Next Article in Journal
Production and Application of Stable Isotope-Labeled Internal Standards for RNA Modification Analysis
Previous Article in Journal
The Versatility of SMRT Sequencing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting DNA Double-Strand Break Repair Pathways to Improve Radiotherapy Response

Division of Radiobiology and Molecular Environmental Research, Department of Radiation Oncology, University of Tuebingen, Roentgenweg 11, 72076 Tuebingen, Germany
Genes 2019, 10(1), 25; https://doi.org/10.3390/genes10010025
Submission received: 9 November 2018 / Revised: 7 December 2018 / Accepted: 27 December 2018 / Published: 4 January 2019
(This article belongs to the Special Issue DNA Damage and Repair in Cancer)

Abstract

:
More than half of cancer patients receive radiotherapy as a part of their cancer treatment. DNA double-strand breaks (DSBs) are considered as the most lethal form of DNA damage and a primary cause of cell death and are induced by ionizing radiation (IR) during radiotherapy. Many malignant cells carry multiple genetic and epigenetic aberrations that may interfere with essential DSB repair pathways. Additionally, exposure to IR induces the activation of a multicomponent signal transduction network known as DNA damage response (DDR). DDR initiates cell cycle checkpoints and induces DSB repair in the nucleus by non-homologous end joining (NHEJ) or homologous recombination (HR). The canonical DSB repair pathways function in both normal and tumor cells. Thus, normal-tissue toxicity may limit the targeting of the components of these two pathways as a therapeutic approach in combination with radiotherapy. The DSB repair pathways are also stimulated through cytoplasmic signaling pathways. These signaling cascades are often upregulated in tumor cells harboring mutations or the overexpression of certain cellular oncogenes, e.g., receptor tyrosine kinases, PIK3CA and RAS. Targeting such cytoplasmic signaling pathways seems to be a more specific approach to blocking DSB repair in tumor cells. In this review, a brief overview of cytoplasmic signaling pathways that have been reported to stimulate DSB repair is provided. The state of the art of targeting these pathways will be discussed. A greater understanding of the underlying signaling pathways involved in DSB repair may provide valuable insights that will help to design new strategies to improve treatment outcomes in combination with radiotherapy.

1. DNA Double-Strand Break Repair

Despite advances in radiotherapy, radioresistance remains a major cause of treatment failure, leading to lower progression-free survival rates in cancers such as lung cancer, pancreatic cancer, and glioblastoma. This failure indicates the necessity to investigate the underlying network of signal transduction pathways known as DNA damage response (DDR) pathways, which stimulate the repair of double-strand breaks (DSBs), the most lethal type of DNA damage due to radiotherapy [1]. Following exposure to ionizing radiation (IR), cells undergo transient cell cycle arrest to perform DNA repair. DSBs are repaired either by classical and alternative non-homologous end joining (C-NHEJ and A-NHEJ) throughout the cell cycle or by homologous recombination (HR) during the S phase and G2 phase [2]. After the induction of DSB, the MRE11-Rad50-NBS1 (MRN) complex and K70/80 act as sensor proteins to recognize the lesions [3,4]. Thereafter, the phosphoinositide 3-kinase (PI3K) family members ataxia telangiectasia mutated (ATM), ataxia telangiectasia and Rad3-related (ATR), and DNA-dependent protein kinase catalytic subunit (DNA-PKcs) are activated [5]. The activation of ATM kinase by autophosphorylation relays the signal to the transducer enzymes, including checkpoint kinase 2 (CHK2) and the transcription factor p53. Cellular p53 regulates the expression of cell cycle regulators such as p21 that, through interaction with the cyclin-dependent kinase (CDK) complex, lead to G1 arrest. Along with this process, chromatin modification occurs, and DNA repair is initiated [6]. In most cancer cells, the G1 checkpoint is defective because of mutations in the key regulators of the G1 checkpoint, such as p53. Therefore, the G2 checkpoint becomes important for cancer cell survival. ATR phosphorylates checkpoint kinase 1 (CHK1). The activation of CHK1 mediates the phosphorylation of cell division cycle 25A (CDC25A) and consequently its degradation, which slows the progression of DNA replication through the S phase. Likewise, CHK1 phosphorylates CDC25C, which stops the cell cycle in the G2 phase [7,8]. After arresting in different cell cycles, cells apply the canonical DSB repair pathways to repair DSBs before cell cycle re-entry. Thus, the underlying components involved in cell cycle arrest and canonical DSB repair pathways can be targeted to block DSB repair and induce radiosensitivity. More details on the canonical DSB repair pathways have been presented in review articles by other investigators [4,9,10,11].
Tumor-specific defects in DDR provide new options for cancer therapy. A significant fraction of cancers have inherited and acquired defects in one of the DNA repair pathways [5,12]. One example is the defect in the HR repair pathway caused by mutations in the tumor suppressor genes BRCA1 and BRCA2, which are related to breast and ovarian cancer. Mutation in the p53 gene, which is referred to as the guardian of the genome and is most commonly mutated in human cancers, is another example of tumor-specific DDR defects. DNA damage repair in cancer cells that have lost one DNA repair component usually requires other repair pathways to allow the cells to survive after therapy. Thus, targeting these rescue repair pathways in cancer cells with repair deficiency, known as synthetic lethality, might be an effective therapeutic approach for such cancers [13]. Synthetic lethality is the underlying mechanism by which BRCA1/2-defective tumor cells are sensitive to poly(ADP-ribose) polymerase (PARP) inhibitors. Likewise, in p53-mutated cells with defects in the G1/S checkpoint, targeting the G2/M checkpoint becomes a very effective therapy to sensitize p53 mutant cells to DNA damage-inducing agents such as ionizing radiation (IR) during radiotherapy. Application of the concept of synthetic lethality to radiotherapy may be an effective approach to enhance the therapeutic ratio by minimizing normal-tissue side effects [14].
Exposure to IR induces the activation of cytoplasmic signaling cascades, such as the PI3K/AKT, RAS-mitogen-activated protein kinase (MAPK), signal transducer and activator of transcription (STAT) and phospholipase C (PLC) pathways [15,16]. According to the functions of such pathways, the radiation-induced activation of these signaling cascades accelerates cell proliferation and the post-irradiation survival of tumor cells [17,18,19,20]. In addition to the well-described role of the cytoplasmic signaling pathway in regulating multiple cellular functions, the activation of these pathways stimulates the repair of radiation-induced DSBs by regulating the expression and/or activation of the components of the canonical DSB repair pathways. It is notable that the described functions of these pathways in DSB repair and radioresistance are not solely dependent on IR-induced stimulation of the underlying pathways. These cascades can also become hyperactivated in tumor cells expressing mutations or overexpressing certain oncogenes or tumor suppressor genes, elevating the DSB repair capacity of these cells. Thus, as depicted in Figure 1, this mode of action in DSB repair may specifically lead to the radioresistance of tumor cells. Likewise, it provides a rationale for mutational screening and expression analyses of previously described oncogenes before radiotherapy. The knowledge obtained from such analyses will help to design appropriate molecular therapy strategies in combination with radiotherapy as part of personalized therapy.
In the present paper, the relevant literature concerning the activation of canonical DSB repair pathways through cytoplasmic signaling cascades is summarized. The information regarding the described signaling cascades is accompanied by a description of proof-of-principle radiosensitization by targeting the components of the described pathway.

2. Receptor Tyrosine Kinases That Mediate DSB Repair after Irradiation

Membrane-bound receptor tyrosine kinases (RTKs) consist of 58 members within 20 subfamilies. Mutation or overexpression of these receptors has been reported in more than 65% of these receptor families in different cancers. These alterations result in deregulated kinase activity and malignant transformation [21] and are associated with poor prognosis, drug resistance, cancer metastasis, and lower survival rate. In addition to the contributions of direct genomic events to the expression and hyperactivation of RTKs, epigenetic modifications of RTKs also modulate RTK activity [22,23].

2.1. ErbB Family of RTKs

The ErbB receptor family, also described as the epidermal growth factor (EGFR) family, includes EGFR or ErbB1/Her1, ErbB2/Her2, ErbB3/Her3, and ErbB4/Her4. These family members are activated after homodimerization or heterodimerization. No ligand has been described for ErbB2, and ErbB3 has impaired kinase activity. Among the RTKs, nuclear localization of the ErbB receptor family members EGFR, HER2, and HER3 has been reported. The first step in the function of RTKs in DSB repair is the nuclear localization of the receptors. Potential mechanisms involved in the nuclear translocation of these receptors, including endosome-mediated nuclear translocation and retro-translocation by the endoplasmic reticulum, have been shown and reviewed elsewhere [24]. RTKs stimulate DNA repair through several mechanisms. In this context, a critical role of EGFR acetylation has been reported to regulate tyrosine phosphorylation of the receptor and its function [23]. Likewise, nuclear EGFR can directly interact with DSB repair proteins such as DNA-PKcs, a core enzyme in the NHEJ repair pathway that stimulates DSB repair after irradiation [25,26,27]. Radiation may also stimulate EGFR activity in the nucleus without requiring translocation from the cytoplasm to the nucleus, although this issue has not yet been investigated (Figure 2A). Stimulating cells with EGFR ligands, e.g., EGF and transforming growth factor α (TGFα), leads to the translocation of EGFR to the nucleus [28]. The expression of oncogenic KRAS stimulates autocrine secretion of TGFα and another EGFR ligand, amphiregulin [29,30,31]. Because the overexpression of oncogenic KRAS leads to enhanced radiation-induced autophosphorylation of DNA-PKcs at S2056 and stimulates DSB repair [29], it is supposed that the translocation of activated EGFR to the nucleus facilitates radiation-induced DNA-PKcs activity and stimulates DSB repair through NHEJ. According to the model outlined in Figure 2B, nuclear EGFR was proposed to play an efficient role in DSB repair through NHEJ in tumor cells expressing oncogenic KRAS.
EGFR also plays a key role in the repair of DSBs through HR. Nowsheen et al. showed that EGFR, in a protein complex with BRCA1, plays an essential role in DSB repair through HR [32]. Lee et al. showed that, upon irradiation, ATM associates with and is phosphorylated by EGFR at Tyr-370 at the site of DSBs. Likewise, EGFR regulates the phosphorylation of ATM at Tyr-370, leading to Chk2 activity [33], which can affect G1 cell cycle arrest, a prerequisite for NHEJ. Additionally, Wang and colleagues showed that chromatin-bound proliferating cell nuclear antigen (PCNA) phosphorylation at Tyr-211 is dependent on the tyrosine kinase activity of EGFR in the nucleus. Phosphorylation at Tyr-211 by EGFR stabilizes the chromatin-bound PCNA protein and its associated functions [34]. EGFR also mediates the Tyr-72 phosphorylation of histone 4 (H4), which leads to the Lys-20 methylation of H4 and the acceleration of DNA synthesis and repair [35]. In the context of epigenetic regulation by RTKs, it was also reported that EGFR acetylation plays a role in the function of the receptor, leading to tumor cell resistance to histone deacetylase inhibitors [23]. These and other lines of evidence strongly indicate that EGFR regulates DDR and stimulates DSB repair through both NHEJ and HR, as summarized in Figure 2. Consistent with the role of EGFR in regulating the DDR response, i.e., stimulating DSB repair through HR and NHEJ [36,37,38,39], EGFR expression is inversely correlated with the IR sensitivity [40] and overall survival of patients after radiotherapy [41].

2.2. Targeting the ErbB Family of RTKs in Combination with Radiotherapy

The ErbB family of receptors has been considered the most important target in oncology [42,43] as well as in radiation oncology [44]. Among the molecular targeting approaches, inhibitors against EGFR, i.e., monoclonal antibodies and tyrosine kinase inhibitors (TKIs), have been thoroughly investigated in several preclinical and clinical settings, as reviewed by Huang et al. [45]. As a proof of concept, the anti-EGFR antibody cetuximab has been shown to improve radiation sensitivity in head and neck squamous cell carcinoma (HNSCC) in multicenter phase III clinical trials [46,47]. To date, the majority of studies performed with anti-EGFR antibodies have been conducted using cetuximab. Most of these clinical trials have been performed in HNSCC patients in combination with chemoradiotherapy. However, as summarized in Table 1, only a single phase III trial for the combination of cetuximab with radiotherapy led to improved overall survival in HNSCC patients [46,47]. The combination of cetuximab with chemoradiotherapy was not positive and sometimes even showed inferior outcome (Table 1) The combination of cetuximab with chemoradiotherapy for other tumor entities was optimized according to the aim of each study [48,49,50]. However, treatment-associated toxicity is often a limiting factor that must be considered [51,52]. The beneficial impact of EGFR TKIs has been mainly investigated by the combination of erlotinib and gefitinib with radiotherapy or chemoradiotherapy in non-small cell lung cancer (NSCLC) as well as HNSCC. Some of the early phase clinical trials combining erlotinib and radiotherapy or chemoradiotherapy are summarized and presented in Table 1.
Radiation-induced DSB is the major mechanism by which radiotherapy induces cell death. Thus, interference with the DSB repair pathway may be the crucial mechanism by which cetuximab improves radiotherapy outcomes. Additionally, EGFR targeting enhances IR-induced apoptosis, inhibits proliferative growth, and downregulates tumor angiogenic response [63].
Although the nuclear localization of Her2 and Her3 has been demonstrated previously [64,65,66], the specific function of these receptors in DSB repair is largely unknown. Consistent with the role of nuclear Her3 in radioresistance [67], Reif et al. showed that the level of nuclear Her3 increases upon hypoxia [66]. Because hypoxia is associated with radioresistance, it may be concluded that the enhanced nuclear translocation of Her3 is linked to DSB repair.

2.3. IGF-1 Receptor

The membrane-bound type 1 insulin growth factor receptor (IGF-1R) is a tyrosine kinase receptor that is stimulated by IGF-1, a polypeptide protein hormone similar in molecular structure to insulin. IGF-1R signaling contributes to the growth of many solid tumors, such as glioblastoma [68], with poor prognosis. IGF-1R is the second most studied receptor whose roles in DSB repair and radioresistance have been reported. Nuclear localization of IGF-1R has been well demonstrated by several investigators [69,70] and is associated with therapy resistance. Likewise, high levels of total or cytoplasmic IGF-1R expression reveal an increased risk of postradiotherapy recurrence in prostate cancer patients [71], indicating that radioresistance is mediated by IGF-1R.

2.4. Targeting IGF-1R in Combination with Radiotherapy

Because suppression of HR sensitizes human tumor cells to IGF-1R inhibition, IGF-1R might be crucial for HR repair of DSBs [72]. In addition to its role in HR, IGF-1R also stimulates NHEJ repair of DSB [73,74]. These results and other evidence of the role of IGF-1R in DSB repair support the idea that this receptor can potentially be an appropriate target in combination with radiotherapy. To date, several studies combining IGF-1R targeting with IR have been reported. Similar to the EGFR targeting strategies, monoclonal antibodies and RTKs have been investigated as potential approaches to block IGF-1R in combination with radiotherapy. The anti-IGF-1R monoclonal antibodies A12 (ImClone Systems, Inc., New York, NY, USA) and CP-751,871 enhance the radiosensitivity of NSCLC via the inhibition of DSB repair [75,76]. In addition to the monoclonal antibodies, the radiosensitizing effect of IGF-1R TKIs has also been investigated. Isebaert et al. demonstrated that the IGF-1R TKI NVP-AEW541 enhances radiosensitivity in phosphatase and tensin homolog (PTEN) wild-type but not PTEN-deficient human prostate cancer cells [77]. NVP-AEW541-induced radiosensitization was associated with downregulation of phospho-AKT levels and high levels of residual DSB [77], indicating a role for IGF-1R in stimulating DSB repair pathways. Chitnis et al. demonstrated that the radiosensitizing effect of the IGF-1R TKI AZ12253801 depends on the expression of DNA-PKcs [73]. AZ12253801 induced radiosensitization in DNA-PKcs-proficient but not DNA-PKcs-deficient glioblastoma cells and did not radiosensitize DNA-PKcs-inhibited DU145 prostate cells [73]. These results suggest that IGF-1R functions in the same pathway as DNA-PKcs, which stimulates the C-NHEJ repair of DSB repair [73]. The radiosensitizing effect of IGF-1R molecular targeting strategies has also been supported by genetic approaches. Zhao and Gu reported that the silencing of IGF-1R enhances the radiation sensitivity of human esophageal squamous cell carcinoma, leading to tumor growth delay and prolonged survival after radiotherapy in a tumor xenograft model in vivo [78]. IGF-1R has a functional interaction with EGFR [79]. This interaction stimulates the activation of common downstream signaling cascades, e.g., the PI3K/AKT pathway. Consistent with this conclusion, investigation of co-targeting IGF-1R and EGFR in gastric cancers has been suggested [80] and might be an efficient approach to improve radiotherapy outcome. Interference with the PI3K/AKT pathway is one of the potential mechanisms by which targeting IGF-1R induces radiosensitization [77]. However, because the PI3K/AKT pathway is a common pathway downstream of all RTK members, including EGFR, the long-term inhibition of IGF-1R may lead to the activation of this pathway through EGFR. This compensatory reactivation of the PI3K/AKT pathway may diminish the radiosensitizing effect achieved by targeting IGF-1R.

2.5. TAM Family of Receptors

The TAM family of RTKs consists of three members: TYRO-3, AXL, and MERTK. TAM receptors play an important role in promoting the growth, survival, and metastatic spread of several tumor types. AXL and MERTK have been described to be overexpressed in HNSCC, triple-negative breast cancer (TNBC), and NSCLC, malignancies that are highly metastatic and lethal [81]. The level of TAM receptor expression correlates with the tumor grade and emergence of chemo- and radioresistance to targeted therapeutics [82]. The AXL receptor and its activating ligand, growth arrest-specific 6 (GAS6), are important drivers of metastasis and therapeutic resistance [83,84]. Thus, based on the stimulatory role of AXL in DNA repair and therapy resistance, targeting AXL might be an effective approach to improve radiotherapy outcome.

2.6. Targeting AXL in Combination with Radiotherapy

The downregulation or inhibition of AXL decreases the expression of DNA repair genes and diminishes the efficiency of HR [85]. Consistent with the function of AXL in DSB repair, the expression of AXL is highly associated with the radiation resistance of human papilloma virus (HPV)-negative HNSCC cells, dependent on the activation of PI3K and the expression of programmed death-ligand 1 (PD-L1) [86]. Based on the role of AXL in DNA repair and radioresistance [85,86], targeting AXL is supposed to be an efficient approach to block DSB repair and induce radiosensitization. An anti-AXL antibody and small molecule TKIs of AXL have been investigated in preclinical trials, and some have entered clinical trials. The antitumor effectiveness of the AXL small-molecule inhibitors TP-0903 [87] and R428 [88] has been tested in hematological malignancies and solid tumors [87,89,90]. However, cancer treatment by targeting AXL is a relatively new field in oncology, and thus far, few investigations have been conducted in combination with radiotherapy. In this context, Brand et al. demonstrated that the AXL inhibitor R428 blocks DSB repair and induces radiosensitivity [91]. Additionally, because AXL mediates the nuclear localization of EGFR and resistance to anti-EGFR therapy [92], co-targeting of AXL and EGFR might be an efficient approach to induce radiosensitization, a strategy that needs to be investigated.

3. Cytoplasmic Signaling Cascades That Stimulate DSB Repair

In addition to the direct function of RTKs in DSB repair as described above, this family of receptors can also stimulate the repair of DSB by activating various cytoplasmic signaling pathways, such as the PI3K/AKT pathway, STAT pathway, and RAS-MAPK pathway. Among the different pathways downstream of RTKs, the PI3K/AKT pathway is the best-studied pathway regulating DNA damage repair. This pathway is one of the major survival pathways and is frequently upregulated in tumors from different entities [93,94,95,96].

3.1. Stimulation of DSB Repair by the PI3K/AKT Pathway

Mutations in RTKs, PTEN, PI3K, AKT, and RAS are the major mutations involved in the constitutive activation of the PI3K/AKT pathway. Additionally, exposure to IR induces the activation of this pathway through the stimulation of upstream RTKs, e.g., EGFR [17,97,98,99,100,101]. Targeting PI3K induces radiosensitization in tumor cells from different entities in vitro and in vivo [17,97,100,102,103,104,105,106]. Initial studies using the PI3K inhibitor LY294002 showed a radiosensitizing effect through interference with DSB repair pathways, i.e., HR and NHEJ [107,108,109].
Among the three classes of PI3K, the class IA isoform of PI3K has been strongly implicated in cancer. This PI3K isoform consists of a p85 regulatory subunit and a p110 catalytic subunit. Despite the therapeutic benefit of targeting this isoform, there have been concerns about the severe adverse effects of this class of drug that have led to difficulties in the application of PI3K inhibitors in combination with radiotherapy [110]. PI3K activity regulates various substrates that are involved in different cellular functions, such as cell cycle progression, different types of cell death, glycolysis, and DNA repair [93,111]. Among the many PI3K substrates, AKT, also known as protein kinase B (PKB), is the key mediator of the PI3K signaling pathway [112]. AKT is involved in the repair of radiation-induced DSBs through both HR and NHEJ [99,107,109,113,114,115,116,117]. Consistent with the preclinical observations, the clinical data support the role of AKT in radioresistance. In this context, AKT activity was reported to be a prognostic marker for the radiotherapy response of head and neck cancer and cervical cancer patients [118,119]. The PI3K-independent reactivation of AKT has been reported in KRAS-mutated NSCLC cells [120,121,122], glioblastoma cells [123], bladder cancer cells [124], and TNBC cells [125] after the blockade of PI3K. Thus, in addition to dose-limiting toxicities, the feedback activation of AKT can be an obstacle for targeting PI3K and limits the success of combining PI3K inhibitors with radiotherapy.
AKT consists of three isoforms—AKT1/PKBα, AKT2/PKBβ, and AKT3/PKBγ —transcribed from separate genes. These AKT isoforms have an N-terminal pleckstrin homology (PH) domain and a kinase domain, which are separated by 39 amino acids [126]. The PH domains and kinase domains in the AKT isoforms are approximately 60% and 85% identical, respectively [127]. Activated AKT regulates the function of numerous substrates in cell growth, proliferation, and survival, as well as regulating metabolism, angiogenesis, and migration [128]. Among the different AKT isoforms, AKT1 was the first isoform that was described to directly interact with DNA-PKcs [116,129] through its C-terminal domain [116]. Furthermore, we showed that, similar to AKT1, AKT3 also interacts with DNA-PKcs [130]. No interaction with DNA-PKcs could be observed for AKT2 in KRAS-mutated NSCLC [130]. Thus, AKT activates DNA-PKcs kinase activity and its phosphorylation at Thr-2609 and Ser-2056, which is essential for DSB repair by NHEJ [131,132]. Activated AKT localizes to DSB sites, as shown by the colocalization of γH2AX foci as a marker of DSB with P-AKT (Ser-473) after irradiation [115,133,134], and stimulates DSB repair, leading to radioresistance [116,130,135,136,137,138]. Accumulating evidence has indicated the role of AKT in stimulating HR. My colleagues and I recently showed that AKT1 knockdown reduces Rad51 protein level, Rad51 foci formation, and its colocalization with H2AX foci after irradiation [117]. These events were associated with unsuccessful HR, as shown by increased BRCA1 foci 24 h post-irradiation [117]. Consistent with the role of AKT in HR, it was shown that, in PTEN-mutated cells with the hyperactivation of PI3K, PI3K inhibition reduces RAD51 foci formation and sensitizes these cells to the PARP inhibitor [139].
AKT, through the GSK3β/β-catenin/LEF pathway, upregulates the expression of Mre11 in the MRN complex and elevates DSB repair capacity [140]. The MRN complex recruits ATM to the DSB sites, where ATM is subsequently activated. Activated ATM, similar to phosphorylating members of the MRN complex, stimulates the phosphorylation of AKT at Ser-473 [141] through RNF168 [115]. Given the role of AKT in MRE11 expression [140], AKT leads to the expression of MRE11 and stimulation of ATM signaling. Because ATM plays a critical role in regulating HR repair [142], AKT stimulates HR as well. Thus, the ATM-dependent mode of action of AKT in HR might be important to regulate the slow component of DNA repair following the fast NHEJ repair process. AKT in non-irradiated cells is expressed in the nucleus. Thus far, there is no convincing data for IR-induced nuclear translocation of AKT. Because AKT needs to be in the nucleus immediately after irradiation to play its role in regulating DNA repair pathways, it can be assumed that nuclear AKT is activated by PI3K components in the nucleus, independent of the cytoplasmic fraction, although this issue needs to be investigated. Overall, the role of AKT in the stimulation of DSB repair through HR as well as NHEJ makes AKT an effective target in combination with radiotherapy. Some of the most important aspects of the function of AKT in DSB repair are summarized in Figure 3.
The hyperactivation of AKT correlates with various clinicopathological parameters and is a prognostic indicator for cancers from different entities [143,144,145,146]. The mutation and overexpression of oncogenes such as RTK families, RAS, and PIK3CA, as well as mutation in the tumor suppressor gene PTEN, lead to stimulated AKT activity. The constitutive activation of KRAS through point mutation in KRAS gene leads to the activation of the PI3K/AKT and MAPK/ERK pathways, as the major pathways regulating growth proliferation and survival. It is well known that mutation in the KRAS gene leads to radioresistance that is linked to the activation of the PI3K/AKT pathway [29,147]. Thus, it is supposed that KRAS or PI3K may serve as a suitable target to overcome radioresistance. To this end, several laboratories have begun testing the antitumor activity of KRAS mutation-specific inhibitors. Ostrem et al. validated a new allosteric regulatory site on KRAS (G12C) that is targetable in a mutant-specific manner, and the described inhibitor blocks the interaction of KRAS protein with effectors such as B-Raf and C-Raf [148]. However, in that study, the impact of KRAS (G12C) inhibition on PI3K/AKT activity was not investigated. In a further study, Misale et al. [149] demonstrated that the reactivation of the PI3K/AKT pathway limits the efficacy of the KRAS (G12C) inhibitor ARS1620 in some of the cell lines tested. These authors suggested the combination of ARS1620 with PI3K inhibitors [149]. This resistance mechanism indicates the rationale for targeting the PI3K/AKT survival pathway alone as well as in combination with IR. Previously, we showed that short-term treatment (2 h) with the PI3K inhibitor PI-103 blocks the phosphorylation of AKT (Ser-473 and Thr-308) as well as the phosphorylation of the AKT substrate PRAS40 (Thr-246) in NSCLC cell lines A549 and H460 expressing KRAS (G12V) and KRAS (Q61H), respectively. The inhibitory effect of PI-103 disappeared at 24 h post-treatment in both KRAS-mutated cells [121]. In KRAS wild-type H661 cells, PI-103 blocked AKT phosphorylation at the 2-h and 24-h post-treatment time points [120,121]. These new observations and previous reports on the role of oncoprotein RAS, especially KRAS, in activating survival pathways warrant further studies to uncover appropriate targeting strategies to overcome RAS-mediated radioresistance.
Among the three AKT isoforms tested, AKT1 and AKT3 interact with DNA-PKcs and regulate DSB repair through NHEJ [130]. Mutations in AKT isoforms lead to the constitutive activation of AKT. The E17K mutation in the AKT1 gene causes the constitutive membrane localization of AKT1. This results in permanent AKT1 kinase activity and phosphorylation at Thr-308 and Ser-473, as well as the consequent activation of downstream target proteins independent of growth factor stimulation, as reported in different cancers such as breast cancer, endometrial cancer, bladder cancer, lung cancer, and colorectal cancer [150,151,152,153]. The expression of AKT1-E17K accelerates DSB repair and improves post-irradiation cell survival [154]. For full activity, AKT3 needs to be phosphorylated at Ser-472 and Thr-305 [127]. Similar to the AKT1 mutation, the E17K mutation in the AKT3 gene leads to the constitutive activation of AKT3.

3.2. Targeting AKT for Radiosensitization

Based on the role of AKT in DSB repair discussed above and the common disruption of the RTK/PI3K/AKT pathway in human cancers, e.g., the hyperactivation of AKT in over 50% of human tumors, AKT is a suitable target in combination with radiotherapy. Thus far, almost all the relevant studies, particularly clinical trials, have been performed using AKT inhibitors that inhibit all three isoforms. The targeting of AKT can be achieved using either allosteric inhibitors or ATP competitive inhibitors. Allosteric inhibitors prevent the plasma membrane localization of AKT through blocking the PH domain. Thus, AKT will not be recruited to the cell membrane, where it is phosphorylated at its threonine and serine residues by PDK1 and PDK2, respectively. Several small-molecule AKT inhibitors are currently undergoing clinical evaluation. The current status of clinical trials with AKT inhibitors as monotherapy or in combination with chemotherapy has been summarized by other investigators [155].
Preclinical data support the radiosensitizing effect of AKT inhibitors through blocking DSB repair [136,156,157]. Among several AKT inhibitors, phase I trials have been performed using the combination of the AKT inhibitor perifosine with radiotherapy. Phase I and pharmacokinetic studies of combined treatment with perifosine and radiation in patients with advanced solid tumors have indicated that perifosine can be safely combined with fractionated radiotherapy [158]. Nelfinavir is a human immunodeficiency virus (HIV) protease inhibitor that has been shown to possess antitumor activity and radiosensitizing effects through the inhibition of phosphorylated AKT [159]. The combination of nelfinavir and chemoradiotherapy in phase I and II clinical trials showed acceptable toxicity and promising activity in patients with NSCLC, rectal cancers, and pancreatic cancers [160,161,162,163,164] (Table 2), which frequently express KRAS mutations. Overall, thus far, mechanistic in vitro studies and existing clinical trials have supported the rationale for future, more in-depth clinical studies investigating the combination of AKT inhibitors with radiotherapy.

4. Conclusions and Prospects

Cytoplasmic signaling cascades downstream of oncoproteins, such as RTKs and RAS, stimulate canonical DSB repair pathways, i.e., NHEJ and HR. Direct targeting of the canonical pathways, e.g., by applying DNA-PKcs or ATM inhibitors, most likely hampers DSB repair in both normal and tumor cells. This effect leads to a limited therapeutic window due to normal-tissue toxicity. By contrast, targeting DSB repair stimulated by cytoplasmic signaling cascades, known as indirect targeting of DSB repair, is advantageous and seems to be a more tumor-specific approach for radiosensitization. However, the issue of tumor heterogeneity can potentially challenge the efficacy of this approach. This problem can be caused by functional crosstalk between different cytoplasmic signaling pathways or, alternatively, by a compensatory activation loop following the long-term inhibition of a target/pathway. This challenge mainly arises in tumors that harbor the overexpression and/or mutation of several oncogenes. Thus, to choose the appropriate target(s), it is necessary to identify the genetic background of tumors and develop biomarkers to predict the targetability of the components of a pathway that regulates DSB repair. This approach will help to select patients who will potentially benefit from the combination of radiotherapy with specific molecular targeting strategies for the indirect targeting of DSB repair.

Funding

The authors’ research is supported by Deutsche Forschungsgemeinschaft (DFG, German Research Council).

Acknowledgments

I would like to thank H. Peter Rodemann for supporting my research in his laboratory. I also would like to thank Klaus Dittmann, Division of Radiobiology and Molecular Environmental Research, Department of Radiation Oncology, University of Tuebingen and Cihan Gani, Department of Radiation Oncology, University of Tuebingen for their valuable comments to improve the manuscript. I acknowledge support by Deutsche Forschungsgemeinschaft and Open Access Publishing Fund of University of Tuebingen.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Khanna, K.K.; Jackson, S.P. DNA double-strand breaks: Signaling, repair and the cancer connection. Nat. Genet. 2001, 27, 247–254. [Google Scholar] [CrossRef] [PubMed]
  2. Iliakis, G.; Wang, H.; Perrault, A.R.; Boecker, W.; Rosidi, B.; Windhofer, F.; Wu, W.; Guan, J.; Terzoudi, G.; Pantelias, G. Mechanisms of DNA double strand break repair and chromosome aberration formation. Cytogenet. Genome Res. 2004, 104, 14–20. [Google Scholar] [CrossRef] [PubMed]
  3. Yuan, J.; Chen, J. MRE11-RAD50-NBS1 complex dictates DNA repair independent of H2AX. J. Biol. Chem. 2010, 285, 1097–1104. [Google Scholar] [CrossRef] [PubMed]
  4. Mladenov, E.; Magin, S.; Soni, A.; Iliakis, G. DNA double-strand break repair as determinant of cellular radiosensitivity to killing and target in radiation therapy. Front. Oncol. 2013, 3, 113. [Google Scholar] [CrossRef] [PubMed]
  5. Nickoloff, J.A.; Jones, D.; Lee, S.H.; Williamson, E.A.; Hromas, R. Drugging the cancers addicted to DNA repair. J. Natl. Cancer Inst. 2017, 109. [Google Scholar] [CrossRef]
  6. Williams, R.S.; Williams, J.S.; Tainer, J.A. Mre11-Rad50-Nbs1 is a keystone complex connecting DNA repair machinery, double-strand break signaling, and the chromatin template. Biochem. Cell Biol. 2007, 85, 509–520. [Google Scholar] [CrossRef] [PubMed]
  7. Ronco, C.; Martin, A.R.; Demange, L.; Benhida, R. ATM, ATR, CHK1, CHK2 and WEE1 inhibitors in cancer and cancer stem cells. Medchemcomm 2017, 8, 295–319. [Google Scholar] [CrossRef]
  8. Weber, A.M.; Ryan, A.J. ATM and ATR as therapeutic targets in cancer. Pharmacol. Ther. 2015, 149, 124–138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Iliakis, G. Backup pathways of NHEJ in cells of higher eukaryotes: Cell cycle dependence. Radiother. Oncol. 2009, 92, 310–315. [Google Scholar] [CrossRef]
  10. Jeggo, P.A.; Geuting, V.; Lobrich, M. The role of homologous recombination in radiation-induced double-strand break repair. Radiother. Oncol. 2011, 101, 7–12. [Google Scholar] [CrossRef]
  11. Martin, L.M.; Marples, B.; Coffey, M.; Lawler, M.; Lynch, T.H.; Hollywood, D.; Marignol, L. DNA mismatch repair and the DNA damage response to ionizing radiation: Making sense of apparently conflicting data. Cancer Treat. Rev. 2010, 36, 518–527. [Google Scholar] [CrossRef]
  12. Kennedy, R.D.; D’Andrea, A.D. DNA repair pathways in clinical practice: Lessons from pediatric cancer susceptibility syndromes. J. Clin. Oncol. 2006, 24, 3799–3808. [Google Scholar] [CrossRef] [PubMed]
  13. Stachelek, G.C.; Peterson-Roth, E.; Liu, Y.; Fernandez, R.J., 3rd; Pike, L.R.; Qian, J.M.; Abriola, L.; Hoyer, D.; Hungerford, W.; Merkel, J.; et al. YU238259 is a novel inhibitor of homology-dependent DNA repair that exhibits synthetic lethality and radiosensitization in repair-deficient tumors. Mol. Cancer Res. 2015. [Google Scholar] [CrossRef] [PubMed]
  14. Thoms, J.; Bristow, R.G. DNA repair targeting and radiotherapy: A focus on the therapeutic ratio. Semin. Radiat. Oncol. 2010, 20, 217–222. [Google Scholar] [CrossRef]
  15. Rodemann, H.P.; Dittmann, K.; Toulany, M. Radiation-induced EGFR-signaling and control of DNA-damage repair. Int. J. Radiat. Biol. 2007, 83, 781–791. [Google Scholar] [CrossRef] [PubMed]
  16. Pueyo, G.; Mesia, R.; Figueras, A.; Lozano, A.; Baro, M.; Vazquez, S.; Capella, G.; Balart, J. Cetuximab may inhibit tumor growth and angiogenesis induced by ionizing radiation: A preclinical rationale for maintenance treatment after radiotherapy. Oncologist 2010, 15, 976–986. [Google Scholar] [CrossRef] [PubMed]
  17. Toulany, M.; Dittmann, K.; Baumann, M.; Rodemann, H.P. Radiosensitization of Ras-mutated human tumor cells in vitro by the specific EGF receptor antagonist BIBX1382BS. Radiother. Oncol. 2005, 74, 117–129. [Google Scholar] [CrossRef] [PubMed]
  18. Amorino, G.P.; Hamilton, V.M.; Valerie, K.; Dent, P.; Lammering, G.; Schmidt-Ullrich, R.K. Epidermal growth factor receptor dependence of radiation-induced transcription factor activation in human breast carcinoma cells. Mol. Biol. Cell 2002, 13, 2233–2244. [Google Scholar] [CrossRef] [PubMed]
  19. Schmidt-Ullrich, R.K.; Mikkelsen, R.B.; Dent, P.; Todd, D.G.; Valerie, K.; Kavanagh, B.D.; Contessa, J.N.; Rorrer, W.K.; Chen, P.B. Radiation-induced proliferation of the human A431 squamous carcinoma cells is dependent on EGFR tyrosine phosphorylation. Oncogene 1997, 15, 1191–1197. [Google Scholar] [CrossRef] [Green Version]
  20. Contessa, J.N.; Hampton, J.; Lammering, G.; Mikkelsen, R.B.; Dent, P.; Valerie, K.; Schmidt-Ullrich, R.K. Ionizing radiation activates Erb-B receptor dependent AKT and p70 S6 kinase signaling in carcinoma cells. Oncogene 2002, 21, 4032–4041. [Google Scholar] [CrossRef] [PubMed]
  21. Blume-Jensen, P.; Hunter, T. Oncogenic kinase signalling. Nature 2001, 411, 355–365. [Google Scholar] [CrossRef]
  22. Spangle, J.M.; Roberts, T.M. Epigenetic regulation of RTK signaling. J. Mol. Med. 2017, 95, 791–798. [Google Scholar] [CrossRef]
  23. Song, H.; Li, C.W.; Labaff, A.M.; Lim, S.O.; Li, L.Y.; Kan, S.F.; Chen, Y.; Zhang, K.; Lang, J.; Xie, X.; et al. Acetylation of EGF receptor contributes to tumor cell resistance to histone deacetylase inhibitors. Biochem. Biophys. Res. Commun. 2011, 404, 68–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wang, S.C.; Hung, M.C. Nuclear translocation of the epidermal growth factor receptor family membrane tyrosine kinase receptors. Clin Cancer Res. 2009, 15, 6484–6489. [Google Scholar] [CrossRef] [PubMed]
  25. Dittmann, K.; Mayer, C.; Fehrenbacher, B.; Schaller, M.; Raju, U.; Milas, L.; Chen, D.J.; Kehlbach, R.; Rodemann, H.P. Radiation-induced epidermal growth factor receptor nuclear import is linked to activation of DNA-dependent protein kinase. J. Biol. Chem. 2005, 280, 31182–31189. [Google Scholar] [CrossRef] [PubMed]
  26. Das, A.K.; Chen, B.P.; Story, M.D.; Sato, M.; Minna, J.D.; Chen, D.J.; Nirodi, C.S. Somatic mutations in the tyrosine kinase domain of epidermal growth factor receptor (EGFR) abrogate EGFR-mediated radioprotection in non-small cell lung carcinoma. Cancer Res. 2007, 67, 5267–5274. [Google Scholar] [CrossRef] [PubMed]
  27. Dittmann, K.; Mayer, C.; Czemmel, S.; Huber, S.M.; Rodemann, H.P. New roles for nuclear EGFR in regulating the stability and translation of mRNAs associated with VEGF signaling. PLoS ONE 2017, 12, e0189087. [Google Scholar] [CrossRef] [PubMed]
  28. Faria, J.; de Andrade, C.; Goes, A.M.; Rodrigues, M.A.; Gomes, D.A. Effects of different ligands on epidermal growth factor receptor (EGFR) nuclear translocation. Biochem. Biophys. Res. Commun. 2016, 478, 39–45. [Google Scholar] [CrossRef] [Green Version]
  29. Minjgee, M.; Toulany, M.; Kehlbach, R.; Giehl, K.; Rodemann, H.P. K-RAS(V12) induces autocrine production of EGFR ligands and mediates radioresistance through EGFR-dependent AKT signaling and activation of DNA-PKcs. Int. J. Radiat. Oncol. Biol. Phys. 2011, 81, 1506–1514. [Google Scholar] [CrossRef]
  30. Saki, M.; Toulany, M.; Rodemann, H.P. Acquired resistance to cetuximab is associated with the overexpression of Ras family members and the loss of radiosensitization in head and neck cancer cells. Radiother. Oncol. 2013, 108, 473–478. [Google Scholar] [CrossRef]
  31. Toulany, M.; Baumann, M.; Rodemann, H.P. Stimulated PI3K-AKT signaling mediated through ligand or radiation-induced EGFR depends indirectly, but not directly, on constitutive K-Ras activity. Mol. Cancer Res. 2007, 5, 863–872. [Google Scholar] [CrossRef]
  32. Nowsheen, S.; Cooper, T.; Stanley, J.A.; Yang, E.S. Synthetic lethal interactions between EGFR and PARP inhibition in human triple negative breast cancer cells. PLoS ONE 2012, 7, e46614. [Google Scholar] [CrossRef]
  33. Lee, H.J.; Lan, L.; Peng, G.; Chang, W.C.; Hsu, M.C.; Wang, Y.N.; Cheng, C.C.; Wei, L.; Nakajima, S.; Chang, S.S.; et al. Tyrosine 370 phosphorylation of ATM positively regulates DNA damage response. Cell Res. 2015, 25, 225–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wang, S.C.; Nakajima, Y.; Yu, Y.L.; Xia, W.; Chen, C.T.; Yang, C.C.; McIntush, E.W.; Li, L.Y.; Hawke, D.H.; Kobayashi, R.; et al. Tyrosine phosphorylation controls PCNA function through protein stability. Nat. Cell Biol. 2006, 8, 1359–1368. [Google Scholar] [CrossRef] [PubMed]
  35. Chou, R.H.; Wang, Y.N.; Hsieh, Y.H.; Li, L.Y.; Xia, W.; Chang, W.C.; Chang, L.C.; Cheng, C.C.; Lai, C.C.; Hsu, J.L.; et al. EGFR modulates DNA synthesis and repair through Tyr phosphorylation of histone H4. Dev. Cell 2014, 30, 224–237. [Google Scholar] [CrossRef] [PubMed]
  36. Myllynen, L.; Rieckmann, T.; Dahm-Daphi, J.; Kasten-Pisula, U.; Petersen, C.; Dikomey, E.; Kriegs, M. In tumor cells regulation of DNA double strand break repair through EGF receptor involves both NHEJ and HR and is independent of p53 and K-Ras status. Radiother. Oncol. 2011, 101, 147–151. [Google Scholar] [CrossRef]
  37. Friedmann, B.J.; Caplin, M.; Savic, B.; Shah, T.; Lord, C.J.; Ashworth, A.; Hartley, J.A.; Hochhauser, D. Interaction of the epidermal growth factor receptor and the DNA-dependent protein kinase pathway following gefitinib treatment. Mol. Cancer Ther. 2006, 5, 209–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Li, L.; Wang, H.; Yang, E.S.; Arteaga, C.L.; Xia, F. Erlotinib attenuates homologous recombinational repair of chromosomal breaks in human breast cancer cells. Cancer Res. 2008, 68, 9141–9146. [Google Scholar] [CrossRef]
  39. Li, Y.H.; Wang, X.; Pan, Y.; Lee, D.H.; Chowdhury, D.; Kimmelman, A.C. Inhibition of non-homologous end joining repair impairs pancreatic cancer growth and enhances radiation response. PLoS ONE 2012, 7, e39588. [Google Scholar] [CrossRef] [PubMed]
  40. Milas, L.; Fan, Z.; Andratschke, N.H.; Ang, K.K. Epidermal growth factor receptor and tumor response to radiation: In vivo preclinical studies. Int. J. Radiat. Oncol. Biol. Phys. 2004, 58, 966–971. [Google Scholar] [CrossRef] [PubMed]
  41. Ang, K.K.; Berkey, B.A.; Tu, X.; Zhang, H.Z.; Katz, R.; Hammond, E.H.; Fu, K.K.; Milas, L. Impact of epidermal growth factor receptor expression on survival and pattern of relapse in patients with advanced head and neck carcinoma. Cancer Res. 2002, 62, 7350–7356. [Google Scholar] [PubMed]
  42. Iida, M.; Bahrar, H.; Brand, T.M.; Pearson, H.E.; Coan, J.P.; Orbuch, R.A.; Flanigan, B.G.; Swick, A.D.; Prabakaran, P.J.; Lantto, J.; et al. Targeting the HER family with pan-HER effectively overcomes resistance to cetuximab. Mol. Cancer Ther. 2016, 15, 2175–2186. [Google Scholar] [CrossRef]
  43. Iida, M.; Brand, T.M.; Starr, M.M.; Li, C.; Huppert, E.J.; Luthar, N.; Pedersen, M.W.; Horak, I.D.; Kragh, M.; Wheeler, D.L. Sym004, a novel EGFR antibody mixture, can overcome acquired resistance to cetuximab. Neoplasia 2013, 15, 1196–1206. [Google Scholar] [CrossRef]
  44. Francis, D.M.; Huang, S.; Armstrong, E.A.; Werner, L.R.; Hullett, C.; Li, C.; Morris, Z.S.; Swick, A.D.; Kragh, M.; Lantto, J.; et al. Pan-HER inhibitor augments radiation response in human lung and head and neck cancer models. Clin. Cancer Res. 2016, 22, 633–643. [Google Scholar] [CrossRef] [PubMed]
  45. Huang, S.; Peter Rodemann, H.; Harari, P.M. Molecular targeting of growth factor receptor signaling in radiation oncology. Recent Results Cancer Res. 2016, 198, 45–87. [Google Scholar] [CrossRef] [PubMed]
  46. Bonner, J.A.; Harari, P.M.; Giralt, J.; Azarnia, N.; Shin, D.M.; Cohen, R.B.; Jones, C.U.; Sur, R.; Raben, D.; Jassem, J.; et al. Radiotherapy plus cetuximab for squamous-cell carcinoma of the head and neck. N. Engl. J. Med. 2006, 354, 567–578. [Google Scholar] [CrossRef]
  47. Bonner, J.A.; Harari, P.M.; Giralt, J.; Cohen, R.B.; Jones, C.U.; Sur, R.K.; Raben, D.; Baselga, J.; Spencer, S.A.; Zhu, J.; et al. Radiotherapy plus cetuximab for locoregionally advanced head and neck cancer: 5-year survival data from a phase 3 randomised trial, and relation between cetuximab-induced rash and survival. Lancet Oncol. 2010, 11, 21–28. [Google Scholar] [CrossRef]
  48. Garg, M.K.; Zhao, F.; Sparano, J.A.; Palefsky, J.; Whittington, R.; Mitchell, E.P.; Mulcahy, M.F.; Armstrong, K.I.; Nabbout, N.H.; Kalnicki, S.; et al. Cetuximab plus chemoradiotherapy in immunocompetent patients with anal carcinoma: A phase II Eastern cooperative oncology group-American College of Radiology Imaging Network Cancer Research Group Trial (E3205). J. Clin. Oncol. 2017, 35, 718–726. [Google Scholar] [CrossRef]
  49. Sparano, J.A.; Lee, J.Y.; Palefsky, J.; Henry, D.H.; Wachsman, W.; Rajdev, L.; Aboulafia, D.; Ratner, L.; Fitzgerald, T.J.; Kachnic, L.; et al. Cetuximab plus chemoradiotherapy for HIV-associated anal carcinoma: A phase II AIDS malignancy consortium trial. J. Clin. Oncol. 2017, 35, 727–733. [Google Scholar] [CrossRef]
  50. Eisterer, W.; De Vries, A.; Ofner, D.; Rabl, H.; Koplmuller, R.; Greil, R.; Tschmelitsch, J.; Schmid, R.; Kapp, K.; Lukas, P.; et al. Preoperative treatment with capecitabine, cetuximab and radiotherapy for primary locally advanced rectal cancer--a phase II clinical trial. Anticancer Res. 2014, 34, 6767–6773. [Google Scholar]
  51. Bonomo, P.; Loi, M.; Desideri, I.; Olmetto, E.; Delli Paoli, C.; Terziani, F.; Greto, D.; Mangoni, M.; Scoccianti, S.; Simontacchi, G.; et al. Incidence of skin toxicity in squamous cell carcinoma of the head and neck treated with radiotherapy and cetuximab: A systematic review. Crit. Rev. Oncol. Hematol. 2017, 120, 98–110. [Google Scholar] [CrossRef] [PubMed]
  52. Deutsch, E.; Lemanski, C.; Pignon, J.P.; Levy, A.; Delarochefordiere, A.; Martel-Lafay, I.; Rio, E.; Malka, D.; Conroy, T.; Miglianico, L.; et al. Unexpected toxicity of cetuximab combined with conventional chemoradiotherapy in patients with locally advanced anal cancer: Results of the UNICANCER ACCORD 16 phase II trial. Ann. Oncol. 2013, 24, 2834–2838. [Google Scholar] [CrossRef] [PubMed]
  53. Bradley, J.D.; Paulus, R.; Komaki, R.; Masters, G.; Blumenschein, G.; Schild, S.; Bogart, J.; Hu, C.; Forster, K.; Magliocco, A.; et al. Standard-dose versus high-dose conformal radiotherapy with concurrent and consolidation carboplatin plus paclitaxel with or without cetuximab for patients with stage IIIA or IIIB non-small-cell lung cancer (RTOG 0617): A randomised, two-by-two factorial phase 3 study. Lancet Oncol. 2015, 16, 187–199. [Google Scholar] [CrossRef] [PubMed]
  54. Crosby, T.; Hurt, C.N.; Falk, S.; Gollins, S.; Mukherjee, S.; Staffurth, J.; Ray, R.; Bashir, N.; Bridgewater, J.A.; Geh, J.I.; et al. Chemoradiotherapy with or without cetuximab in patients with oesophageal cancer (SCOPE1): A multicentre, phase 2/3 randomised trial. Lancet Oncol. 2013, 14, 627–637. [Google Scholar] [CrossRef]
  55. Ang, K.K.; Zhang, Q.; Rosenthal, D.I.; Nguyen-Tan, P.F.; Sherman, E.J.; Weber, R.S.; Galvin, J.M.; Bonner, J.A.; Harris, J.; El-Naggar, A.K.; et al. Randomized phase III trial of concurrent accelerated radiation plus cisplatin with or without cetuximab for stage III to IV head and neck carcinoma: RTOG 0522. J. Clin. Oncol. 2014, 32, 2940–2950. [Google Scholar] [CrossRef] [PubMed]
  56. Gillison, M.L.; Trotti, A.M.; Harris, J.; Eisbruch, A.; Harari, P.M.; Adelstein, D.J.; Sturgis, E.M.; Burtness, B.; Ridge, J.A.; Ringash, J.; et al. Radiotherapy plus cetuximab or cisplatin in human papillomavirus-positive oropharyngeal cancer (NRG Oncology RTOG 1016): A randomised, multicentre, non-inferiority trial. Lancet 2018. [Google Scholar] [CrossRef]
  57. Chang, C.C.; Chi, K.H.; Kao, S.J.; Hsu, P.S.; Tsang, Y.W.; Chang, H.J.; Yeh, Y.W.; Hsieh, Y.S.; Jiang, J.S. Upfront gefitinib/erlotinib treatment followed by concomitant radiotherapy for advanced lung cancer: A mono-institutional experience. Lung Cancer 2011, 73, 189–194. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, J.; Xia, T.Y.; Wang, Y.J.; Li, H.Q.; Li, P.; Wang, J.D.; Chang, D.S.; Liu, L.Y.; Di, Y.P.; Wang, X.; et al. Prospective study of epidermal growth factor receptor tyrosine kinase inhibitors concurrent with individualized radiotherapy for patients with locally advanced or metastatic non-small-cell lung cancer. Int. J. Radiat. Oncol. Biol. Phys. 2011, 81, e59–e65. [Google Scholar] [CrossRef]
  59. Iyengar, P.; Kavanagh, B.D.; Wardak, Z.; Smith, I.; Ahn, C.; Gerber, D.E.; Dowell, J.; Hughes, R.; Abdulrahman, R.; Camidge, D.R.; et al. Phase II trial of stereotactic body radiation therapy combined with erlotinib for patients with limited but progressive metastatic non-small-cell lung cancer. J. Clin. Oncol. 2014, 32, 3824–3830. [Google Scholar] [CrossRef]
  60. Casal Rubio, J.; Firvida-Perez, J.L.; Lazaro-Quintela, M.; Baron-Duarte, F.J.; Alonso-Jaudenes, G.; Santome, L.; Afonso-Afonso, F.J.; Amenedo, M.; Huidobro, G.; Campos-Balea, B.; et al. A phase II trial of erlotinib as maintenance treatment after concurrent chemoradiotherapy in stage III non-small-cell lung cancer (NSCLC): A Galician Lung Cancer Group (GGCP) study. Cancer Chemother. Pharmacol. 2014, 73, 451–457. [Google Scholar] [CrossRef]
  61. Hainsworth, J.D.; Spigel, D.R.; Greco, F.A.; Shipley, D.L.; Peyton, J.; Rubin, M.; Stipanov, M.; Meluch, A. Combined modality treatment with chemotherapy, radiation therapy, bevacizumab, and erlotinib in patients with locally advanced squamous carcinoma of the head and neck: A phase II trial of the Sarah Cannon oncology research consortium. Cancer J. 2011, 17, 267–272. [Google Scholar] [CrossRef] [PubMed]
  62. Clarke, J.L.; Molinaro, A.M.; Phillips, J.J.; Butowski, N.A.; Chang, S.M.; Perry, A.; Costello, J.F.; DeSilva, A.A.; Rabbitt, J.E.; Prados, M.D. A single-institution phase II trial of radiation, temozolomide, erlotinib, and bevacizumab for initial treatment of glioblastoma. Neuro-Oncology 2014, 16, 984–990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Harari, P.M.; Huang, S.M. Head and neck cancer as a clinical model for molecular targeting of therapy: Combining EGFR blockade with radiation. Int. J. Radiat. Oncol. Biol. Phys. 2001, 49, 427–433. [Google Scholar] [CrossRef]
  64. Xie, Y.; Hung, M.C. Nuclear localization of p185neu tyrosine kinase and its association with transcriptional transactivation. Biochem. Biophys. Res. Commun. 1994, 203, 1589–1598. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, S.C.; Lien, H.C.; Xia, W.; Chen, I.F.; Lo, H.W.; Wang, Z.; Ali-Seyed, M.; Lee, D.F.; Bartholomeusz, G.; Ou-Yang, F.; et al. Binding at and transactivation of the COX-2 promoter by nuclear tyrosine kinase receptor ErbB-2. Cancer Cell 2004, 6, 251–261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Reif, R.; Adawy, A.; Vartak, N.; Schroder, J.; Gunther, G.; Ghallab, A.; Schmidt, M.; Schormann, W.; Hengstler, J.G. Activated ErbB3 translocates to the nucleus via clathrin-independent endocytosis, which is associated with proliferating cells. J. Biol. Chem. 2016, 291, 3837–3847. [Google Scholar] [CrossRef]
  67. Huang, S.; Li, C.; Armstrong, E.A.; Peet, C.R.; Saker, J.; Amler, L.C.; Sliwkowski, M.X.; Harari, P.M. Dual targeting of EGFR and HER3 with MEHD7945A overcomes acquired resistance to EGFR inhibitors and radiation. Cancer Res. 2013, 73, 824–833. [Google Scholar] [CrossRef]
  68. Hagerstrand, D.; Lindh, M.B.; Pena, C.; Garcia-Echeverria, C.; Nister, M.; Hofmann, F.; Ostman, A. PI3K/PTEN/AKT pathway status affects the sensitivity of high-grade glioma cell cultures to the insulin-like growth factor-1 receptor inhibitor NVP-AEW541. Neuro-Oncology 2010, 12, 967–975. [Google Scholar] [CrossRef]
  69. Waraky, A.; Lin, Y.; Warsito, D.; Haglund, F.; Aleem, E.; Larsson, O. Nuclear insulin-like growth factor 1 receptor phosphorylates proliferating cell nuclear antigen and rescues stalled replication forks after DNA damage. J. Biol. Chem. 2017, 292, 18227–18239. [Google Scholar] [CrossRef]
  70. Codony-Servat, J.; Cuatrecasas, M.; Asensio, E.; Montironi, C.; Martinez-Cardus, A.; Marin-Aguilera, M.; Horndler, C.; Martinez-Balibrea, E.; Rubini, M.; Jares, P.; et al. Nuclear IGF-1R predicts chemotherapy and targeted therapy resistance in metastatic colorectal cancer. Br. J. Cancer 2017, 117, 1777–1786. [Google Scholar] [CrossRef]
  71. Aleksic, T.; Verrill, C.; Bryant, R.J.; Han, C.; Worrall, A.R.; Brureau, L.; Larre, S.; Higgins, G.S.; Fazal, F.; Sabbagh, A.; et al. IGF-1R associates with adverse outcomes after radical radiotherapy for prostate cancer. Br. J. Cancer 2017, 117, 1600–1606. [Google Scholar] [CrossRef] [PubMed]
  72. Lodhia, K.A.; Gao, S.; Aleksic, T.; Esashi, F.; Macaulay, V.M. Suppression of homologous recombination sensitizes human tumor cells to IGF-1R inhibition. Int. J. Cancer 2015, 136, 2961–2966. [Google Scholar] [CrossRef] [PubMed]
  73. Chitnis, M.M.; Lodhia, K.A.; Aleksic, T.; Gao, S.; Protheroe, A.S.; Macaulay, V.M. IGF-1R inhibition enhances radiosensitivity and delays double-strand break repair by both non-homologous end-joining and homologous recombination. Oncogene 2014, 33, 5262–5273. [Google Scholar] [CrossRef] [PubMed]
  74. Turney, B.W.; Kerr, M.; Chitnis, M.M.; Lodhia, K.; Wang, Y.; Riedemann, J.; Rochester, M.; Protheroe, A.S.; Brewster, S.F.; Macaulay, V.M. Depletion of the type 1 IGF receptor delays repair of radiation-induced DNA double strand breaks. Radiother. Oncol. 2012, 103, 402–409. [Google Scholar] [CrossRef] [PubMed]
  75. Allen, G.W.; Saba, C.; Armstrong, E.A.; Huang, S.M.; Benavente, S.; Ludwig, D.L.; Hicklin, D.J.; Harari, P.M. Insulin-like growth factor-I receptor signaling blockade combined with radiation. Cancer Res. 2007, 67, 1155–1162. [Google Scholar] [CrossRef] [PubMed]
  76. Iwasa, T.; Okamoto, I.; Suzuki, M.; Hatashita, E.; Yamada, Y.; Fukuoka, M.; Ono, K.; Nakagawa, K. Inhibition of insulin-like growth factor 1 receptor by CP-751,871 radiosensitizes non-small cell lung cancer cells. Clin. Cancer Res. 2009, 15, 5117–5125. [Google Scholar] [CrossRef] [PubMed]
  77. Isebaert, S.F.; Swinnen, J.V.; McBride, W.H.; Haustermans, K.M. Insulin-like growth factor-type 1 receptor inhibitor NVP-AEW541 enhances radiosensitivity of PTEN wild-type but not PTEN-deficient human prostate cancer cells. Int. J. Radiat. Oncol. Biol. Phys. 2011, 81, 239–247. [Google Scholar] [CrossRef]
  78. Zhao, H.; Gu, X. Silencing of insulin-like growth factor-1 receptor enhances the radiation sensitivity of human esophageal squamous cell carcinoma in vitro and in vivo. World J. Surg. Oncol. 2014, 12, 325. [Google Scholar] [CrossRef] [Green Version]
  79. Liu, C.; Zhang, Z.; Tang, H.; Jiang, Z.; You, L.; Liao, Y. Crosstalk between IGF-1R and other tumor promoting pathways. Curr. Pharm. Des. 2014, 20, 2912–2921. [Google Scholar] [CrossRef]
  80. Matsubara, J.; Yamada, Y.; Nakajima, T.E.; Kato, K.; Hamaguchi, T.; Shirao, K.; Shimada, Y.; Shimoda, T. Clinical significance of insulin-like growth factor type 1 receptor and epidermal growth factor receptor in patients with advanced gastric cancer. Oncology 2008, 74, 76–83. [Google Scholar] [CrossRef]
  81. McDaniel, N.K.; Cummings, C.T.; Iida, M.; Hulse, J.; Pearson, H.E.; Vasileiadi, E.; Parker, R.E.; Orbuch, R.A.; Ondracek, O.J.; Welke, N.B.; et al. MERTK mediates intrinsic and adaptive resistance to AXL-targeting agents. Mol. Cancer Ther. 2018. [Google Scholar] [CrossRef]
  82. Kasikara, C.; Kumar, S.; Kimani, S.; Tsou, W.I.; Geng, K.; Davra, V.; Sriram, G.; Devoe, C.; Nguyen, K.N.; Antes, A.; et al. Phosphatidylserine sensing by TAM receptors regulates AKT-dependent chemoresistance and PD-L1 expression. Mol. Cancer Res. 2017, 15, 753–764. [Google Scholar] [CrossRef] [PubMed]
  83. Lin, J.Z.; Wang, Z.J.; De, W.; Zheng, M.; Xu, W.Z.; Wu, H.F.; Armstrong, A.; Zhu, J.G. Targeting AXL overcomes resistance to docetaxel therapy in advanced prostate cancer. Oncotarget 2017, 8, 41064–41077. [Google Scholar] [CrossRef] [PubMed]
  84. Ludwig, K.F.; Du, W.; Sorrelle, N.B.; Wnuk-Lipinska, K.; Topalovski, M.; Toombs, J.E.; Cruz, V.H.; Yabuuchi, S.; Rajeshkumar, N.V.; Maitra, A.; et al. Small-molecule inhibition of Axl targets tumor immune suppression and enhances chemotherapy in pancreatic cancer. Cancer Res. 2018, 78, 246–255. [Google Scholar] [CrossRef] [PubMed]
  85. Balaji, K.; Vijayaraghavan, S.; Diao, L.; Tong, P.; Fan, Y.; Carey, J.P.; Bui, T.N.; Warner, S.; Heymach, J.V.; Hunt, K.K.; et al. AXL inhibition suppresses the DNA damage response and sensitizes cells to PARP inhibition in multiple cancers. Mol. Cancer Res. 2017, 15, 45–58. [Google Scholar] [CrossRef] [PubMed]
  86. Skinner, H.D.; Giri, U.; Yang, L.P.; Kumar, M.; Liu, Y.; Story, M.D.; Pickering, C.R.; Byers, L.A.; Williams, M.D.; Wang, J.; et al. Integrative analysis identifies a novel AXL-PI3 Kinase-PD-L1 signaling axis associated with radiation resistance in head and neck cancer. Clin. Cancer Res. 2017, 23, 2713–2722. [Google Scholar] [CrossRef] [PubMed]
  87. Sinha, S.; Boysen, J.; Nelson, M.; Secreto, C.; Warner, S.L.; Bearss, D.J.; Lesnick, C.; Shanafelt, T.D.; Kay, N.E.; Ghosh, A.K. Targeted Axl inhibition primes chronic lymphocytic leukemia B cells to apoptosis and shows synergistic/additive effects in combination with BTK inhibitors. Clin. Cancer Res. 2015, 21, 2115–2126. [Google Scholar] [CrossRef]
  88. Holland, S.J.; Pan, A.; Franci, C.; Hu, Y.; Chang, B.; Li, W.; Duan, M.; Torneros, A.; Yu, J.; Heckrodt, T.J.; et al. R428, a selective small molecule inhibitor of Axl kinase, blocks tumor spread and prolongs survival in models of metastatic breast cancer. Cancer Res. 2010, 70, 1544–1554. [Google Scholar] [CrossRef]
  89. Aveic, S.; Corallo, D.; Porcu, E.; Pantile, M.; Boso, D.; Zanon, C.; Viola, G.; Sidarovich, V.; Mariotto, E.; Quattrone, A.; et al. TP-0903 inhibits neuroblastoma cell growth and enhances the sensitivity to conventional chemotherapy. Eur. J. Pharmacol. 2018, 818, 435–448. [Google Scholar] [CrossRef]
  90. Zhen, Y.; Lee, I.J.; Finkelman, F.D.; Shao, W.H. Targeted inhibition of Axl receptor tyrosine kinase ameliorates anti-GBM-induced lupus-like nephritis. J. Autoimmun. 2018, 93, 37–44. [Google Scholar] [CrossRef]
  91. Brand, T.M.; Iida, M.; Stein, A.P.; Corrigan, K.L.; Braverman, C.M.; Coan, J.P.; Pearson, H.E.; Bahrar, H.; Fowler, T.L.; Bednarz, B.P.; et al. AXL is a logical molecular target in head and neck squamous cell carcinoma. Clin. Cancer Res. 2015, 21, 2601–2612. [Google Scholar] [CrossRef] [PubMed]
  92. Brand, T.M.; Iida, M.; Corrigan, K.L.; Braverman, C.M.; Coan, J.P.; Flanigan, B.G.; Stein, A.P.; Salgia, R.; Rolff, J.; Kimple, R.J.; et al. The receptor tyrosine kinase AXL mediates nuclear translocation of the epidermal growth factor receptor. Sci. Signal. 2017, 10. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, P.; Cheng, H.; Roberts, T.M.; Zhao, J.J. Targeting the phosphoinositide 3-kinase pathway in cancer. Nat. Rev. Drug Discov. 2009, 8, 627–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Liu, Z.; Roberts, T.M. Human tumor mutants in the p110α subunit of PI3K. Cell Cycle 2006, 5, 675–677. [Google Scholar] [CrossRef] [PubMed]
  95. Lui, V.W.; Hedberg, M.L.; Li, H.; Vangara, B.S.; Pendleton, K.; Zeng, Y.; Lu, Y.; Zhang, Q.; Du, Y.; Gilbert, B.R.; et al. Frequent mutation of the PI3K pathway in head and neck cancer defines predictive biomarkers. Cancer Discov. 2013, 3, 761–769. [Google Scholar] [CrossRef] [PubMed]
  96. Cancer Genome Atlas Research, N. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature 2008, 455, 1061–1068. [Google Scholar] [CrossRef]
  97. Zhang, T.; Cui, G.B.; Zhang, J.; Zhang, F.; Zhou, Y.A.; Jiang, T.; Li, X.F. Inhibition of PI3 kinases enhances the sensitivity of non-small cell lung cancer cells to ionizing radiation. Oncol. Rep. 2010, 24, 1683–1689. [Google Scholar] [PubMed]
  98. Valerie, K.; Yacoub, A.; Hagan, M.P.; Curiel, D.T.; Fisher, P.B.; Grant, S.; Dent, P. Radiation-induced cell signaling: Inside-out and outside-in. Mol. Cancer Ther. 2007, 6, 789–801. [Google Scholar] [CrossRef]
  99. Toulany, M.; Minjgee, M.; Kehlbach, R.; Chen, J.; Baumann, M.; Rodemann, H.P. ErbB2 expression through heterodimerization with erbB1 is necessary for ionizing radiation- but not EGF-induced activation of AKT survival pathway. Radiother. Oncol. J. Eur. Soc. Ther. Radiol. Oncol. 2010, 97, 338–345. [Google Scholar] [CrossRef]
  100. Li, H.F.; Kim, J.S.; Waldman, T. Radiation-induced AKT activation modulates radioresistance in human glioblastoma cells. Radiat. Oncol. 2009, 4, 43. [Google Scholar] [CrossRef]
  101. Winograd-Katz, S.E.; Levitzki, A. Cisplatin induces PKB/AKT activation and p38(MAPK) phosphorylation of the EGF receptor. Oncogene 2006, 25, 7381–7390. [Google Scholar] [CrossRef] [PubMed]
  102. Yu, C.C.; Hung, S.K.; Lin, H.Y.; Chiou, W.Y.; Lee, M.S.; Liao, H.F.; Huang, H.B.; Ho, H.C.; Su, Y.C. Targeting the PI3K/AKT/mTOR signaling pathway as an effectively radiosensitizing strategy for treating human oral squamous cell carcinoma in vitro and in vivo. Oncotarget 2017, 8, 68641–68653. [Google Scholar] [CrossRef] [PubMed]
  103. Bussink, J.; van der Kogel, A.J.; Kaanders, J.H. Activation of the PI3-K/AKT pathway and implications for radioresistance mechanisms in head and neck cancer. Lancet Oncol. 2008, 9, 288–296. [Google Scholar] [CrossRef]
  104. Toulany, M.; Dittmann, K.; Kruger, M.; Baumann, M.; Rodemann, H.P. Radioresistance of K-Ras mutated human tumor cells is mediated through EGFR-dependent activation of PI3K-AKT pathway. Radiother. Oncol. 2005, 76, 143–150. [Google Scholar] [CrossRef] [PubMed]
  105. Zhan, M.; Han, Z.C. Phosphatidylinositide 3-kinase/AKT in radiation responses. Histol. Histopathol. 2004, 19, 915–923. [Google Scholar] [PubMed]
  106. Liu, Y.; Cui, B.; Qiao, Y.; Zhang, Y.; Tian, Y.; Jiang, J.; Ma, D.; Kong, B. Phosphoinositide-3-kinase inhibition enhances radiosensitization of cervical cancer in vivo. Int. J. Gynecol. Cancer 2011, 21, 100–105. [Google Scholar] [CrossRef] [PubMed]
  107. Kao, G.D.; Jiang, Z.; Fernandes, A.M.; Gupta, A.K.; Maity, A. Inhibition of phosphatidylinositol-3-OH kinase/AKT signaling impairs DNA repair in glioblastoma cells following ionizing radiation. J. Biol. Chem. 2007, 282, 21206–21212. [Google Scholar] [CrossRef]
  108. Choi, E.J.; Ryu, Y.K.; Kim, S.Y.; Wu, H.G.; Kim, J.S.; Kim, I.H.; Kim, I.A. Targeting epidermal growth factor receptor-associated signaling pathways in non-small cell lung cancer cells: Implication in radiation response. Mol. Cancer Res. 2010, 8, 1027–1036. [Google Scholar] [CrossRef]
  109. Toulany, M.; Kasten-Pisula, U.; Brammer, I.; Wang, S.; Chen, J.; Dittmann, K.; Baumann, M.; Dikomey, E.; Rodemann, H.P. Blockage of epidermal growth factor receptor-phosphatidylinositol 3-kinase-AKT signaling increases radiosensitivity of K-RAS mutated human tumor cells in vitro by affecting DNA repair. Clin. Cancer Res. 2006, 12, 4119–4126. [Google Scholar] [CrossRef]
  110. Greenwell, I.B.; Ip, A.; Cohen, J.B. PI3K inhibitors: Understanding toxicity mechanisms and management. Oncology 2017, 31, 821–828. [Google Scholar]
  111. Hemmings, B.A.; Restuccia, D.F. PI3K-PKB/AKT pathway. Cold Spring Harb. Perspect. Biol. 2012, 4, a011189. [Google Scholar] [CrossRef] [PubMed]
  112. Fayard, E.; Xue, G.; Parcellier, A.; Bozulic, L.; Hemmings, B.A. Protein kinase B (PKB/AKT), a key mediator of the PI3K signaling pathway. Curr. Top. Microbiol. Immunol. 2010, 346, 31–56. [Google Scholar] [CrossRef]
  113. Toulany, M.; Kehlbach, R.; Florczak, U.; Sak, A.; Wang, S.; Chen, J.; Lobrich, M.; Rodemann, H.P. Targeting of AKT1 enhances radiation toxicity of human tumor cells by inhibiting DNA-PKcs-dependent DNA double-strand break repair. Mol. Cancer Ther. 2008, 7, 1772–1781. [Google Scholar] [CrossRef] [Green Version]
  114. Golding, S.E.; Morgan, R.N.; Adams, B.R.; Hawkins, A.J.; Povirk, L.F.; Valerie, K. Pro-survival AKT and ERK signaling from EGFR and mutant EGFRvIII enhances DNA double-strand break repair in human glioma cells. Cancer Biol. Ther. 2009, 8, 730–738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Fraser, M.; Harding, S.M.; Zhao, H.; Coackley, C.; Durocher, D.; Bristow, R.G. MRE11 promotes AKT phosphorylation in direct response to DNA double-strand breaks. Cell Cycle 2011, 10, 2218–2232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Toulany, M.; Lee, K.J.; Fattah, K.R.; Lin, Y.F.; Fehrenbacher, B.; Schaller, M.; Chen, B.P.; Chen, D.J.; Rodemann, H.P. AKT promotes post-irradiation survival of human tumor cells through initiation, progression, and termination of DNA-PKcs-dependent DNA double-strand break repair. Mol. Cancer Res. 2012, 10, 945–957. [Google Scholar] [CrossRef] [PubMed]
  117. Mueck, K.; Rebholz, S.; Harati, M.D.; Rodemann, H.P.; Toulany, M. AKT1 stimulates homologous recombination repair of DNA double-strand breaks in a Rad51-dependent manner. Int. J. Mol. Sci. 2017, 18, 2473. [Google Scholar] [CrossRef]
  118. Gupta, A.K.; McKenna, W.G.; Weber, C.N.; Feldman, M.D.; Goldsmith, J.D.; Mick, R.; Machtay, M.; Rosenthal, D.I.; Bakanauskas, V.J.; Cerniglia, G.J.; et al. Local recurrence in head and neck cancer: Relationship to radiation resistance and signal transduction. Clin. Cancer Res. 2002, 8, 885–892. [Google Scholar]
  119. Kim, T.J.; Lee, J.W.; Song, S.Y.; Choi, J.J.; Choi, C.H.; Kim, B.G.; Lee, J.H.; Bae, D.S. Increased expression of pAKT is associated with radiation resistance in cervical cancer. Br. J. Cancer 2006, 94, 1678–1682. [Google Scholar] [CrossRef]
  120. Toulany, M.; Minjgee, M.; Saki, M.; Holler, M.; Meier, F.; Eicheler, W.; Rodemann, H.P. ERK2-dependent reactivation of AKT mediates the limited response of tumor cells with constitutive K-RAS activity to PI3K inhibition. Cancer Biol. Ther. 2014, 15, 317–328. [Google Scholar] [CrossRef]
  121. Toulany, M.; Iida, M.; Keinath, S.; Iyi, F.F.; Mueck, K.; Fehrenbacher, B.; Mansour, W.Y.; Schaller, M.; Wheeler, D.L.; Rodemann, H.P. Dual targeting of PI3K and MEK enhances the radiation response of K-RAS mutated non-small cell lung cancer. Oncotarget 2016, 7, 43746–43761. [Google Scholar] [CrossRef] [Green Version]
  122. Zhang, J.; Park, D.; Shin, D.M.; Deng, X. Targeting KRAS-mutant non-small cell lung cancer: Challenges and opportunities. Acta Biochim. Biophys. Sin. 2016, 48, 11–16. [Google Scholar] [CrossRef] [PubMed]
  123. Kuger, S.; Graus, D.; Brendtke, R.; Gunther, N.; Katzer, A.; Lutyj, P.; Polat, B.; Chatterjee, M.; Sukhorukov, V.L.; Flentje, M.; et al. Radiosensitization of glioblastoma cell lines by the dual PI3K and mTOR inhibitor NVP-BEZ235 depends on drug-irradiation schedule. Transl. Oncol. 2013, 6, 169–179. [Google Scholar] [CrossRef]
  124. Sathe, A.; Chalaud, G.; Oppolzer, I.; Wong, K.Y.; von Busch, M.; Schmid, S.C.; Tong, Z.; Retz, M.; Gschwend, J.E.; Schulz, W.A.; et al. Parallel PI3K, AKT and mTOR inhibition is required to control feedback loops that limit tumor therapy. PLoS ONE 2018, 13, e0190854. [Google Scholar] [CrossRef] [PubMed]
  125. Clement, E.; Inuzuka, H.; Nihira, N.T.; Wei, W.; Toker, A. Skp2-dependent reactivation of AKT drives resistance to PI3K inhibitors. Sci. Signal. 2018, 11. [Google Scholar] [CrossRef] [PubMed]
  126. Barnett, S.F.; Defeo-Jones, D.; Fu, S.; Hancock, P.J.; Haskell, K.M.; Jones, R.E.; Kahana, J.A.; Kral, A.M.; Leander, K.; Lee, L.L.; et al. Identification and characterization of pleckstrin-homology-domain-dependent and isoenzyme-specific AKT inhibitors. Biochem. J. 2005, 385, 399–408. [Google Scholar] [CrossRef] [PubMed]
  127. Masure, S.; Haefner, B.; Wesselink, J.J.; Hoefnagel, E.; Mortier, E.; Verhasselt, P.; Tuytelaars, A.; Gordon, R.; Richardson, A. Molecular cloning, expression and characterization of the human serine/threonine kinase AKT-3. Eur. J. Biochem. 1999, 265, 353–360. [Google Scholar] [CrossRef] [PubMed]
  128. Manning, B.D.; Cantley, L.C. AKT/PKB signaling: Navigating downstream. Cell 2007, 129, 1261–1274. [Google Scholar] [CrossRef] [PubMed]
  129. Park, J.; Feng, J.; Li, Y.; Hammarsten, O.; Brazil, D.P.; Hemmings, B.A. DNA-dependent protein kinase-mediated phosphorylation of protein kinase B requires a specific recognition sequence in the C-terminal hydrophobic motif. J. Biol. Chem. 2009, 284, 6169–6174. [Google Scholar] [CrossRef]
  130. Toulany, M.; Maier, J.; Iida, M.; Rebholz, S.; Holler, M.; Grottke, A.; Juker, M.; Wheeler, D.L.; Rothbauer, U.; Rodemann, H.P. AKT1 and AKT3 but not AKT2 through interaction with DNA-PKcs stimulate proliferation and post-irradiation cell survival of K-RAS-mutated cancer cells. Cell. Death Discov. 2017, 3, 17072. [Google Scholar] [CrossRef] [PubMed]
  131. Povirk, L.F.; Zhou, R.Z.; Ramsden, D.A.; Lees-Miller, S.P.; Valerie, K. Phosphorylation in the serine/threonine 2609-2647 cluster promotes but is not essential for DNA-dependent protein kinase-mediated nonhomologous end joining in human whole-cell extracts. Nucleic Acids Res. 2007, 35, 3869–3878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Chen, B.P.; Uematsu, N.; Kobayashi, J.; Lerenthal, Y.; Krempler, A.; Yajima, H.; Lobrich, M.; Shiloh, Y.; Chen, D.J. Ataxia telangiectasia mutated (ATM) is essential for DNA-PKcs phosphorylations at the Thr-2609 cluster upon DNA double strand break. J. Biol. Chem. 2007, 282, 6582–6587. [Google Scholar] [CrossRef] [PubMed]
  133. Toulany, M.; Schickfluss, T.A.; Fattah, K.R.; Lee, K.J.; Chen, B.P.; Fehrenbacher, B.; Schaller, M.; Chen, D.J.; Rodemann, H.P. Function of erbB receptors and DNA-PKcs on phosphorylation of cytoplasmic and nuclear Akt at S473 induced by erbB1 ligand and ionizing radiation. Radiother. Oncol. 2011, 101, 140–146. [Google Scholar] [CrossRef] [PubMed]
  134. Bozulic, L.; Surucu, B.; Hynx, D.; Hemmings, B.A. PKBalpha/AKT1 acts downstream of DNA-PK in the DNA double-strand break response and promotes survival. Mol. Cell 2008, 30, 203–213. [Google Scholar] [CrossRef] [PubMed]
  135. Szymonowicz, K.; Oeck, S.; Krysztofiak, A.; van der Linden, J.; Iliakis, G.; Jendrossek, V. Restraining AKT1 phosphorylation attenuates the repair of radiation-induced DNA double-strand breaks and reduces the survival of irradiated cancer cells. Int. J. Mol. Sci. 2018, 19, 2233. [Google Scholar] [CrossRef]
  136. Holler, M.; Grottke, A.; Mueck, K.; Manes, J.; Jucker, M.; Rodemann, H.P.; Toulany, M. Dual targeting of Akt and mTORC1 impairs repair of dna double-strand breaks and increases radiation sensitivity of human tumor cells. PLoS ONE 2016, 11, e0154745. [Google Scholar] [CrossRef]
  137. Toulany, M.; Rodemann, H.P. Potential of Akt mediated DNA repair in radioresistance of solid tumors overexpressing erbB-PI3K-AKT pathway. Transl. Cancer Res. 2013, 3, 190–202. [Google Scholar] [CrossRef]
  138. Toulany, M.; Rodemann, H.P. Phosphatidylinositol 3-kinase/Akt signaling as a key mediator of tumor cell responsiveness to radiation. Semin. Cancer Biol. 2015. [Google Scholar] [CrossRef]
  139. Philip, C.A.; Laskov, I.; Beauchamp, M.C.; Marques, M.; Amin, O.; Bitharas, J.; Kessous, R.; Kogan, L.; Baloch, T.; Gotlieb, W.H.; et al. Inhibition of PI3K-AKT-mTOR pathway sensitizes endometrial cancer cell lines to PARP inhibitors. BMC Cancer 2017, 17, 638. [Google Scholar] [CrossRef] [Green Version]
  140. Deng, R.; Tang, J.; Ma, J.G.; Chen, S.P.; Xia, L.P.; Zhou, W.J.; Li, D.D.; Feng, G.K.; Zeng, Y.X.; Zhu, X.F. PKB/Akt promotes DSB repair in cancer cells through upregulating Mre11 expression following ionizing radiation. Oncogene 2011, 30, 944–955. [Google Scholar] [CrossRef]
  141. Viniegra, J.G.; Martinez, N.; Modirassari, P.; Hernandez Losa, J.; Parada Cobo, C.; Sanchez-Arevalo Lobo, V.J.; Aceves Luquero, C.I.; Alvarez-Vallina, L.; Ramon y Cajal, S.; Rojas, J.M.; et al. Full activation of PKB/Akt in response to insulin or ionizing radiation is mediated through ATM. J. Biol. Chem. 2005, 280, 4029–4036. [Google Scholar] [CrossRef] [PubMed]
  142. Golding, S.E.; Rosenberg, E.; Khalil, A.; McEwen, A.; Holmes, M.; Neill, S.; Povirk, L.F.; Valerie, K. Double strand break repair by homologous recombination is regulated by cell cycle-independent signaling via ATM in human glioma cells. J. Biol. Chem. 2004, 279, 15402–15410. [Google Scholar] [CrossRef] [PubMed]
  143. Lim, J.; Kim, J.H.; Paeng, J.Y.; Kim, M.J.; Hong, S.D.; Lee, J.I.; Hong, S.P. Prognostic value of activated AKT expression in oral squamous cell carcinoma. J. Clin. Pathol. 2005, 58, 1199–1205. [Google Scholar] [CrossRef] [PubMed]
  144. Ringel, M.D.; Hayre, N.; Saito, J.; Saunier, B.; Schuppert, F.; Burch, H.; Bernet, V.; Burman, K.D.; Kohn, L.D.; Saji, M. Overexpression and overactivation of Akt in thyroid carcinoma. Cancer Res. 2001, 61, 6105–6111. [Google Scholar]
  145. Dobashi, Y.; Kimura, M.; Matsubara, H.; Endo, S.; Inazawa, J.; Ooi, A. Molecular alterations in AKT and its protein activation in human lung carcinomas. Hum. Pathol. 2012, 43, 2229–2240. [Google Scholar] [CrossRef] [PubMed]
  146. Altomare, D.A.; Tanno, S.; De Rienzo, A.; Klein-Szanto, A.J.; Skele, K.L.; Hoffman, J.P.; Testa, J.R. Frequent activation of AKT2 kinase in human pancreatic carcinomas. J. Cell Biochem. 2002, 87, 470–476. [Google Scholar] [CrossRef] [PubMed]
  147. McKenna, W.G.; Muschel, R.J.; Gupta, A.K.; Hahn, S.M.; Bernhard, E.J. The RAS signal transduction pathway and its role in radiation sensitivity. Oncogene 2003, 22, 5866–5875. [Google Scholar] [CrossRef] [Green Version]
  148. Ostrem, J.M.; Peters, U.; Sos, M.L.; Wells, J.A.; Shokat, K.M. K-Ras(G12C) inhibitors allosterically control GTP affinity and effector interactions. Nature 2013, 503, 548–551. [Google Scholar] [CrossRef] [Green Version]
  149. Misale, S.; Fatherree, J.P.; Cortez, E.; Li, C.; Bilton, S.; Timonina, D.; Myers, D.T.; Lee, D.; Gomez-Caraballo, M.; Greenberg, M.; et al. KRAS G12C NSCLC models are sensitive to direct targeting of KRAS in combination with PI3K inhibition. Clin. Cancer Res. 2018. [Google Scholar] [CrossRef]
  150. Cohen, Y.; Shalmon, B.; Korach, J.; Barshack, I.; Fridman, E.; Rechavi, G. AKT1 pleckstrin homology domain E17K activating mutation in endometrial carcinoma. Gynecol. Oncol. 2010, 116, 88–91. [Google Scholar] [CrossRef]
  151. Askham, J.M.; Platt, F.; Chambers, P.A.; Snowden, H.; Taylor, C.F.; Knowles, M.A. AKT1 mutations in bladder cancer: Identification of a novel oncogenic mutation that can co-operate with E17K. Oncogene 2010, 29, 150–155. [Google Scholar] [CrossRef] [PubMed]
  152. Malanga, D.; Scrima, M.; De Marco, C.; Fabiani, F.; De Rosa, N.; De Gisi, S.; Malara, N.; Savino, R.; Rocco, G.; Chiappetta, G.; et al. Activating E17K mutation in the gene encoding the protein kinase AKT1 in a subset of squamous cell carcinoma of the lung. Cell Cycle 2008, 7, 665–669. [Google Scholar] [CrossRef] [PubMed]
  153. Bleeker, F.E.; Felicioni, L.; Buttitta, F.; Lamba, S.; Cardone, L.; Rodolfo, M.; Scarpa, A.; Leenstra, S.; Frattini, M.; Barbareschi, M.; et al. AKT1(E17K) in human solid tumours. Oncogene 2008, 27, 5648–5650. [Google Scholar] [CrossRef] [PubMed]
  154. Oeck, S.; Al-Refae, K.; Riffkin, H.; Wiel, G.; Handrick, R.; Klein, D.; Iliakis, G.; Jendrossek, V. Activating AKT1 mutations alter DNA double strand break repair and radiosensitivity. Sci. Rep. 2017, 7, 42700. [Google Scholar] [CrossRef] [PubMed]
  155. Brown, J.S.; Banerji, U. Maximising the potential of AKT inhibitors as anti-cancer treatments. Pharmacol. Ther. 2017, 172, 101–115. [Google Scholar] [CrossRef] [PubMed]
  156. Narayan, R.S.; Fedrigo, C.A.; Brands, E.; Dik, R.; Stalpers, L.J.; Baumert, B.G.; Slotman, B.J.; Westerman, B.A.; Peters, G.J.; Sminia, P. The allosteric AKT inhibitor MK2206 shows a synergistic interaction with chemotherapy and radiotherapy in glioblastoma spheroid cultures. BMC Cancer 2017, 17, 204. [Google Scholar] [CrossRef]
  157. Vink, S.R.; Lagerwerf, S.; Mesman, E.; Schellens, J.H.; Begg, A.C.; van Blitterswijk, W.J.; Verheij, M. Radiosensitization of squamous cell carcinoma by the alkylphospholipid perifosine in cell culture and xenografts. Clin. Cancer Res. 2006, 12, 1615–1622. [Google Scholar] [CrossRef]
  158. Vink, S.R.; Schellens, J.H.; Beijnen, J.H.; Sindermann, H.; Engel, J.; Dubbelman, R.; Moppi, G.; Hillebrand, M.J.; Bartelink, H.; Verheij, M. Phase I and pharmacokinetic study of combined treatment with perifosine and radiation in patients with advanced solid tumours. Radiother. Oncol. 2006, 80, 207–213. [Google Scholar] [CrossRef]
  159. Gupta, A.K.; Cerniglia, G.J.; Mick, R.; McKenna, W.G.; Muschel, R.J. HIV protease inhibitors block AKT signaling and radiosensitize tumor cells both in vitro and in vivo. Cancer Res. 2005, 65, 8256–8265. [Google Scholar] [CrossRef]
  160. Wilson, J.M.; Fokas, E.; Dutton, S.J.; Patel, N.; Hawkins, M.A.; Eccles, C.; Chu, K.Y.; Durrant, L.; Abraham, A.G.; Partridge, M.; et al. ARCII: A phase II trial of the HIV protease inhibitor Nelfinavir in combination with chemoradiation for locally advanced inoperable pancreatic cancer. Radiother. Oncol. 2016, 119, 306–311. [Google Scholar] [CrossRef] [Green Version]
  161. Hill, E.J.; Roberts, C.; Franklin, J.M.; Enescu, M.; West, N.; MacGregor, T.P.; Chu, K.Y.; Boyle, L.; Blesing, C.; Wang, L.M.; et al. Clinical trial of oral nelfinavir before and during radiation therapy for advanced rectal cancer. Clin. Cancer Res. 2016, 22, 1922–1931. [Google Scholar] [CrossRef] [PubMed]
  162. Buijsen, J.; Lammering, G.; Jansen, R.L.; Beets, G.L.; Wals, J.; Sosef, M.; Den Boer, M.O.; Leijtens, J.; Riedl, R.G.; Theys, J.; et al. Phase I trial of the combination of the Akt inhibitor nelfinavir and chemoradiation for locally advanced rectal cancer. Radiother. Oncol. 2013, 107, 184–188. [Google Scholar] [CrossRef] [PubMed]
  163. Rengan, R.; Mick, R.; Pryma, D.; Rosen, M.A.; Lin, L.L.; Maity, A.M.; Evans, T.L.; Stevenson, J.P.; Langer, C.J.; Kucharczuk, J.; et al. A phase I trial of the HIV protease inhibitor nelfinavir with concurrent chemoradiotherapy for unresectable stage IIIA/IIIB non-small cell lung cancer: A report of toxicities and clinical response. J. Thorac. Oncol. 2012, 7, 709–715. [Google Scholar] [CrossRef] [PubMed]
  164. Brunner, T.B.; Geiger, M.; Grabenbauer, G.G.; Lang-Welzenbach, M.; Mantoni, T.S.; Cavallaro, A.; Sauer, R.; Hohenberger, W.; McKenna, W.G. Phase I trial of the human immunodeficiency virus protease inhibitor nelfinavir and chemoradiation for locally advanced pancreatic cancer. J. Clin. Oncol. 2008, 26, 2699–2706. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Cytoplasmic signaling cascades mediate post-irradiation cell survival and radioresistance in tumor cells by stimulating the canonical DSB repair pathways in the nucleus. Downstream effectors of cytoplasmic signaling cascades translocate to the nucleus and stimulate the canonical DSB repair pathway through either direct physical interaction with the repair protein or indirect stimulation of the pathway. IR: ionizing radiation; DSB: double-strand break; NHEJ: non-homologous end joining; HR: homologous recombination.
Figure 1. Cytoplasmic signaling cascades mediate post-irradiation cell survival and radioresistance in tumor cells by stimulating the canonical DSB repair pathways in the nucleus. Downstream effectors of cytoplasmic signaling cascades translocate to the nucleus and stimulate the canonical DSB repair pathway through either direct physical interaction with the repair protein or indirect stimulation of the pathway. IR: ionizing radiation; DSB: double-strand break; NHEJ: non-homologous end joining; HR: homologous recombination.
Genes 10 00025 g001
Figure 2. Overview of potential mechanisms by which epidermal growth factor (EGFR) stimulates DNA double-strand break repair, leading to radioresistance. (A) Ionizing radiation (IR) stimulates membrane-bound EGFR, which may lead to translocation of the receptor to the nucleus. Ionizing radiation may induce the activation of EGFR directly in the nucleus. (B) Exogenous stimulation of EGFR by related ligands, such as EGF and transforming growth factor α (TGFα), as well as autocrine secretion of EGFR ligands, such as in tumor cells harboring the KRAS mutation, induce nuclear translocation of EGFR. Following irradiation, activated EGFR in the nucleus stimulates DNA repair machinery by stimulating the phosphorylation/activation of proteins involved in DNA damage response (DDR) and DSB repair. EGFR may also function as a transcription factor regulating proteins involved in DSB repair.
Figure 2. Overview of potential mechanisms by which epidermal growth factor (EGFR) stimulates DNA double-strand break repair, leading to radioresistance. (A) Ionizing radiation (IR) stimulates membrane-bound EGFR, which may lead to translocation of the receptor to the nucleus. Ionizing radiation may induce the activation of EGFR directly in the nucleus. (B) Exogenous stimulation of EGFR by related ligands, such as EGF and transforming growth factor α (TGFα), as well as autocrine secretion of EGFR ligands, such as in tumor cells harboring the KRAS mutation, induce nuclear translocation of EGFR. Following irradiation, activated EGFR in the nucleus stimulates DNA repair machinery by stimulating the phosphorylation/activation of proteins involved in DNA damage response (DDR) and DSB repair. EGFR may also function as a transcription factor regulating proteins involved in DSB repair.
Genes 10 00025 g002
Figure 3. AKT1 stimulates DSB repair through the NHEJ and HR repair pathways. Exposure to IR induces the activation of cytoplasmic AKT1 that may translocate to the nucleus. Alternatively, IR can activate AKT1 by stimulating nuclear receptor tyrosine kinases (RTKs) independent of cytoplasmic AKT1. Activated AKT1 in the nucleus stimulates the DNA repair machinery by increasing the expression of repair proteins such as Mre11 and Rad51 or inducing the phosphorylation/activation of proteins involved in DDR and DSB repair.
Figure 3. AKT1 stimulates DSB repair through the NHEJ and HR repair pathways. Exposure to IR induces the activation of cytoplasmic AKT1 that may translocate to the nucleus. Alternatively, IR can activate AKT1 by stimulating nuclear receptor tyrosine kinases (RTKs) independent of cytoplasmic AKT1. Activated AKT1 in the nucleus stimulates the DNA repair machinery by increasing the expression of repair proteins such as Mre11 and Rad51 or inducing the phosphorylation/activation of proteins involved in DDR and DSB repair.
Genes 10 00025 g003
Table 1. Phase III clinical trials of EGFR monoclonal antibody cetuximab and early phase clinical trials of the EGFR tyrosine kinase inhibitor (TKI) erlotinib in combination with radiotherapy or chemoradiotherapy.
Table 1. Phase III clinical trials of EGFR monoclonal antibody cetuximab and early phase clinical trials of the EGFR tyrosine kinase inhibitor (TKI) erlotinib in combination with radiotherapy or chemoradiotherapy.
Target/DrugCombination Tumor TypeOutcomeReference
EGFR/CetuximabRT HNSCCImproved OS[46,47]
CRTNSCLC/Stage IIINo improved OS[53]
CRTEsophageal carcinomaReduced OS[54]
CRTHNSCCNo improved OS[55]
RT vs. CRTHPV-positive oropharyngeal carcinomaLower PFS after cetuximab + RT compared to CRT[56]
EGFR/ErlotinibRTNSCLCOS 62.5% (3 years)[57]
RTAdvanced or metastatic NSCLCOS 30% (3 years)[58]
SBRTNSCLCPFS and OS greater than historical values[59]
CRTNSCLCEffective maintenance therapy PFS 63.5%[60]
Bavacizumab + CRTHNSCCOS 71% and PFS 82% (3 years)[61]
CRTGMNo improvement in OS and PFS[62]
RT: radiotherapy; HNSCC: head and neck squamous cell carcinoma; OS: overall survival; CRT: chemoradiotherapy; NSCLC: non-small cell lung cancer; PFS: progression-free survival; SBRT: stereotactic body radiation therapy; GM: glioblastoma multiforme; HPV: human papilloma virus.
Table 2. Clinical trials of the AKT antagonists in combination with radiotherapy or chemoradiotherapy.
Table 2. Clinical trials of the AKT antagonists in combination with radiotherapy or chemoradiotherapy.
Target/DrugCombinationTumor TypeOutcomeReference
AKT/NelfinavirCRT/Phase IIPancreatic cancerAcceptable toxicity and promising activity[160]
RT/Phase IRectal cancerWell-tolerated and good tumor regression[161]
CRT/Phase IRectal cancerNelfinavir 750 mg recommended phase II[162]
CRT/Phase INSCLCAcceptable toxicity and promising activity[163]
CRT/Phase IPancreatic cancerAcceptable toxicity and promising activity[164]
AKT/PerifosineRT/Phase INSCLC, prostate, esophageal, colon, and bladder cancerRecommended phase II, 150 mg/day, started one week prior to RT[158]

Share and Cite

MDPI and ACS Style

Toulany, M. Targeting DNA Double-Strand Break Repair Pathways to Improve Radiotherapy Response. Genes 2019, 10, 25. https://doi.org/10.3390/genes10010025

AMA Style

Toulany M. Targeting DNA Double-Strand Break Repair Pathways to Improve Radiotherapy Response. Genes. 2019; 10(1):25. https://doi.org/10.3390/genes10010025

Chicago/Turabian Style

Toulany, Mahmoud. 2019. "Targeting DNA Double-Strand Break Repair Pathways to Improve Radiotherapy Response" Genes 10, no. 1: 25. https://doi.org/10.3390/genes10010025

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