The structures of geldanamycin (47) and a synthetic derivative, 17-AAG (48).
Figure 13. The structures of geldanamycin (47) and a synthetic derivative, 17-AAG (48).
Ijms 22 02408 g013
Figure 14. Structures of the eEF2K activators ritonavir, lopinavir, and resveratrol (4951).
Figure 14. Structures of the eEF2K activators ritonavir, lopinavir, and resveratrol (4951).
Ijms 22 02408 g014
Table 1. Original names and discovery method of reported eEF2K inhibitors.
Table 1. Original names and discovery method of reported eEF2K inhibitors.
CompoundOriginal Name in Published
Report
Discovery Method aDevelopment Status for Cancer Therapy bReference
1NH125KIORecent[73,80,81,82,83]
2TX-1918KIONA[84]
5A-484954HTSRecent[82,85]
10compound 9KIONA[86]
11compound 11lKIORecent[87]
14TS-2KIONA[88,89]
15TS-4KIONA[88,89]
16compound 1HTSNA[90,91,92]
17compound 2HTSNA[90,91,92]
21compound 34KIONA[93]
22compound 9CADDRecent[94]
28compound 21lCADDRecent[94]
29fluoxetineDRRRecent[95]
30rottlerinNPDDNA[96,97,98,99,100,101,102,103,104,105,106]
31thymoquinoneNPDDRecent[107,108,109,110,111,112,113,114,115]
326-hydroxystaurosporinoneNPDDNA[116,117]
33myriaporone 3/4NPDDNA[118,119]
34leptosin MNPDDNA[120]
35calyxin YNPDDRecent[121]
44compound 18iMTDDRecent[122]
4513-hydroxy-dihydro-
trichodermolide
NPDDRecent[123]
4610,11,27,28-tetrahydro-
trisorbicillinone C
NPDDRecent[123]
47geldanamycinNPDDRecent[124,125]
4817-AAGNPDDRecent[124,126]
a CADD = Computer-Assisted Drug Design, DRR = Drug Repurposing/Repositioning, HTS = High-Throughput Screening, KIO = Kinase Inhibitor Optimization, MTDD = Multi-Targeted Drug Design, NPDD = Natural Product Discovery and Derivatization; b Recent = this compound has had literature reports or new developments in cancer research and development published in the past five years, NA = no apparently relevant literature reports or new developments have been found for the same time period.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, B.; Zou, J.; Zhang, Q.; Wang, Z.; Wang, N.; He, S.; Zhao, Y.; Naman, C.B. Progress in the Development of Eukaryotic Elongation Factor 2 Kinase (eEF2K) Natural Product and Synthetic Small Molecule Inhibitors for Cancer Chemotherapy. Int. J. Mol. Sci. 2021, 22, 2408. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052408

AMA Style

Zhang B, Zou J, Zhang Q, Wang Z, Wang N, He S, Zhao Y, Naman CB. Progress in the Development of Eukaryotic Elongation Factor 2 Kinase (eEF2K) Natural Product and Synthetic Small Molecule Inhibitors for Cancer Chemotherapy. International Journal of Molecular Sciences. 2021; 22(5):2408. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052408

Chicago/Turabian Style

Zhang, Bin, Jiamei Zou, Qiting Zhang, Ze Wang, Ning Wang, Shan He, Yufen Zhao, and C. Benjamin Naman. 2021. "Progress in the Development of Eukaryotic Elongation Factor 2 Kinase (eEF2K) Natural Product and Synthetic Small Molecule Inhibitors for Cancer Chemotherapy" International Journal of Molecular Sciences 22, no. 5: 2408. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22052408

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

Article Metrics

Back to TopTop