Abstract
Radiation therapy (RT) is commonly used for the treatment of localized prostate cancer (PCa). However, cancer cells often develop resistance to radiation through unknown mechanisms and pose an intractable challenge. Radiation resistance is highly unpredictable, rendering the treatment less effective in many patients and frequently causing metastasis and cancer recurrence. Understanding the molecular events that cause radioresistance in PCa will enable us to develop adjuvant treatments for enhancing the efficacy of RT. Radioresistant PCa depends on the elevated DNA repair system and the intracellular levels of reactive oxygen species (ROS) to proliferate, self-renew, and scavenge anti-cancer regimens, whereas the elevated heat shock protein 90 (HSP90) and the epithelial-mesenchymal transition (EMT) enable radioresistant PCa cells to metastasize after exposure to radiation. The up-regulation of the DNA repairing system, ROS, HSP90, and EMT effectors has been studied extensively, but not targeted by adjuvant therapy of radioresistant PCa. Here, we emphasize the effects of ionizing radiation and the mechanisms driving the emergence of radioresistant PCa. We also address the markers of radioresistance, the gene signatures for the predictive response to radiotherapy, and novel therapeutic platforms for targeting radioresistant PCa. This review provides significant insights into enhancing the current knowledge and the understanding toward optimization of these markers for the treatment of radioresistant PCa.
Keywords: radiotherapy, prostate cancer, DNA repair, oxidative stress, EMT, HSP90
I. INTRODUCTION
A. Clinical Problem with Prostate Cancer Therapy
Prostate cancer (PCa) is the most commonly diagnosed cancer in males in Western countries. Risk factors for PCa development include age, race, and family history and possible risk factors include diet, lifestyle, androgens, and inflammation. One in six men will get PCa during his lifetime. The American Cancer Society estimated that there would be approximately 161,360 new cases of PCa diagnosed in the United States in 2017, resulting in an estimated 26,730 deaths.1 Although serum markers for prostate malignancy may indicate the presence of cancer, they do not predict those patients in which localized cancer is predestined to form infiltrative or metastatic cancer.
Early-stage confined PCa can be treated by radical prostatectomy, radiotherapy (RT), or watchful waiting.2 Surgery, RT, or a combination of both are the primary treatment of localized PCa. Androgen deprivation therapy (ADT) is the standard approach for patients with advanced metastatic PCa and poor prognosis.3 Despite the initial success of ADT, castration-resistant prostate cancer (CRPC) will eventually evolve, resulting in tumor recurrence and patient morbidity. To date, survival of patients with CRPC has significantly improved as a result of the development of second-generation anti-androgen therapies such as abiraterone and enzalutamide.4,5
To minimize toxic effects from treatment, low-risk patients (Gleason scores ≤ 6, prostate-specific antigen [PSA] concentrations < 10 ng/mL, or T1-T2a) are often offered an active surveillance. In contrast, treatments of patients with intermediate-risk PCa (Gleason scores = 7, PSA concentrations of 10–20 ng/mL, or T2b—c) and high-risk PCa (Gleason scores ≥ 8, PSA concentrations ≥ 20 ng/mL, or T3/4) include radical prostatectomy or some type of RT.6,7 To improve RT outcomes, adjuvant radiotherapy (co-administration of therapeutic agents with RT) and neoadjuvant therapy (administration of therapeutic agents before RT) are employed. RTOG 8610 was the first randomized phase III trial to evaluate neoadjuvant ADT that started 2 months before and then 2 months concurrently with external beam radiotherapy (EBRT) in men with locally advanced PCa.8 The ADT plus EBRT arm had statistically significant improvements in 10-year overall survival (43% vs. 34%, p = 0.12) and disease-free survival (11% vs. 3%, p < 0.001) and statistically significant reduced 10-year PCa-specific mortality (23% vs. 36%, p = 0.01), distant metastasis (35% vs. 47%, p = 0.006), and biochemical failure (defined as a rise in the blood level of PSA after treatment with surgery or RT, 65% vs. 80%, p < 0.00019) compared with the radiation alone arm. The 10-year results of the European Organization for Research and Treatment of Cancer support the addition of long-term ADT to EBRT in treating high-risk PCa.10 Although these adjuvant therapies have shown significant clinical benefit, the side effects and late complications limited the full potential of RT.
The therapeutic response to RT is determined by the intrinsic radiosensitivity (differential response of cancers to low-dose RT), that is, levels of DNA damage-repairing enzymes, and/or acquired radiation resistance, that is, increased reactive oxygen species (ROS) or antioxidants, with factors driving radioresistance linked to oxidative stress. In this review, we consider the molecular and therapeutic effects of ionizing radiation (IR) on PCa and the underlying mechanisms that could result in an increased capacity of irradiated cancer cells to overcome the anti-proliferative effects of RT and/or repair radiation damage. We also discuss recent advancements in the identification of markers of radioresistance such as DNA repair enzymes, oxidative stress markers, and protein chaperones; the impact of the tumor microenvironment; and translational strategies that integrate cancer genomics, redox state, and epithelial-mesenchymal transition (EMT) toward the development of therapeutic platforms for overcoming radioresistant PCa. The changes in these markers correspond to the resistant phenotype among different tumors.
II. EFFECTS OF IR: THE GOOD AND THE BAD
The primary cellular target of IR is chromosomal DNA. Photon radiation damages intracellular molecules by direct ionization and through indirect ionizations mediated by water radiolysis products (Fig. 1).11,12 The most common DNA damage is DNA breaks: single-strand breaks (SSBs), and double-strand breaks (DSBs).13–15 Most DSB repair is accomplished by two mechanisms: homologous recombination (HR) and nonhomologous end joining (NHEJ). NHEJ is the predominant DSB repair mechanism during the G1 phase; it repairs a direct, two-ended radiation-induced DSBs. HR is crucial for the repair of complex DNA DSBs.16 BRCA1 (a phosphoprotein that assists in 5’ to 3’ resection of DSBs and loading of RAD51) and BRCA2 are large nuclear proteins that play an integral role in the HR pathway.17 Mutations in BRCA genes have been associated with increasing the risk of PCa as well as outcomes and response to therapy.18 Therefore, these DSBs, if not re-annealed at the site of the original break, may re-anneal with other breaks on the same chromosome or a different chromosome toform chromosomal aberrations.19 γH2AX is often used as a biomarker for DBS.20 Analysis of DSB repair kinetics revealed that cancer cells had increased levels of γH2AX formation 2 h after IR.21
FIG. 1:
Effect of radiotherapy on DNA repair, mitochondrial function, cellular ROS, and EMT. IR initially causes ionization and excitation of water, leading directly and indirectly to the formation of free radicals. (A) Signaling of IR-induced DNA damage and repair. DNA can either be damaged directly by IR or indirectly through radiation-induced free radicals, which causes radical base formation, SSBs, DSBs, and DNA damage. (B) Signaling pathway of IR-induced modification of mitochondrial function. IR induces mitochondrial leakage of electrons, which interact with 0 2 to generate ROS. ROS can travel into the nucleus to cause DNA damage and G2M arrest and activate EMT signaling or to the cytoplasm to activate growth signaling and other cellular responses. Up-regulation of the mitochondrial content and mitochondrial OXPHOS subsequently leads to amplification of mitochondrial ROS and ATP production. (C) Signaling pathway of IR-induced cellular responses. Other cellular responses, including activation of ROS-generating enzymes, antioxidant proteins, growth/autophagy signaling, and chaperone proteins, are activated to contest IR-mediated cell death. (D) Signaling IR-induced EMT. ROS are implicated in IR-induced EMT via activation of signaling pathways (i.e., TGF-β, Wnt, Hedgehog, Notch, and EGFR/PI3K/Akt), which in turn activate transcription factors (i.e., Snail, Twist, and ZEB1) and induce a loss of epithelial morphology.
As mentioned above, one of the indirect actions of radiation is a multicellular effect by water radiolysis that produces free radicals in a very short period of time (< 1 sec), such as hydrated electron-ionized water, the hydroperoxyl radical, the hydrogen radical, and the hydroxyl radical.22,23 Hydroxyl radicals are highly reactive and can diffuse far enough to reach and damage biomolecules, including DNA.24,25 At least 50% of radiation-mediated DNA damage is caused by indirect effects from free radicals derived from low linear energy IR.26 To a large extent, these free radicals can initiate a chain of events that results in biological damage or further induce ROS. It has been demonstrated that ROS formation increased immediately after irradiation and continued for several hours.27 Production of hydrogen peroxide and two important species, solvated electron and hydroxyl radical, occurs in ~10−12 sec and these species remain in biological systems for times ranging from 0.1 to 1 ms.
It has been reported that IR induces secondary ROS generation via electron leakage from the mitochondrial electron transport chain complexes.28,29 It was demonstrated that IR elicits state 4 respiration and mitochondrial hyperpolarization, which results in mitochondrial ROS production.30 Therefore, IR-induced damage of mitochondrial integrity leads to mitochondrial dysfunction, ultimately inducing local and systemic oxidative stress and cell death. Conversely, McDermott et al. reported that mitochondrial ROS accumulation contributes to the survival of irradiated PCa cells.31 Radiation-mediated mitochondrial ROS production was associated with reduced levels of DNA damage, CD44 expression, and apoptotic cells.
Other sources of secondarily generated ROS after IR include NADPH oxidase, Ras- related C3 botulinum toxin substrate 1 GTPase, and ER-stress-mediated ROS mechanisms.32–34 Paradoxically, IR also promotes EMT to allow cancer cells to evade stress within the microenvironment.35,36 IR-induced ROS generation can increase the activity of transcription regulators of EMT, such as Snail, hypoxia-inducible factor-1 (HIF-1), and transforming growth factor-beta (TGF-β).37 This secondary ROS production leads to an array of biological effects, including aberrant apoptosis, genomic instability, and radiation-induced bystander effects, ultimately affecting cancer cell integrity and survival. Among the ROS, H2O2 is considered to be a molecule that activates cell signal or damages biomolecules due to its ability to diffuse across membranes.24,25
III. BIOCHEMICAL MECHANISMS UNDERLYING RADIOADAPTIVE CHANGES
The development of resistance to radiation is one of the complications of PCa treatment. Several molecular entities associated with radioresistance have been identified in PCa38–40; nevertheless, the underlying mechanisms are still inconclusive. We proposed that radioresistance phenotypes are driven by diverse molecular mechanisms including altered DNA repair, adapted mitochondrial and cytoplasmic redox states, and favorable tumor environment (Fig. 1). It is worth noting that these mechanisms are not exclusive to radioresistant PCa development.
B. Repairing of DNA Damage
DNA damage and DNA repair decide the fate of irradiated cells.41,42 The DNA damage inducible checkpoints occur in all phases of the cell cycle; G1, S, and G2 and are initiated by Ataxia telangiectasia mutated (ATM) and ATM and RAD3-related (ATR).43 ATM, in cooperation with the trimeric protein complex composed of MRE11-RAD50-NBS1 (MRN), is the earliest responder to DNA DSBs.44 Aberrations in ATM were reported in CRPC patients,45 which affected the tumor suppressor functions of ATM; that is, the damage checkpoints, programmed cell death, and DNA repair pathways.13,46 ATR is a DNA damage sensor and its absence in ATR results in deficient HR. Because RT triggers cell cycle checkpoints to block cells from entering mitosis, the apparent overexpression of ATM/ATR could simply reflect changes in the radioresistant activity of PCa. Interestingly, it has been reported that oxidative stress can also activate ATM by a mechanism independent of DNA DSBs.47
PARP-1, which is functionally involved in DNA damage repair and transcription factor regulation, is up-regulated in PCa and has been implicated in PCa radioresistance.48,49 The elevated PARP-1 expression among the tumor cells undergoing EMT implicates that DNA repair mechanisms may potentially reverse the cytotoxic effects of RT-induced DNA breaks via phenotypic programming.50,51 Moreover, the IR-mediated ROS is capable of inducing 8-hydroxy-2’-deoxyguanosine (80HdG) and reducing DNA repair52 via cysteine modification of 8-oxogua-nine DNA glycosylase 1 (OGG1).53 Accordingly, the interaction between OGG1 and PARP-1 has been described recently.54 A further example of IR-induced PARP-1 is Apurinic-apyrimidinic endonuclease 1 (APE-1), which functions as DNA repair and transcriptional regulatory proteins by facilitating binding of transcription factors to DNA.55 IR-mediated ROS production after Ca2+ mobilization via purinergic receptor-induced extracellular ATP stimulation is responsible for the localization of APE-156 and this interaction can stimulate PARP-1 activity and potentially cancer radioresistance.57
C. Alteration of Cellular Dynamics
IR-induced mitochondrial dysfunction, especially decreased electron transport chain complex activities, produces a feed-forward loop that contributes to persistent oxidative stress.58 Because the mitochondrion is an energy-generating organelle, mitochondrial dysfunction due to IR exposure would mediate alterations or adaptive responses of metabolic pathways involved in radioresistance development.59–61 Up-regulation of OXPHOS and mitochondrial ROS are correlated with radiosensitivity in several cancer cell lines.62,63 The increase in mitochondrial (mtDNA) content and mass after IR points to the up-regulation of mitochondrial biogenesis64 with increased point mutations and deletions of mitochondria in patients treated with RT.65 Petros et al. have demonstrated an association among four different mutations in cytochrome oxidase subunit I (m.6253T>C, m.6340C>T, m.6261G>A, and m.6663A>G), which has been implicated in PCa radioresistance, potentially via IR-induced increase of mitochondrial mass.66,67 Studies from other investigators demonstrated an increase of mtDNA after IR.28 Because mtDNA is prone to being attacked by IR-induced ROS, it is also potentially a source of the acquired radioresistant phenotype. Interestingly, cells depleted of mtDNA, ρ0 (rho-zero) cells are radioresistant, possibly due to suppression of the G2 checkpoint.68,69 Yamamori et al. demonstrated that IR increases mitochondrial membrane potential, mitochondrial respiration and content, and ATP production,70 which eventually lead to up-regulation of mitochondrial OXPHOS. Consistent with other evidence,70–72 Lu et al. revealed that cancer cells quickly adapt to radiation damage via mammalian target of rapamycin (mTOR)-mediated reprogramming of bioenergetics from predominantly aerobic glycolysis to mitochondrial OXPHOS via relocation of mTOR to the mitochondria and inhibiting hexokinase II.62 Conversely, mitochondria are prone to radiation damage as a result of: (1) lack of histone protection and repair mechanisms73,74 because mtDNA is 10 times more susceptible to DNA damage than nuclear DNA and (2) the proximity to the major source of these ROS. Therefore, a decrease in mtDNA content has been reported as a result of RT.75 The dichotomous role(s) of mitochondria in radiation sensitivity has been investigated in vitro and in vivo.76–79 Collectively, these data support the concept that IR targeting mitochondria plays a role, at least in part, in radiosensitivity.
A rapidly expanding body of evidence suggests that glycolysis might impede radiation treatment in radioresistant cancer cells.80 In a study by Chang et al., the glycolysis pathway was found to be activated in radioresistant DU145 and PC3 cells and radioresistant PCa xenografts.81 The investigators further found that ALDOA, a glycolytic enzyme that catalyzes the reversible reaction of fructose-1,6-bisphosphate to glyceraldehydes-3-phosphate and dihydroxyacetone phosphate,82 is up-regulated using the label-free LC-MS method in radioresistant DU145 and PC3 cells. Shimura et al. found that AKT-mediated enhanced aerobic glycolysis causes acquired radioresistance by tumor cells.83 Mechanistic evidence established that targeting H IF-lα and tumor glucose metabolism reduced the antioxidant capacity of tumors, affected the microenvironment, and sensitized various solid tumors to IR.84 Lactate dehydrogenase-5-isoenzyme, a marker of cancer anaerobic metabolism, is also significantly linked to highly proliferating prostate with biochemical failure and local relapse after RT.85
Metabolic alterations in radioresistant PCa may extend beyond an increase in glycolysis. A recent study indicated that PCa undergoes exacerbated endogenous fatty acid biosynthesis through inhibition of lipogenic enzymes such as the fatty acid synthase, ATP-citrate lysate, acetyl-coenzyme A, carboxylase 3, or stearoyl-CoA desaturase, resulting in decreased proliferation and increased apoptosis of radioresistant cancer cells.86,87 Furthermore, using an integrative molecular epidemiology approach, Kelly et al. characterized metabolic signatures at the mRNA level in prostate tumors and identified pyrimidine metabolism as the most deregulated pathways in lethal cancers, including radioresistant phenotypes.88 Taken together, these results indicate that glycolysis, lipid synthesis, and pyrimidine pathways are involved in radioresistance and targeting these pathways is likely to have broad therapeutic applications for cancer radioresistance.
D. Oxidative Stress Overloaded
The association between PCa progression and oxidative stress has been well recognized. Altered mitochondrial bioenergetics likely underlies the development of PCa and the more aggressive phenotype is directly proportional to a degree of ROS generation. PCa cells have a higher level of ROS, including oxidative stress markers, compared with normal prostate cells and the level of oxidative stress is associated with PCa occurrence, recurrence, and progression (Fig. 2).89–92 Studies have demonstrated that maintaining a balance of ROS levels in cancer is an important mechanism of radioresistance.93 The level of ROS is also critical to the radiation response; therefore, IR tends to induce ROS levels and adaptive antioxidant defense systems, which may lead to radioresistance.94–96 We have demonstrated that cancer cells are exposed to high oxidative stress at an early stage of tumorigenesis, in part due to the inhibition of various antioxidant enzyme activities such as manganese superoxide dismutase (MnSOD).92,97,98 ROS/reactive nitrogen species (RNS) cause oxidative and nitrosative/nitrative stress in cells, functioning as secondary messengers in signal transduction cascades (Fig. 2). There is a close connection between radioresistance of PCa and oxidative stress. Oxidative stress activates DNA damage, which promotes an anti-survival role, and DNA damage repair, which promotes an anti-apoptotic role. It is well established that intracellular ROS levels are increased significantly after IR exposure and that elevated levels of ROS are persistent for several hours after initial IR.99 In fact, oxidative stress in the G2/M phase lasts three times longer in irradiated cancer cells compared with non-irradiated cells.100,101 Consistent with this, H2O2 levels are increased significantly in the radioresistant phenotype of PC3 and DU145 cells,102 whereas increased lipid oxidation and malondialdehyde formation, as byproducts of IR-mediated ROS, are detected in both irradiated and bystander cells and are correlated with a significant decrease in MnSOD activity.103 Although no IR-caused changes in cellular oxidative stress have been reported,104,105 depending on tolerance level to oxidative stress, different cancer cells may exhibit a differential response to ROS elevation.
FIG. 2:
Proposed regulation of PCa radioresistance by the redox state. During the aging process, expression of antioxidant proteins (APs) decreases. The high influx of ROS/RNS and H2O2 and inactive APs secondary to aging lead to increasing oxidizing condition, oxidative stress markers (4HNE, PrxSO3), and PTMs of cysteine (Cys) residues. This PTM has a profound effect on tumor suppressors and transcription factors, which leads to mutations of redox-regulated proteins and a further decline of APs. The persistent influx of ROS/RNS drives the progression of prostatic intraepithelial neoplasia to low-grade PCa and subsequently to high-grade PCa, as indicated by higher oxidative stress markers correlated with PCa stage. A better measurement of baseline levels of the redox state would enable the development of more sensitive biomarkers and more effective redox-related treatment of PCa emerging as radioresistant disease.
Persistent oxidative stress can regulate various transcription factors/activators, such as nuclear factor (erythroid-derived 2)-related factor 2 (Nrf2), nuclear factor kappa-beta (NF-κB), and p53, thereby influencing the cell cycle and DNA repair expression and modulating cellular survival signaling pathways.93 Among redox-sensitive transcription factors, Nrf2 is recognized as a key feature for protecting cells from apoptosis, xenobiotic metabolism enzymes, and drug efflux pumps leading to the radioresistant phenotype.106 The up-regulation of Nrf2 target genes, such as heme oxygenase-1, NAD(P)H dehydrogenase (NQO1), and peroxiredoxin (Prx), have also been confirmed in relation to cancer radioresis-tance.107 Overactivation of the Nrf2-mediated defense system leads to cancer cell protection from their inherently stressed microenvironment and anti-cancer treatments including PCa, lung, breast, gall bladder, ovarian, and colorectal cancers.108 Accompanied Nrf2, NQO1, an ubiquitous enzyme involved in detoxifying pathways, can be activated by IR, thus contributing to the radioresistant phenotype.108 Significantly enough, the expression of both N Q O l and Nrf2 is increased dramatically in PCa cells intrinsically resistant to R T109 Prxs are thiol-specific antioxidant proteins that are classified largely on the basis of having either one (1-Cys) or two (2-Cys) conserved cysteine residues.110 Prx1, a major member of the 2-Cys subfamily, is also a target gene of Nrf2.111 Decreased Prx1 expression is associated with augmentation of radiosensitivity,112,113 whereas activation of Prx enhances radioresistance of cancer cells.114,115 Prx6 protein expression is increased significantly in radioresistant PC3 and D U −145 cells compared with parental cells.81 Taken together, these results strongly support that IR-generated ROS can stimulate Nrf2-associated detoxification enzymes and antioxidants, which promote cell survival under conditions of oxidative stress.108
In addition to IR-induced Nrf2 activity, the elevation of NF-κB activity in certain cancers has been associated with resistance to treatment.116 Mechanistically, ATM activation of NF-κB serves as an interexchange molecule for IR-activated NF-κB.117 It was demonstrated that ATM is essential for activation of the entire NF-κB pathway, including IKK activation, IκB-α degradation, and induction of NF-κB DNA-binding activity, in both human cells and mouse tissues after DSB activation.118 There is no direct evidence, however, to suggest that ATM-mediated NF-κB activation is required for the adaptive radioresistance phenotype. DNA-dependent protein kinase, a member of the phosphoinositide 3-kinase (PI3K)-like family of protein kinases, has also been implicated in the activation of NF-κB after exposure to IR.119 Interestingly, inhibition of IR-induced NF-κB binding is cell-type specific. For example, in PCa cells, inhibition of NF-κB by a dom inant-negative IκB m utant after exposure to a radiation dose that is equivalent to medical diagnostic use enhances apoptosis in DU145 but not in PC3 PCa cells.120,121 The induction of radiosensitivity is also mediated by several genes that are regulated by NF-κB, such as XIAP, A20, FLIP, and Bcl-xL (www.bu.edu/nf-kb/), to either regulate activation of survival pathways or to inhibit apoptotic signaling pathways.122 Overall, the critical role of NF-κB in the emergence of radioresistance in tumors after RT and fractional IR is widely recognized.
Due to IR-mediated ROS production, depletion of intracellular antioxidants such as glutathione (GSH), thioredoxin, Prx, and superoxide dismutase (SOD), can actually enhance radiation sensitivity; in contrast, up-regulation of these redox-regulating enzymes can protect radiation damage.123 Overexpression of antioxidant enzymes often occurs in cancer resistance to RT, so blocking these antioxidant enzymes enhances radiation sensitivity by promoting radiation-induced apoptosis.108,124 Radiation-resistant mice exhibit higher levels of SOD and catalase (CAT) activities compared with radiation-sensitive mice.125 MnSOD is one of the most important antioxidant enzymes located exclusively in the mitochondria and modulates cellular redox status and influences the effects of radiotherapy.126 MnSOD up-regulation has been implicated in adaptive response induced by fractionated doses of IR, leading to radioresistance.38,127 We proposed that MnSOD is one of a key NF-κB effector genes in radioadaptive resistance. NF-κB and MnSOD expressions are up-regulated in PCa cells after exposure to low-dose IR.128 Inactivation of NF-κB with an IKK-β inhibitor (IMD-0354) suppressed low-dose IR, induced expression of MnSOD, and diminished the adaptive radioresistance.128 These results indicate that NF-κB-mediated MnSOD contributes to the scavenging of radiation-induced ROS and the signaling network leading to adaptive radioresistance. We demonstrated previously that selective inhibition of RelB (proto-oncogene, NF-κB subunit) after irradiation concomitantly decreased MnSOD and sensitized PCa cells to IR.38,129–131
In Fig. 3, we propose that radioresistant PCa cells maintain their redox state toward high levels of ROS and antioxidants to enable their escape and consistent proliferation post-radiation. Therefore, the ability of radioresistant PCa cells to survive and sustain a high proliferative activity is associated with persistent ROS production, up-regulation of redox-sensitive transcription factors, and higher antioxidant levels that scavenge ROS.
FIG. 3:
Mechanistic models of redox state in radioresistant PCa. Cancer cells have higher ROS and antioxidant proteins compared with normal cells; an even higher level of ROS and antioxidants are detected in radioresistant PCa. ROS at low levels is physiologically functional, whereas at high levels, it is toxic to cells. ROS production is also activated by radiation exposure, leading to more persistent oxidative stress in cancer cells. We propose that radioresistant PCa cells adapt their redox state into higher oxidative stress via up-regulation of endogenous ROS-generating enzymes and antioxidant proteins.
E. EMT Landscape
EMT confers mesenchymal properties on epithelial cells. The ability for cancer cells to become established in regional and distant metastatic sites is dependent upon EMT. It is characterized by the loss of epithelial morphology and markers (i.e., E-cadherin, desmoplakin, Muc-1, and cytokeratin-18) and by the acquisition of mesenchymal markers (i.e., N-cadherin, vimentin, fibronectin, and vitronectin). IR is also known to enhance the metastatic potential of cancer cells by inducing EMT, which potentially mediates the emergence of the radioresistant phenotype.132 Fractionated IR has been shown to induce EMT in prostate tumors by activation of the cAMP response element binding protein and cytoplasmic sequestration of the activating transcription factor 2.133 Loss of E-cadherin is associated with attenuated radiation-induced DNA damages during hypoxia, thus contributing to radioresistance of tumor cells.134 Recently, we identified a marked decrease in E-cadherin protein expression paralleled with an increase in N-cadherin and vimentin expressions in post-RT specimens compared with pre-RT of PCa.135 The EMT phenotypic signature has been associated with biochemical recurrence and radioresistance.136
Mechanistically, IR-induced EMT is mediated by transcription factors (i.e., E12/E47, Snail/Slug, STAT3, Twistl/2, ZEB l/δEFl, and ZEB2/SIP1) that are activated by an array of signaling pathways (i.e., Hedgehog, Notch, TGF-β, Wnt, EGFR/PI3K/Akt, CXCL12/CXCR4, PAI-1, and MAPK) (Fig. 1). Among the transcriptional regulators of the EMT phenotypic landscape, Snail has been shown to play a crucial role in IR-induced EMT, with mitochondrial ROS stimulating activation of Snail and SMAD pathway through the MAPK/ERK cascade.137,138 Sustained elevation of Snail promotes mesenchymal transition after irradiation of cancer cells, which is associated with the metastatic phenotype in cancers. Furthermore, Snail shRNA prevented IR-induced mitochondrial repression and EMT switch, confirming that IR converts EMT via Snail activation.139,140 Moreover, IR can indirectly induce Wnt/β-catenin signaling to promote Snail and EMT, which leads to the invasiveness of progeny from irradiated cancer cells.141,142 Notch signaling is also implicated in IR-induced EMT; IR activates the interleukin-6 (IL-6)/JAK/signal STAT3 pathway to up-regulate Notch-1 and subsequently induces EMT.143,144 IR-induced EMT is also mediated by EGF, TGF-β, vascular endothelial growth factor, ROS, and others.22,145–147 For example, TGF-β-mediated EMT, in accordance with ATM, is capable of activating Homeobox B9 for enhancement of DNA damage and repair responses, leading to radioresistance.148 M addition, TGF-β navigates disruption of the mitochondrial complex IV activity, thus prolonging the production of mitochondrial ROS which delay cell growth as a mechanism for cell death resistance.149 Moreover, the balance between superoxide radical (O2•−) and H2O2 can potentially determine pathways that drive the EMT process and its reversal to MET.25 Work by Quiros-Gonzalez et al. showed that MnSOD affects EMT inter-conversion by modulating the rate of H2O2 production and the balance between O2•− and H2O2 in androgen-dependent prostate cancer LNCaP cells with stable expression of MnSOD.150 Indeed, treatment with the N-acetylcysteine, a general ROS scavenger, prevents IR-induced EMT.151 Activation of the redox-sensitive transcription factor NF-κB by ROS is also another route for ROS-mediated EMT,152 supporting an inter-exchange role for ROS in IR-induced EMT.
IV. PREDICTIVE MARKERS OF RESISTANCE TO RADIOTHERAPY
It is well recognized that substantial heterogeneity of the radiation response in normal tissues and PCa exists among individual patients.153–155 Even within a single cancer, different regions can have a gradient of radiosensitivity depending on the microenvironment, distribution of cancer stem cells, and/or genetic alterations.156 Several lines of evidence suggest difference in selected proteins of PCa patients after radiation exposure.157,158 Identification of radiation response markers could be used for monitoring the progress of RT and prediction of RT therapeutic outcome.159 The clinical and genomic databases can be used to develop and validate a gene expression. A signature to categorize patients as responders or non-responders to local therapy on initial diagnostic biopsies would be of significant clinical use.
A. Gene Signature Predictive of PCa Radioresistance
Utilization of the genetically diverse NCI-60 cell lines, led to the first microarray analysis of gene expression in response to IR and correlation of the gene profiling with clonogenic survival. The following genes were identified: 22 genes associated with low survival after 2 Gy γ-rays, 14 genes associated with low survival after 8 Gy, and 25 genes with radiation responses dependent on wild-type p53.160 Further, mapping of known DNA interactions with promoters of mitosis genes revealed a regulatory network centered on the transcription factors E2F4, E2F1, RBL2 (the p130 retinoblastoma-like protein), and TAF1. Among them, E2F4 and RBL2 act together as a repressor complex to prevent gene transcription.161 Induction of E2F4 protein has been reported (8–24 h) after irradiation of PCa cell lines162 and its activity is changes during normal cell cycle progression163 and in response to IR.164
Recent studies in PCa patients who received either image-guided radiation therapy (IGRT) or radical prostatectomy as primary therapy revealed that RNA-based gene signatures could differentiate indolent and non-indolent and lethal PCa based on the alterations in the following genes: MYC, NKX3–1, PTEN, and STAR.165–168 Interestingly enough, TMPRSS2-ERG fusion status does not predict prognosis after radical prostatectomy or IGRT169,170 Zhao et al. utilized high-throughput gene expression and clinical data to develop a 24-gene postoperative radiotherapy outcomes score (PORTOS) for monitoring the likelihood of developing postoperative radiotherapy metastasis at 10 years.171 PORTOS is further designed to enrich for intrinsic radiation response by including genes related to radiation or DNA damage response.172 Accordingly, PORTOS (overexpression of the genes DRAM1, GSEA, KRT14, PTPN22, ZMAT3, ARHGAP1, IL-1B, ANLN, RPS27A, MUM1, TOP2A, CDKN3, HCLS1, DTL, IL-7R, UBA7, NEK1, CDKN2AIP, APEX2, KIF23, SULF2, PLK2, EME1, and BIN2) is the first validated gene signature developed to predict response to therapy in PCa, with a potentially high clinical value in selecting patients for radiotherapy.
B. Oxidative Stress Markers
1. Oxidative Damage Products
Oxidative damage products are a nonspecific systemic index of oxidative stress in the body. In cancer patients undergoing any cancer treatment, markers of oxidative stress in the form of lipid peroxidation and protein carbonyl content are often elevated.173,174 Alteration of antioxidants and accumulation of oxidative damage products have been demonstrated during PCa development in human91,92 and mouse (Nkx3.1 or Nkx3.1/PTEN knockout) models.175 Cells with PTEN loss exhibited decreased antioxidants and redox regulatory proteins and increased oxidative damage products.176 Our data from PCa cell culture and human tissues clearly show an association between increased oxidative stress markers including 8OHdG, lipid peroxidation 4-hydroxylnonenal (4HNE), and 3-nitrotyrosine (3NT), with both high Gleason score and tumor stage.91 Levels of 8OHdG were focally elevated in the nuclei of primary cancer compared with the benign epithelium, whereas metastatic PCa showed high levels of nuclear 8OHdG among cells. In contrast, metastatic PCa exhibited nuclear 4HNE protein adducts, whereas 3NT was detected in the cytoplasm. As a result, the persistent markers of increased oxidative stress may contribute to redox imbalance and promote PCa progression and recurrence after RT.
2. Hypoxia
Given the essential role of oxygen in the response of IR-induced ROS, hypoxia becomes a major problem with RT.177,178 It is well established that tumor radioresistance is conversely correlated with the distance from blood vessels; cancer cells near blood vessels are sensitive to IR compared with distant tumors.179 IR resulted in acute hypoxia in the tumor due to transient changes in blood flow and this acute hypoxia increased radioresistance in the tumor.180,181 Higher radiosensitivity of normoxic compared with hypoxic cells is most likely due tofixation of ROS under normoxic conditions.182,183 Preclinical studies support that hypoxia leads to a radioresistant and metastatic phenotype of prostate tumors.184 PCa patients with hypoxic tumors had recurrent disease after treatment with radical prostatectomy or IGRT.185 A review by Deep and Panigrahi has summarized the crucial role of the major hypoxia signals, HIF, PI3K/Akt/mTOR, NOX, Wnt/β-catenin, and Hedgehog, in PCa growth and progression.186 Among these signals, extracellular vesicles called “exosomes” are the latest marker in mediating hypoxia-induced PCa progression, potentially via tumor microenvironment remodeling. Exosomes have been reported in hypoxia-induced angiogenesis, cancer sternness, activation of cancer-associated fibroblasts, and EMT.187 The exosomes secreted under hypoxia enhance invasiveness and stemness of PCa cells by targeting adherent junction molecules and increased level of diverse signaling molecules (TGF-β2, TNFlα, IL-6, TSG101, Akt, ILK1, MMP, and β-catenin).187 Therefore, exosomes serve as vehicles that transfer bioactive molecules between cells and mediate cell-cell communication during hypoxia-derived RT.
Recent radical prostatectomy mRNA cohorts (108 Memorial Sloan Kettering Cancer Center cohort patients and 110 Cambridge patients) by Lalonde et al. revealed that hypoxia RNA signatures are univariately prognostic.188 However, when patients were separated into four groups on the basis of high versus low percentage of genome alterations and high versus low hypoxia values, it was found that hypoxia increased the prognostic accuracy of the percentage genome alteration for risk of biochemical relapse after radical prostatectomy.185,189 Therefore, measurement of cancer hypoxia and genomic instability can improve the prognostic capability of a pretreatment biopsy.189 Ample evidence supports the crucial role of these hypoxia RNA signatures. Evaluation of cancer hypoxia using [18F] fluoromisonidazole-positron emission tomography prior to radiochemotherapy has become a powerful discriminator between groups with good versus poor prognosis after RT.190 Because hypoxia diminishes the therapeutic efficacy of RT, the development of a platform to increase oxygen concentrations within the tumor, including fractioned RT, hypoxic cell sensitization, and pharmacological reduction of oxygen consumption, has been pursued.191–193
3. IL-8
IL-8, also known as CXCL8, is a chemoattractant chemokine that has oxidative stress as one of its secretion stimulators.194 The effects of IL-8/IL-8 receptor signaling pathways in PCa progression and radiation sensitivity may be orchestrated by communication and/or interaction with chemokines and their receptors such as CXCR1–7.195 Increased IL-8 expression is associated with a high Gleason score, tumor pathologic stage, and markers of advanced PCa including castration resistance and radioresistance.196 Because the transition of PCa to radioresistant PCa is partially due to IL-8-signaling-induced HSP90 activation, identifying IL-8 expression and its down-regulated signaling pathways may be used as a marker for radioresistant PCa regardless of AR response.197
C. “Marker Dictations” by the Tumor Microenvironment
1. EMT Signature
The expression of critical proteins of defining the phenotypic landscape of EMT, E-cadherin, N-cadherin, and vimentin are indicators of PCa cells with invasive characteristics. Based on knowledge about the EMT signaling pathway, several molecules could be up-regulated upon radioresistant phenotype development; these include TGF-β, Wnt, Hedgehog, Notch, EGFR/PI3K/Akt, MAPK, and p21-PAK1.132 In turn, these signals contribute to the induction of Snail expression to promote EMT and invasion.198 For example, vimentin, which is a symbol of the acquisition of mesenchymal characteristics, is found to be increased in radioresistant PCa cells (PC3-RR, DU145-RR, and LNCaP-RR) compared with their parental PCa counterpart (up to a 7-fold increase).81 Slug, also known to inhibit p53-mediated apoptosis, is up-regulated after IR of radioresistant cancers.199 ZEB1 is known to confer radioresistance by binding to the USP7 deubiquitinase to stabilize CHK1, thereby activating the recombination-dependent DNA repair response.200 Further, high expression of Wnt signaling activity is associated with increased radioresistance in colorectal cancer cells and intestinal stem cells,201 whereas nuclear β-catenin correlates with poor clinical outcomes after RT.202
An EMT biomarker signature tailored to the emergence of radioresistance can be created and globally applied for the radioresistant response of various cancers, including PCa. It is well established that loss of E-cadherin is considered a hallmark of EMT. Therefore, loss of E-cadherin combines with other EMT signature markers that are correlated with a radioresistant phenotype would be potential candidates for radioresistant PCa markers. For instance, if E-cadherin is negative and Snail is positive, then the signals that contribute to induced Snail expression (i.e., Wnt, Hedgehog, Notch, EGFR/PI3K/Akt, and MAPK) maybe defined as surrogate markers of the radioresistant phenotype in cancer patients.
2. TGF-β
TGF-β is a ubiquitous growth factor that plays an important role in regulating injury response and in the progression of PCa. There are three TGF-β isoforms: TGF-βl, TGF-β2, andTGF- β3. Functionally, TGF-β is considered to be a molecular switch responsible for excess synthesis and deposition of collagen and other ECM proteins via the TGF-β receptors and the SMAD3 pathway. Once activated, SMADs translocate into the nucleus, where they regulate the transcription of genes involved in cell cycle arrest. During the early phases of tumor progression including tumor initiation, TGF-β plays role as a tumor suppressor.203 Work from our group demonstrated that disruption of TGF-β signaling in vivo accelerated pathologic malignant changes in the prostate by altering the tumor growth kinetics and inducing EMT.204 Interestingly, at the later stage, TGF-β promotes processes associated with tumor aggressiveness such as cell invasion, dissemination, EMT, and immune evasion.205,206 IR induces the release and activation of TGF-β in cells and tissues.207 A mechanistic study in a cell-free system demonstrated that oxidation of the TGF-β latent complex acts as a sensor of oxidative stress to mediate the release and activation of TGF-β and orchestrates cellular responses to IR.208 In turn, TGF-β stimulates ROS production by activation of p22 (phox) subunit of NADPH oxidases/ATM/p53 axis, which initiates downstream signaling targets (e.g., SMADs and EGFR) and results in the expression of a subset of pro-fibrotic genes.209 TGF-β also induces a stromal redox state imbalance as a result of elevated NADPH oxidase 4-dependent ROS production and inhibits the expression of the MnSOD and CAT.210
TGF-β exerts differential effects during the early versus late stages of PCa progression. As a predictive marker, TGF-β up-regulation has been associated with tumor grade, pathologic stage, and lymph node metastasis in PCa patients.211,212 Although no difference in the serum levels of TGF-β has been detected between benign prostate hyperplasia and PCa, elevated levels of plasma and urinary TGF-β were found in patients with poor clinical prognosis.213 Considering the role of TGF-β in radiation-induced PCa progression, TGF-β could serve as a marker of radioresistant PCa.
3. PARP- 1
PARP-1 is a nuclear protein with multiple cellular functions including DNA damage repair.14 Herein, we redirected our focus to PARP-1 as an IR-mediated EMT marker. A recent study demonstrated that inactivation of PARP-1 by gene-targeted deletion in an in vivo model of PCa leads to EMT induction toward high-grade PCa.214,215 PARP-1 immunostaining increased dramatically in PCa cells associated with the EMT phenotype (mesenchymal) in response to RT.48 Statistical evaluation of PARP-1 expression data among the mesenchymal cells in PCa after RT revealed that loss of PARP-1 is associated with biochemical recurrence and poor response to RT.216 Work from our group revealed that loss of PARP-1 is strongly associated with prostate tumorigenesis in vivo via TGF-β-mduced EMT48 and poor therapeutic response to RT.135 Recent clinical studies utilizing PARP-1 inhibitors in the treatment of CRPC support targeting PARP-1 as a means to sensitize PCa to AR-directed therapeutics.217 PARP-1 inhibitors can be of clinical value in combination with chemotherapy or RT to sensitize cancer cells to genotoxic insult.218 Evidence on tumor radiosensitization by PARP inhibition provides indirect support for a role of PARP-1 in promoting radioresistance in PCa.50
D. Chaperone Proteins: Heat Shock Protein 90
Among the chaperone proteins, HSP90 is of particular interest due to the ~100-fold elevation in human cancers.219 HSP90 is an ATP-dependent chaperone that plays a central role in refolding and maintains the stability of diverse proteins (~725 client proteins), many of which are essential for cancer growth and confer therapeutic resistance (e.g., Akt, Src kinase, mutant p53, cyclin-dependent kinase 4, and H IF-lα220,221). HSP90 is often overexpressed in cancer due to mutation, amplification, deletion, methylation, and post-translational modifications (PTMs).222 According to the Cancer Genome Atlas, the copy number of an HSP90AA1 gene is not altered significantly, but its protein expression is increased significantly in PCa.223 Using a proteomics approach, it was shown that HSP90-mediated PI3K/Akt/mTOR signaling pathway is an important signaling pathway regulating radioresistant PCa cells.224 The HSP90-mediated PI3K/Akt/mTOR signaling pathway was identified as a main pathway associated with PCa cell radioresistance.81 In addition, Quanz et al. reported that HSP90α is both a chaperone and substrate of DNA-dependent protein kinase (DNA-PK). After TRAIL-induced apoptosis, DNA-PK is the sole kinase that phosphorylates HSP90α (at the threonine-5 and −7 positions).225
V. NOVEL TARGETING TO OVERCOME PCA RADIORESISTANCE
Identification of a patient’s susceptibility for developing adverse effects from RT is an important prerequisite for the personalized treatment era. Previously, p53/MDM2, Bcl-2/Bax, PTEN/Akt, and COX-2 were investigated extensively as radiation sensitization targets of PCa (as reviewed by Palacios et al.226). From 1997 to 2008, 11 biomarkers for radiation resistance were studied by the Radiation Therapy Oncology Group.227 mRNA microarray studies revealed more genes altered after IR, ineluding IL-2 and sequence-specific DNA-binding proteins.160 Targeting a single signaling pathway that drives radioresistance is not sufficient to improve RT efficacy, but a multi-targeted approach by simultaneously targeting DNA repair and chaperone protein would improve therapeutic outcome for radioresistant PCa.
A. PARP-1 in Charge of DNA Repair
The lethality of unrepaired DNA DSBs, commanded the development of agents to target selectively proteins involved in the DNA damage response. Gavande et al. developed a radiation sensitivity molecular signature based on gene expression profiling datasets and identified 10 genes as DSB repair genes.228 Several clinical trials combining PARPI’s with radiation have been established.13,229 Agents that inhibit DNA repair indirectly are being combined with PARP inhibitors (PARPi’s), such as inhibitors of EGFR, PI3K, HSP90, and agents that inhibit the G2 checkpoint, CHK1, and WEE1.230–232 Agent pairs such as HSP90-PARP are more effective when combined with RT,233 indicating that radiation-induced DNA damage can potentiate therapeutic response. The comprehensive genetic characterization of PCa identified a significant number of germline and somatic mutations in DNA repair genes.234 Given the frequency of these DNA mutations in metastatic CRPC, the paradigm of PARPi synthetic is a potentially important addition to the management of patients with radioresistance disease.235 Current PARPi’s undergoing clinical evaluation are veliparib, olaparib, niraparib, talazoparib, rucaparib, CEP-9722, and SC10914.236 A clinical benefit of PARPi’s in PCa was detected in a phase I trial of olaparib in 60 patients with solid tumors who were germline BRCA mutation carriers.237 Selected mutations could inhibit PARPi effectiveness; for example, the somatic loss of functional mutations in the tumor suppressor PTEN results in impaired PARPi sensitivity in PCa cells.238–241 Moreover, PARPi’s can impair AR recruitment to target promoters.242 In contrast, a phase II trial (NCI 9012), reported in the Journal of Clinical Oncology, found no benefit of adding veliparib to abiraterone (Zytiga) plus prednisone in CRPC patients with ETS fusion-positive tumors. Therefore, ETS status was not predictive of response to veliparib. Interestingly, exploratory analysis led to the novel and unexpected finding that a DNA-damage repair defect (not PARP-1) was associated with improved outcomes with abiraterone (Zytiga) plus prednisone treatment, possibly through induction of a synthetic lethality in the context of HR defects.243
B. Cellular Redox State: MnSOD, the Player
MnSOD is one of the important antioxidant enzymes that is located exclusively in the mitochondria244 and its expression and activity modulate cellular redox status, influencing the effects of radiotherapy. In vivo, the administration of MnSOD-liposome (MnSOD-PL) reduced GPx activity significantly in conjunction with radiation, but did not affect radiation-induced reduction of GSH in oral cavity orthotopic tumor tissue. Intraoral administration of the MnSOD-PL therapy decreased the number of ulcerations significantly 5 days after irradiation compared with control mice.245 Overexpression of MnSOD failed to prevent bladders from the early off-target effects of IR, but enhanced recovery at 4 weeks after IR.246 In vivo, intraperitoneal administration of recombinant MnSOD (rMnSOD) after irradiation with subsequent daily injection of rMnSOD led to a significant decrease of IR-induced off-target injury and increased survival up to 30 days after IR compared with controls, which died 7–8 days after IR.247 Radioprotective thiols can modulate TNF-α-induced MnSOD gene expression with NF-κB activation.248 MnSOD affects the release of cytochrome C from mitochondria due to alterations in ROS levels in mitochondria.249 Our recent studies, which are supported by the work of other investigators, demonstrated that selective inhibition of RelB-induced MnSOD after irradiation can sensitize PCa cells to radiation treatment.130,250–252
Given the significance of MnSOD, manganese porphyrin compounds have been developed with variable O2•−-scavenging properties, pharmacokinetics on tissues, and subcellular localization.253 The impact of SOD mimetics on increasing radiosensitivity in cancers stems from evidence that Mn(III) 5.10.15.20-tetrakis(N-ethylpyridinium −2-yl)por- phyrin) (MnP-2-PyP5+) inhibits the radioprotective effects of cancer-cell-conditioned medium on endothelial cells in vitro and enhanced radiation-induced damage to tumor vasculature and delayed tumor growth in vivo.254 These results suggesting that MnP-2-PyP5+ may disrupt communication between cancer and endothelial cells, leading to the breakdown of tumor vasculature and disruption of the microenvironment after IR. Furthermore, Mn(III) 5.10.15.20-tetrakis(N-hexylpyridinium-2-yl)por phyrin) (MnTnHex-2-PyP5+), a more potent SOD mimetic due to its greater lipophilicity and higher biodistribution,255 exerted a radiosensitizing effect on HeLa cells, murine mammary carcinoma 4T1 cells, and melanoma B16 cells in vitro and in vivo via attenuated DNA damage repair and triggered a shift from pro-survival pathways to apoptosis.127,256 In addition to their radiation enhancer effects, SOD mimetics exert radioprotective effects in normal cells, the most powerful SOD mimetic, Mn(III) meso-tetrakis(N-n-butoxyethyl-pyridinium −2yl) porphyrin, MnTnBuOE-2-PyP5+) (MnP), has been demonstrated to protect bone marrow suppression from irradiation via inhibiting oxidative damage.257 Recent studies by Oberley-Deegan et al. demonstrated that MnP protects against prostate radiation-induced injury to male reproductive system via up-regulation of Nrf2.258 Therefore, SOD mimetics boost ROS generation in cancer cells selectively toward oxidative stress overload while stimulating adaptive responses in normal cells and thus emerge as attractive candidates for adjuvant therapy with RT in PCa patients (Fig. 4).
FIG. 4:
Schema of cell death induction by pro-oxidants. Generally, levels of ROS/RNS are significantly higher in cancer cells than in normal cells. Pro-oxidant treatments with, for example, SOD mimetics prompt an influx of ROS/RNS into cells and shift the cellular redox state of cancer cells into oxidative stress and that of normal cells into a mild pro-oxidant state. Moderate levels of ROS in normal cells induce adaptive responses that activate Nrf2-mediated antioxidant proteins, which protect normal cells from ROS overload. Cancer cells harboring endogenously high levels of ROS, and with further increased ROS, could undergo apoptosis consequential to protein degradation and DNA damage via intrinsic mitochondrial-mediated apoptosis signaling.
C. Redox-Sensitive Transcription Factor: RelB
NF-κB transactivation is induced in many types of cancers by therapeutics, including radiation. The non-canonical dimer p52/RelB plays a more important role in protecting PCa cells against radiation compared with p50/RelA.130 Studies from our group indicated that selective inhibition of RelB enhances the radiosensitivity of PCa cells.129,252 Mechanistic evidence indicates that selective blockade of RelB nuclear import by a cell-permeable SN52 peptide, a variant of the SN50 peptide blocking nuclear import of NF-κB family members, is highly effective in sensitizing PCa to radiotherapy.251,258 The RelB/sirtuin (SIRT3)/MnSOD axis has been identified as a critical contributor to ascorbic-acid-induced radiosensitization of PCa cells and radioprotection of normal prostate cells. RelB suppression decreases expression of SIRT3 and MnSOD, which in turn increases oxidative and metabolic stresses in PCa cells; in contrast, ascorbic acid enhances RelB expression in normal prostate, improving antioxidant and metabolic defenses against radiation. Therefore, targeting RelB in PCa is a viable approach to enhance RT efficacy selectively.61
D. TGF-β in Control of the EMT Landscape
Radioresistance augments prostate tumor progression by promoting EMT, metastatic spread, and evasion of immune surveillance, all processes that are regulated by TGF-β. Considering that IR induces the release and activation of TGF-β and the association of poor prognosis in PCa and blood levels of TGF-β,259 the use of TGF-β inhibitors to ameliorate radiosensitization in PCa is expected. A small-molecule inhibitor of the TGFβ type I receptor kinase (LY364947) has been shown to increase radiosensitivity in vitro and in vivo.260 LY364947 promotes clonogenic cell death and decreases radiation-induced phosphorylation of γH2AX and p53. Several studies indicated that inhibition of TGF-β signaling with a selective inhibitor of a TGF-β type I receptor, ALK5, enhances radiosensitivity in glioblastoma by disrupting EMT and self-renewal capabilities, blocking DNA damage response, and increasing apoptosis.261 Administration of a 2G7-neutralizing pan-TGF-β IgG2 antibody blocked IR-induced circulating levels of TGF-β 1, circulating tumor cells, and lung metastases in transgenic model of metastatic breast cancer.262 Reduction in metastasis is attributed to auto crine TGF-β production in the tumor microenvironment.263 Galunisertib (a TGF-β receptor inhibitor; www.clinicaltrials.gov identifiers NCT01246986 and NCT01373164) is undergoing a phase II trial for pancreatic ductal adenocarcinoma and hepatocellular carcinoma.264 A putatively hypofunctional TGF-β 1 haplotype is associated with lower TGF-β plasma levels and increased sensitivity to radiation-induced chromosomal aberrations and apoptosis in lymphoid cells.265,266 Increased circulating TGF-β levels during RT have been correlated with poor prognoses in patients with non-small-cell lung cancer.267 RT-induced TGF-β is implicated in tissue damage such as fibrosis268 and the use of TGF-β inhibitors has been shown to ameliorate IR toxicity to normal tissues.269 Blocking of TGF-β signaling before irradiation attenuates DNA damage, increases cell death, delays tumor growth, and prolongs survival in several human cancers.60,260 Taken together, this evidence suggests that targeted disruptions of TGF-β signaling can radiosensitize prostate tumor cells while inhibiting metastasis.
E. Protein Homeostasis: All Eyes on HSP90
HSP90 has emerged as a promising target in cancer treatment, including RT, for more than a decade. To date, there are at least 20 HSP 90 inhibitors in oncology trials (www.clinicaltrials.gov). Numerous studies have indicated that HSP90 inhibitors can sensitize tumor cell lines and tumors growing in vivo to clinically relevant doses of X-ray or photon exposure.270 The HSP90 inhibitors NVP-AUY922 and 17-(dimethylaminoethylamino)-17-demethoxygeldanamycin have been shown to induce G2M arrest by down-regulating Cdk1 and Cdk4 and alter cell cycle distribution by affecting cell cycle checkpoints.271,272 Currently, ganetespib, which is safe and well tolerated, has the potential to become the first-in-kind HSP90 inhibitor to be approved as a radiosensitizer in PCa273 because it was shown to be effective in clinical trials of rectal cancer (www.clinicaltrials.gov identifier NCT01554969) and esophageal cancer (www.clinicaltrials.gov identifier NCT02389751).
If HSP90 is cleaved, cross-linked, or changed in conformation due to site-specific PTM, it could prevent the binding of ATP, client proteins, or co-chaperones and activate ubiquitination, sumoylation, or selected AAA-type proteinase, leading to radioresistant cancer cell death.222,274 Therefore, utilizing ROS to modify the HSP90 protein level and function via ROS-mediated PTM is a logical tactic to additionally inhibit HSP90. Clinically, ErbB3 expression is correlated with resistance to HSP90-inhibitor-induced tumor cell radiosensitization.275 The distinct pattern of HSP90 subcellular compartmentalization in the cytoplasm, nucleus, endoplasmic reticulum, and mitochondria can regulate different sets of client proteins dynamically.276 In PCa, mitochondrial HSP90 interacts with proteins essential for survival and metabolism.277 Gamitrinib, a mitochondria-targeted, small-molecule HSP90 inhibitor, is active against all tumor types tested and efficiently killed metastatic, CRPC, and multidrug-resistant PCa cells characterized by overexpression of the ATP-binding transporter P-glycoprotein.278 Targeted inhibition of mitochondrial HSP90 suppresses localized and metastatic PCa in the TRAMP mouse model of PCa progression.279 Shown in Fig. 5 is the approach of simultaneously targeting cytoplasmic and mitochondrial HSP90 during RT. Inhibition of HSP90 can sensitize cancer cells to fractionated IR and block selection of residual tumor-radioresistant cells.
FIG. 5:
Multi-targeting HSP90 therapeutic approach. Mechanistically, gamitrinib prevents mitochondrial HSP90-binding activity, thus inhibiting mitochondrial function by inducing acute mitochondrial dysfunction in PCa cells, with a loss of organelle membrane potential and release of cytochrome C and caspase activity independently of proapoptotic Bcl-2 associated proteins, Bax/Bak. In contrast, ganetespib prevents cytoplasmic HSP90-binding activity and thus inhibits multiple processes, including: (1) growth signaling; cyclin proteins, and MET/EGFR; (2) angiogenesis: HIF-lα, vascular endothelial growth factor (VEGF), and Src; and (3) survival signaling navigated by AKT and IGF-1R. Systemic administrations of both gamitrinib and ganetespib during RT would target effects of radioresistance and prevent the adaptive feedback response or compensation mechanism due to targeting a single HSP90.
VI. CONCLUSION
PCa is one of the most lethal cancers in men in the United States today. Understanding the underlying mechanisms of therapeutic resistance in PCa patients, including radioresistance, will aid in strategies to overcome treatment failure, minimize toxicity, and increase patient survival. The challenges surrounding increased radiation intensity to improve total control of cancer cell growth in prostate tumors relate to significant risks of major side effects associated with radiotherapy. An attractive radiation enhancer would provide exploitation of the intrinsic differences in the radioresponse of cancer epithelial cells and normal cells. DNA repair systems, oxidative stress, HSP90, and EMT are critical for the clinical radiation response of PCa. The multilayered complexity and the functional interactions of these networks support a multi-targeted platform to improve the therapeutic outcomes for radioresistant PCa. Concurrent targeting of the expression and the topological localizations of effectors such as HSP90 in mitochondrial and cytoplasmic compartments (Fig. 5) may provide concomitant protection of normal tissues while sensitizing PCa to IR. ROS, at least in part, regulate DNA repair systems, HSP90, and EMT, thus pushing oxidative stress beyond the cancer cell antioxidant capacity to alter multiple ROS-associated proteins. This represents a promising platform for enhancing RT efficacy by killing cancer cells while protecting normal cells (Fig. 4).
In summary, several combination strategies are currently in clinical trials with great anticipation for overcoming radioresistant PCa and improving patient survival and quality of life. As technology advances, the framework of molecular signatures could be embedded commercially in the clinical platform, integrating it as a predictive factor for the outcome of post-RT to facilitate biomarker-driven treatment approach toward personalized therapies.
ACKNOWLEDGMENTS
Research in the authors’ laboratory was supported in part by the National Institutes of Health (Grants CA188792 to L.C., CA205400–02 to D.S., and National Institute of General Medical Sciences COBRE Program P20 Grant GM121327 to L.C. and D.S.) and the James F. Hardymon Foundation (N.K.).
ABBREVIATIONS:
- 3NT
3-nitrotyrosine
- 4HNE
4-hydroxylnonenal
- ADT
androgen-deprivation therapy
- APR-1
apurinicapyrimidinic endonuclease 1
- AR
androgen receptor
- ATM
ataxia telangiectasia mutated
- ATR
ATM and RAD3-related
- BER
base excision repair
- BRCA1
phosphoprotein that assists in 5’ to 3’ resection of DSBs, loading of RAD51
- BRCA2
phosphoprotein that assists with RAD51 loading
- CAT
catalase
- CRPC
castration-resistant prostate cancer
- DSB
double-strand breaks
- e−aq
hydrated electron
- EBRT
external beam radiotherapy
- EMT
epithelial mesenchymal transition
- GSH
glutathione
- H•
hydrogen radical
- H2O+
ionized water
- H2O2
hydrogen peroxide
- HIF-lα
hypoxia inducible factor 1 alpha
- HO2
hydroperoxide radical
- HR
homologous recombination
- HSP90
heat shock protein 90
- IGRT
image-guided radiation therapy
- IL
interleukin
- IMRT
intensity-modulated radiation therapy
- IR
ionizing radiation
- LNCaP
androgen-dependent prostate cancer cell
- MDC1
mediator of DNA damage checkpoint protein 1
- MnP
Mn(III) meso-tetrakis(N-n-butoxyethyl-pyridinium-2yl)porphyrin
- MnP-2-PyP5+
(Mn(III) 5,10,15,20-tetrakis(N-ethylpyridinium-2-yl)porphyrin
- MnSOD
manganese superoxide dismutase
- MnSOD-LP
MnSOD-liposome
- MnTnHex-2-PyP5+
Mn(III) 5,10,15,20-tetrakis(N-hexylpyridinium-2-yl)porphyrin
- MRN
MRE11-RAD50-NBS1
- mtDNA
mitochondrial DNA
- mTOR
mammalian target of rapamycin
- NEIL2
nei-like DNA glycosylase
- NF-κB
nuclear factor kappa-beta
- NHEJ
nonhomologous end joining
- NQO1
NAD(P)H dehydrogenase
- Nrf2
nuclear factor (erythroidderived 2)-related factor 2; O2•−, superoxide radical
- O H•
hydroxyl radical
- OXPHOS
oxidative phosphorylation
- PARP
polyadenosine diphosphate (ADP) ribose polymerase
- PARPi
PARP1/2 inhibitor
- PCa
prostate cancer
- PI3K
phosphoinositide 3-kinase
- Prx
peroxiredoxin
- PrxSO3
oxidation form of Prx
- PTEN
signaling, phosphatase and tensin homolog
- PTM
posttranslational modification
- rMnSOD
recombinant MnSOD
- RNS
reactive nitrogen species
- ROS
reactive oxygen species
- RT
radiotherapy
- SBRT
stereotactic body radiotherapy
- SIRT3
sirtuin
- SOD
superoxide dismutase
- SSB
single-strand break
- TGF-β
transforming growth factor-beta
- ρ0
rho-zero cells
REFERENCES
- 1.Siegel RL, Miller KD, Jemal A. Cancer Statistics, 2017. CA Cancer J Clin. 2017;67(1):7–30. [DOI] [PubMed] [Google Scholar]
- 2.Altwein JE. Therapeutic options in locally defined or advanced prostate cancer. Eur Urol. 1999;35 Suppl 1: 9–15; discussion 6. [PubMed] [Google Scholar]
- 3.Heidenreich A, Aus G, Bolla M, Joniau S, Matveev VB, Schmid HP, Zattoni F ; European Association of Urology. EAU guidelines on prostate cancer. Eur Urol. 2008;53(1):68–80. [DOI] [PubMed] [Google Scholar]
- 4.Alberti C Prostate cancer immunotherapy, particularly in combination with androgen deprivation or radiation treatment: Customized pharmacogenomic approaches to overcome immunotherapy cancer resistance. G Chir 2017;37(5):225–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Merseburger AS, Haas GP, von Klot CA. An update on enzalutamide in the treatment of prostate cancer. Ther Adv Urol 2015;7(1):9–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Shao YH, Demissie K, Shih W, Mehta AR, Stein MN, Roberts CB, Dipaola RS, Lu-Yao GL. Contemporary risk profile of prostate cancer in the United States. J Natl Cancer Inst. 2009;101(18):1280–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Nichol AM, Warde P, Bristow RG. Optimal treatment of intermediate-risk prostate carcinoma with radiotherapy: Clinical and translational issues. Cancer. 2005;104(5):891–905. [DOI] [PubMed] [Google Scholar]
- 8.Roach M 3rd, Bae K, Speight J, Wolkov HB, Rubin P, Lee RJ, Lawton C, Valicenti R, Grignon D, Pilepich MV. Short-term neoadjuvant androgen deprivation therapy and external-beam radiotherapy for locally advanced prostate cancer: Long-term results of RTOG 8610. J Clin Oncol 2008;26(4):585–91. [DOI] [PubMed] [Google Scholar]
- 9.Delaney CP. Outcome of discharge within 24 to 72 hours after laparoscopic colorectal surgery. Dis Colon Rectum. 2008;51(2):181–5. [DOI] [PubMed] [Google Scholar]
- 10.Bolla M, Van Tienhoven G, Warde P, Dubois JB, Mirimanoff RO, Storme G, Bernier J, Kuten A, Sternberg C, Billiet I, Torecilla JL, Pfeffer R, Cutajar CL, Van der Kwast T, Collette L. External irradiation with or without long-term androgen suppression for prostate cancer with high metastatic risk: 10-year results of an EORTC randomised study. Lancet Oncol 2010;11(11):1066—73. [DOI] [PubMed] [Google Scholar]
- 11.Mettler FA. Medical effects and risks of exposure to ionising radiation. J Radiol Prot. 2012;32(1):N9–13. [DOI] [PubMed] [Google Scholar]
- 12.Weisiger RA, Fridovich I. Mitochondrial superoxide simutase. Site of synthesis and intramitochondrial localization. J Biol Chem. 1973;248(13):4793–6. [PubMed] [Google Scholar]
- 13.Morgan MA, Lawrence TS. Molecular pathways: Overcoming radiation resistance by targeting DNA damage response pathways. Clin Cancer Res. 2015;21(13): 2898–904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Ray Chaudhuri A, Nussenzweig A. The multifaceted roles of PARP1 in DNA repair and chromatin remodelling. Nat Rev Mol Cell Biol. 2017;18(10):610–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Wei H, Yu X. Functions of PARylation in DNA damage repair pathways. Genom Proteom Bioinformat. 2016;14(3):131–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Jeggo PA, Geuting V, Lobrich M. The role of homologous recombination in radiation-induced double-strand break repair. Radiother Oncol. 2011;101(1):7–12. [DOI] [PubMed] [Google Scholar]
- 17.Venkitaraman AR. Cancer susceptibility and the functions of BRCA1 and BRCA2. Cell. 2002;108(2):171–82. [DOI] [PubMed] [Google Scholar]
- 18.Alanee SR, Glogowski EA, Schrader KA, Eastham JA, Offit K. Clinical features and management of BRCA1 and BRCA2-associated prostate cancer. Front Biosci (Elite Ed). 2014;6:15–30. [DOI] [PubMed] [Google Scholar]
- 19.Lieber MR. The mechanism of double-strand DNA break repair by the nonhomologous DNA end-joining pathway. Annu Rev Biochem. 2010;79:181–211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Hall EJ. The bystander effect. Health Phys. 2003;85(1):31–5. [DOI] [PubMed] [Google Scholar]
- 21.Sharma PM, Ponnaiya B, Taveras M, Shuryak I, Turner H, Brenner DJ. High throughput measurement of gammaH2AX DSB repair kinetics in a healthy human population. PLoS One. 2015;10(3):e0121083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Barcellos-Hoff MH, Park C, Wright EG. Radiation and the microenvironment: Tumorigenesis and therapy. Nat Rev Cancer. 2005;5(11):867–75. [DOI] [PubMed] [Google Scholar]
- 23.Dayal R, Singh A, Pandey A, Mishra KP. Reactive oxygen species as mediator of tumor radiosensitivity. J Cancer Res Ther. 2014;10(4):811–8. [DOI] [PubMed] [Google Scholar]
- 24.Bienert GP, Chaumont F. Aquaporin-facilitated transmembrane diffusion of hydrogen peroxide. Biochim Biophys Acta. 2014;1840(5): 1596–604. [DOI] [PubMed] [Google Scholar]
- 25.Lennicke C, Rahn J, Lichtenfels R, Wessjohann LA, Seliger B. Hydrogen peroxide - production, fate and role in redox signaling of tumor cells. Cell Commun Signal. 2015;13:39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Ravanat JL, Breton J, Douki T, Gasparutto D, Grand A, Rachidi W, Sauvaigo S. Radiation-mediated formation of complex damage to DNA: A chemical aspect overview. Br J Radiol. 2014;87(1035):20130715. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Ogawa Y, Kobayashi T, Nishioka A, Kariya S, Hamasato S, Seguchi H, Yoshida S. Radiation-induced reactive oxygen species formation prior to oxidative DNA damage in human peripheral T cells. Int J Mol Med. 2003;11(2):149–52. [PubMed] [Google Scholar]
- 28.Szumiel I Ionizing radiation-induced oxidative stress, epigenetic changes and genomic instability: The pivotal role of mitochondria. Int J Radiat Biol. 2015;91(1): 1—12. [DOI] [PubMed] [Google Scholar]
- 29.Brand MD. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radic Biol Med. 2016;100:14–31. [DOI] [PubMed] [Google Scholar]
- 30.Yamamori T, Yasui H, Yamazumi M, Wada Y, Nakamura Y, Nakamura H, Inanami O. Ionizing radiation induces mitochondrial reactive oxygen species production accompanied by upregulation of mitochondrial electron transport chain function and mitochondrial content under control of the cell cycle checkpoint. Free Radic Biol Med. 2012;53(2):260–70. [DOI] [PubMed] [Google Scholar]
- 31.McDermott N, Meunier A, Mooney B, Nortey G, Hernandez C, Hurley S, Lynam-Lennon N, Barsoom SH, Bowman KJ, Marples B, Jones GD, Marignol L. Fractionated radiation exposure amplifies the radioresistant nature of prostate cancer cells. Sci Rep. 2016;6:34796. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Hein AL, Post CM, Sheinin YM, Lakshmanan I, Natarajan A, Enke CA, Batra SK, Ouellette MM, Yan Y. RAC1 GTPase promotes the survival of breast cancer cells in response to hyper-fractionated radiation treatment. Oncogene. 2016;35(49):6319–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Kesanakurti D, Maddirela D, Banasavadi-Siddegowda YK, Lai TH, Qamri Z, Jacob NK, Sampath D, Mohanam S, Kaur B, Puduvalli VK. A novel interaction of PAK4 with PPARgamma to regulate Nox1 and radiation-induced epithelial-to-mesenchymal transition in glioma. Oncogene. 2017;36(37):5309–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Kim KW, Moretti L, Mitchell LR, Jung DK, Lu B. Endoplasmic reticulum stress mediates radiation-induced autophagy by perk-eIF2alpha in caspase-3/7-deficient cells. Oncogene. 2010;29(22):3241–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.De Bacco F, Luraghi P, Medico E, Reato G, Girolami F, Perera T, Gabriele P, Comoglio PM, Boccaccio C. Induction of MET by ionizing radiation and its role in radioresistance and invasive growth of cancer. J Natl Cancer Inst. 2011;103(8):645–61. [DOI] [PubMed] [Google Scholar]
- 36.Zhang X, Li X, Zhang N, Yang Q, Moran MS. Low doses ionizing radiation enhances the invasiveness of breast cancer cells by inducing epithelial-mesenchymal transition. Biochem Biophys Res Commun. 2011;412(1):188–92. [DOI] [PubMed] [Google Scholar]
- 37.Montanari M, Rossetti S, Cavaliere C, D ‘Aniello C, Malzone MG, Vanacore D, Di Franco R, La Mantia E, Iovane G, Piscitelli R, Muscariello R, Berretta M, Perdona S, Muto P, Botti G, Bianchi AAM, Veneziani BM, FacchiniG. Epithelial-mesenchymal transition in prostate cancer: An overview. Oncotarget. 2017;8(21):35376–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Holley AK, Xu Y, St Clair DK, St Clair WH. RelB regulates manganese superoxide dismutase gene and resistance to ionizing radiation of prostate cancer cells. Ann N Y Acad Sci. 2010;1201:129–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Kim BY, Kim KA, Kwon O, Kim SO, Kim MS, Kim BS, Oh WK, Kim GD, Jung M, Ahn JS. NF-kappaB inhibition radiosensitizes Ki-Ras-transformed cells to ionizing radiation. Carcinogen. 2005;26(8):1395–403. [DOI] [PubMed] [Google Scholar]
- 40.Yang C, He H, Zhang T, Chen Y, Kong Z. Decreased DAB2IP gene expression, which could be induced by fractionated irradiation, is associated with resistance to gammarays and alphaparticles in prostate cancer cells. Mol Med Rep. 2016;14(1):567–73. [DOI] [PubMed] [Google Scholar]
- 41.Li Z, Pearlman AH, Hsieh P. DNA mismatch repair and the DNA damage response. DNA Repair (Amst). 2016;38:94–101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Sancar A, Lindsey-Boltz LA, Unsal-Kacmaz K, Linn S. Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annu Rev Biochem. 2004;73:39–85. [DOI] [PubMed] [Google Scholar]
- 43.Bakkenist CJ, Kastan MB. DNA damage activates ATM through intermolecular autophosphorylation and dimer dissociation. Nature. 2003;421(6922):499–506. [DOI] [PubMed] [Google Scholar]
- 44.Yoo HY, Kumagai A, Shevchenko A, Shevchenko A, Dunphy WG. The Mre11-Rad50-Nbs1 complex mediates activation of TopBP1 by ATM. Mol Biol Cell. 2009;20(9):2351–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Khemlina G, Ikeda S, Kurzrock R. Molecular landscape of prostate cancer: Implications for current clinical trials. Cancer Treat Rev. 2015;41(9):761–6. [DOI] [PubMed] [Google Scholar]
- 46.Rogakou EP, Pilch DR, Orr AH, Ivanova VS, Bonner WM. DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139. J Biol Chem. 1998; 273(10):5858–68. [DOI] [PubMed] [Google Scholar]
- 47.Ditch S, Paull TT. The ATM protein kinase and cellular redox signaling: Beyond the DNA damage response. Trends in Biochem Sci. 2012;37(1):15–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Pu H, Horbinski C, Hensley PJ, Matuszak EA, Atkinson T, Kyprianou N. PARP-1 regulates epithelial-mesenchymal transition (EMT) in prostate tumorigenesis. Carcinogen. 2014;35(11):2592–601. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Lonn P, van der Heide LP, Dahl M, Hellman U, Heldin CH, Moustakas A. PARP-1 attenuates Smad-mediated transcription. Mol Cell. 2010;40(4):521–32. [DOI] [PubMed] [Google Scholar]
- 50.Chatterjee P, Choudhary GS, Sharma A, Singh K, Heston WD, Ciezki J, Klein EA, Almasan A. PARP inhibition sensitizes to low dose-rate radiation TMPRSS2-ERG fusion gene-expressing and PTEN-deficient prostate cancer cells. PLoS One. 2013;8(4):e60408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Kyjacova L, Hubackova S, Krejcikova K, Strauss R, Hanzlikova H, Dzijak R, Imrichova T, Simova J, Reinis M, Bartek J, Hodny Z. Radiotherapy-induced plasticity of prostate cancer mobilizes stem-like non-adherent, Erk signaling-dependent cells. Cell Death Differ. 2015;22(6):898–911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Fortini P, Pascucci B, Parlanti E, DΈrrico M, Simonelli V, Dogliotti E. 8-Oxoguanine DNA damage: A t the crossroad of alternative repair pathways. Mutat Res. 2003;531(1–2):127–39. [DOI] [PubMed] [Google Scholar]
- 53.Palivec V, Pluharova E, Unger I, Winter B, Jungwirth P. DNA lesion can facilitate base ionization: Vertical ionization energies of aqueous 8-oxoguanine and its nucleoside and nucleotide. J Phys Chem B. 2014;118(48):13833–7. [DOI] [PubMed] [Google Scholar]
- 54.Noren Hooten N, Kompaniez K, Barnes J, Lohani A, Evans MK. Poly(ADP-ribose) polymerase 1 (PARP-1) binds to 8-oxoguanine-DNA glycosylase (OGG1). J Biol Chem. 2011;286(52):44679–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Kelley MR, Georgiadis MM, Fishel ML. APE1/Ref-1 role in redox signaling: Translational applications of targeting the redox function of the DNA repair/redox protein APE1/Ref-1. Curr Mol Pharmacol. 2012;5(1):36–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Pines A, Perrone L, Bivi N, Romanello M, Damante G, Gulisano M, Kelley MR, Quadrifogliof, Tell G. Activation of APE1/Ref-1 is dependent on reactive oxygen species generated after purinergic receptor stimulation by ATP. Nucleic Acids Res. 2005;33(14):4379–94. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Prasad R, Dyrkheeva N, Williams J, Wilson SH. Mammalian base excision repair: Functional partnership between PARP-1 and APE1 in AP-site repair. PLoS One. 2015;10(5):e0124269. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Yoshida T, Goto S, Kawakatsu M, Urata Y, Li TS. Mitochondrial dysfunction, a probable cause of persistent oxidative stress after exposure to ionizing radiation. Free Radic Res. 2012;46(2):147–53. [DOI] [PubMed] [Google Scholar]
- 59.Prise KM, O’Sullivan JM. Radiation-induced bystander signalling in cancer therapy. Nat Rev Cancer. 2009;9(5):351–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Miao L, Holley AK, Zhao Y, St Clair WH, St Clair DK. Redox-mediated and ionizing-radiation-induced inflammatory mediators in prostate cancer development and treatment. Antioxid Redox Signal. 2014;20(9):1481–500. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Holley AK, Miao L, St Clair DK, St Clair WH. Redox-modulated phenomena and radiation therapy: The central role of superoxide dismutases. Antioxid Redox Signal. 2014;20(10):1567–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Lu CL, Qin L, Liu HC, Candas D, Fan M, Li JJ. Tumor cells switch to mitochondrial oxidative phosphorylation under radiation via mTOR-mediated hexokinase II inhibition: A Warburg-reversing effect. PLoS One. 2015;10(3):e0121046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Yu L, Lu M, Jia D, Ma J, Ben-Jacob E, Levine H, Kaipparettu BA, Onuchic JN. Modeling the genetic regulation of cancer metabolism: Interplay between glycolysis and oxidative phosphorylation. Cancer Res. 2017;77(7):1564–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Bartoletti-Stella A, Mariani E, Kurelac I, Maresca A, Caratozzolo MF, Iommarini L, Carelli V, Eusebi LH, Guido A, Cenacchi G, Fuccio L, Rugolo M, Tullo A, Porcelli AM, Gasparre G. Gamma rays induce a p53-independent mitochondrial biogenesis that is counter-regulated by HIF1alpha. Cell Death Dis. 2013;4:e663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Wardell TM, Ferguson E, Chinnery PF, Borthwick GM, Taylor RW, Jackson G, Craft A, Lightowlers RN, Howell N, Turnbull DM. Changes in the human mitochondrial genome after treatment of malignant disease. Mutat Res. 2003;525(1–2):19–27. [DOI] [PubMed] [Google Scholar]
- 66.Nugent SM, Mothersill CE, Seymour C, McClean B, Lyng FM, Murphy JE. Increased mitochondrial mass in cells with functionally compromised mitochondria after exposure to both direct gamma radiation and bystander factors. Radiat Res. 2007;168(1):134–42. [DOI] [PubMed] [Google Scholar]
- 67.Petros JA, Baumann AK, Ruiz-Pesini E, Amin MB, Sun CQ, Hall J, Lim S, Issa MM, Flanders WD, Hosseini SH, Marshall FF, Wallace DC. mtDNA mutations increase tumorigenicity in prostate cancer. Proc Natl Acad Sci U S A. 2005;102(3):719–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Cloos CR, Daniels DH, Kalen A, Matthews K, Du J, Goswami PC, Cullen JJ. Mitochondrial DNA depletion induces radioresistance by suppressing G2 checkpoint activation in human pancreatic cancer cells. Radiat Res. 2009;171(5):581–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.van Gisbergen MW, Voets AM, Starmans MH, de Coo IF, Yadak R, Hoffmann RF, Boutros PC, Smeets HJ, Dubois L, Lambin P. How do changes in the mtDNA and mitochondrial dysfunction influence cancer and cancer therapy? Challenges, opportunities and models. Mutat Res Rev Mutat Res. 2015;764:16–30. [DOI] [PubMed] [Google Scholar]
- 70.Yamamori T, Sasagawa T, Ichii O, Hiyoshi M, Bo T, Yasui H, Kon Y, Inanami O. Analysis of the mechanism of radiation-induced upregulation of mitochondrial abundance in mouse fibroblasts. J Radiat Res. 2017;58(3):292—301. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Yu J, Wang Q, Chen N, Sun Y, Wang X, Wu L, Chen S, Yuan H, Xu A, Wang J. Mitochondrial transcription factor A regulated ionizing radiation-induced mitochondrial biogenesis in human lung adenocarcinoma A549 cells. J Radiat Res. 2013;54(6):998–1004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Yamamori T, Ike S, Bo T, Sasagawa T, Sakai Y, Suzuki M, Yamamoto K, Nagane M, Yasui H, Inanami O. Inhibition of the mitochondrial fission protein dynamin-related protein 1 (Drpl) impairs mitochondrial fission and mitotic catastrophe after x-irradiation. Mol Biol Cell. 2015;26(25):4607–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Yakes FM, Van Houten B. Mitochondrial DNA damage is more extensive and persists longer than nuclear DNA damage in human cells following oxidative stress. Proc Natl Acad Sci U S A. 1997;94(2):514–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Alexeyev M, Shokolenko I, Wilson G, LeDoux S. The maintenance of mitochondrial DNA integrity: critical analysis and update. Cold Spring Harb Perspect Biol. 2013;5(5):a012641. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Jiang WW, Rosenbaum E, Mambo E, Zahurak M, Masayesva B, Carvalho AL, Zhou S, Westra WH, Alberg AJ, Sidransky D, Koch W, Califano JA. Decreased mitochondrial DNA content in posttreatment salivary rinses from head and neck cancer patients. Clin Cancer Res. 2006;12(5):1564–9. [DOI] [PubMed] [Google Scholar]
- 76.Tang JT, Yamazaki H, Inoue T, Koizumi M, Yoshida K, Ozeki S, Inoue T. Mitochondrial DNA influences radiation sensitivity and induction of apoptosis in human fibroblasts. Anticancer Res. 1999;19(6B):4959–64. [PubMed] [Google Scholar]
- 77.Yoshida K, Yamazaki H, Ozeki S, Inoue T, Yoshioka Y, Inoue T. Role of mitochondrial DNA in radiation exposure. Radiat Med. 2000;18(2):87–91. [PubMed] [Google Scholar]
- 78.Yoshioka Y, Yamazaki H, Yoshida K, Ozeki S, Inoue T, Yoneda M, Inoue T. Impact of mitochondrial DNA on radiation sensitivity of transformed human fibroblast cells: Clonogenic survival, micronucleus formation and cellular ATP level. Radiat Res. 2004;162(2):143–7. [DOI] [PubMed] [Google Scholar]
- 79.Jain MR, Li M, Chen W, Liu T, de Toledo SM, Pandey BN, Li H, Rabin BM, Azzam EI. In vivo space radiation-induced non-targeted responses: Late effects on molecular signaling in mitochondria. Curr Mol Pharmacol. 2011;4(2):106–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Bing Z, Yang G, Zhang Y, Wang F, Ye C, Sun J, Zhou G, Yang L. Proteomic analysis of effects by x-rays and heavy ion in HeLa cells. Radiol Oncol. 2014;48(2): 142–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Chang L, Ni J, Beretov J, Wasinger VC, Hao J, Bucci J, Malouf D, Gillatt D, Graham PH, Li Y. Identification of protein biomarkers and signaling pathways associated with prostate cancer radioresistance using label-free LC-MS/MS proteomic approach. Sci Rep. 2017;7:41834. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Du S, Guan Z, Hao L, Song Y, Wang L, Gong L, Liu L, Qi X, Hou Z, Shao S. Fructose-bisphosphate aldolase a is a potential metastasis-associated marker of lung squamous cell carcinoma and promotes lung cell tumorigenesis and migration. PLoS One. 2014;9(1):e85804. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Shimura T, Noma N, Sano Y, Ochiai Y, Oikawa T, Fukumoto M, Kunugita N. AKT-mediated enhanced aerobic glycolysis causes acquired radioresistance by human tumor cells. Radiother Oncol. 2014;112(2):302–7. [DOI] [PubMed] [Google Scholar]
- 84.Meijer TW, Kaanders JH, Span PN, Bussink J. Targeting hypoxia, HIF-1, and tumor glucose metabolism to improve radiotherapy efficacy. Clin Cancer Res. 2012;18(20):5585–94. [DOI] [PubMed] [Google Scholar]
- 85.Koukourakis MI, Giatromanolaki A, Panteliadou M, Pouliliou SE, Chondrou PS, Mavropoulou S, Sivridis E. Lactate dehydrogenase 5 isoenzyme overexpression defines resistance of prostate cancer to radiotherapy. Br J Cancer. 2014;110(9):2217–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Fritz V, Benfodda Z, Henriquet C, Hure S, Cristol JP, Michel F, Carbonneau MA, Casas F, Fajas L. Metabolic intervention on lipid synthesis converging pathways abrogates prostate cancer growth. Oncogene. 2013;32(42):5101–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Fritz V, Fajas L. Metabolism and proliferation share common regulatory pathways in cancer cells. Oncogene. 2010;29(31):4369–77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Kelly RS, Sinnott JA, Rider JR, Ebot EM, Gerke T, Bowden M, Pettersson A, Loda M, Sesso HD, Kantoff PW, Martin NE, Giovannucci EL, Tyekucheva S, Heiden MV, Mucci LA. The role of tumor metabolism as a driver of prostate cancer progression and lethal disease: Results from a nested case-control study. Cancer Metab. 2016;4:22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Chaiswing L, Bourdeau-Heller JM, Zhong W, Oberley TD. Characterization of redox state of two human prostate carcinoma cell lines with different degrees of aggressiveness. Free Radic Biol Med. 2007;43(2):202–15. [DOI] [PubMed] [Google Scholar]
- 90.Chaiswing L, Zhong W, Oberley TD. Distinct redox profiles of selected human prostate carcinoma cell lines: Implications for rational design of redox therapy. Cancers. 2011;3(3):3557–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Bostwick DG, Alexander EE, Singh R, Shan A, Qian J, Santella RM, Oberley LW, Yan T, Zhong W, Jiang X, Oberley TD. Antioxidant enzyme expression and reactive oxygen species damage in prostatic intraepithelial neoplasia and cancer. Cancer. 2000;89(1):123–34. [PubMed] [Google Scholar]
- 92.Chaiswing L, Zhong W, Oberley TD. Increasing discordant antioxidant protein levels and enzymatic activities contribute to increasing redox imbalance observed during human prostate cancer progression. Free Radic Biol Med. 2014;67:342–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Reuter S, Gupta SC, Chaturvedi MM, Aggarwal BB. Oxidative stress, inflammation, and cancer: How are they linked? Free Radic Biol Med. 2010;49(11):1603–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Qu Y, Zhang H, Zhao S, Hong J, Tang C. The effect on radioresistance of manganese superoxide dismutase in nasopharyngeal carcinoma. Oncol Rep. 2010;23(4):1005–11. [DOI] [PubMed] [Google Scholar]
- 95.Riley PA. Free radicals in biology: Oxidative stress and the effects of ionizing radiation. Int J Radiat Biol. 1994;65(1):27–33. [DOI] [PubMed] [Google Scholar]
- 96.Guo G, Yan-Sanders Y, Lyn-Cook BD, Wang T, Tamae D, Ogi J, Khaletskiy A, Li Z, Weydert C, Longmate JA, Huang TT, Spitz DR, Oberley LW, Li JJ. Manganese superoxide dismutase-mediated gene expression in radiation-induced adaptive responses. Mol Cell Biol. 2003;23(7):2362–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Bensalah K, Lotan Y, Karam JA, Shariat SF. New circulating biomarkers for prostate cancer. Prostate Cancer Prostatic Dis. 2008;11(2):112–20. [DOI] [PubMed] [Google Scholar]
- 98.Shan W, Zhong W, Zhao R, Oberley TD. Thioredoxin 1 as a subcellular biomarker of redox imbalance in human prostate cancer progression. Free Radic Biol Med. 2010;49(12):2078–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Chen Q, Chai YC, Mazumder S, Jiang C, Macklis RM, Chisolm GM, Almasan A. The late increase in intracellular free radical oxygen species during apoptosis is associated with cytochrome c release, caspase activation, and mitochondrial dysfunction. Cell Death Differ. 2003;10(3):323–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Azzam EI, Jay-Gerin IP, Pain D. Ionizing radiation-induced metabolic oxidative stress and prolonged cell injury. Cancer Lett. 2012;327(1–2):48–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.de Toledo SM, Asaad N, Venkatachalam P, Li L, Howell RW, Spitz DR, Azzam EI. Adaptive responses to low-dose/low-dose-rate gamma rays in normal human fibroblasts: The role of growth architecture and oxidative metabolism. Radiat Res. 2006;166(6):849–57. [DOI] [PubMed] [Google Scholar]
- 102.Jayakumar S, Kunwar A, Sandur SK, Pandey BN, Chaubey RC. Differential response of DU145 and PC3 prostate cancer cells to ionizing radiation: Role of reactive oxygen species, GSH and Nrf2 in radiosensitivity. Biochim Biophys Acta. 2014;1840(1):485–94. [DOI] [PubMed] [Google Scholar]
- 103.Przybyszewski WM, Widel M, Szurko A, Lubecka B, Matulewicz L, Maniakowski Z, Polaniak R, Birkner E, Rzeszowska-Wolny J. Multiple bystander effect of irradiated megacolonies of melanoma cells on non-irradiated neighbours. Cancer Lett. 2004;214(1):91–102. [DOI] [PubMed] [Google Scholar]
- 104.Nugent S, Mothersill CE, Seymour C, McClean B, Lyng FM, Murphy JE. Altered mitochondrial function and genome frequency post exposure to gamma-radiation and bystander factors. Int J Radiat Biol. 2010;86(10):829–41. [DOI] [PubMed] [Google Scholar]
- 105.Battino M, Ferri E, Gattavecchia E, Breccia A, Genova ML, Littarru GP, Lenaz G. Mitochondrial respiratory chain features after gamma-irradiation. Free Radic Res. 1997;26(5):431–8. [DOI] [PubMed] [Google Scholar]
- 106.Leinonen HM, Kansanen E, Polonen P, Heinaniemi M, Levonen AL. Role of the Keap1-Nrf2 pathway in cancer. Adv Cancer Res. 2014;122:281–320. [DOI] [PubMed] [Google Scholar]
- 107.Kato K, Takahashi K, Monzen S, Yamamoto H, Maruyama A, Itoh K, Kashiwakura I. Relationship between radiosensitivity and Nrf2 target gene expression in human hematopoietic stem cells. Radiat Res. 2010;174(2): 177–84. [DOI] [PubMed] [Google Scholar]
- 108.Zhou S, Ye W, Shao Q, Zhang M, Liang J. Nrf2 is a potential therapeutic target in radioresistance in human cancer. Crit Rev Oncol Hematol. 2013;88(3):706–15. [DOI] [PubMed] [Google Scholar]
- 109.Zhang P, Singh A, Yegnasubramanian S, Esopi D, Kombairaju P, Bodas M, Wu H, Bova SG, Biswal S. Loss of Kelch-like ECH-associated protein 1 function in prostate cancer cells causes chemoresistance and radioresistance and promotes tumor growth. Mol Cancer Ther. 2010;9(2):336–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Rhee SG, Kil IS. Multiple functions and regulation of mammalian peroxiredoxins. Annu Rev Biochem. 2017; 86:749–75. [DOI] [PubMed] [Google Scholar]
- 111.Kim YJ, Ahn JY, Liang P, Ip C, Zhang Y, Park YM. Human prx1 gene is a target of Nrf2 and is up-regulated by hypoxia/reoxygenation: Implication to tumor biology. Cancer Res. 2007;67(2): 546–54. [DOI] [PubMed] [Google Scholar]
- 112.Kim JH, Bogner PN, Baek SH, Ramnath N, Liang P, Kim HR, Andrews C, Park YM. Up-regulation of peroxiredoxin 1 in lung cancer and its implication as a prognostic and therapeutic target. Clin Cancer Res. 2008;14(8):2326–33. [DOI] [PubMed] [Google Scholar]
- 113.Riddell JR, Bshara W, Moser MT, Spernyak JA, Foster BA, Gollnick SO. Peroxiredoxin 1 controls prostate cancer growth through Toll-like receptor 4-dependent regulation of tumor vasculature. Cancer Res. 2011;71(5):1637–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Guo Q, Huang X, Zhang J, Luo Y, Peng Z, Li S. Downregulation of peroxiredoxin I by a novel fully human phage display recombinant antibody induces apoptosis and enhances radiation sensitization in A549 lung carcinoma cells. Cancer Biother Radiopharm. 2012;27(5):307–16. [DOI] [PubMed] [Google Scholar]
- 115.Zhang B, Wang Y, Liu K, Yang X, Song M, Wang Y, Bai Y. Adenovirus-mediated transfer of siRNA against peroxiredoxin I enhances the radiosensitivity of human intestinal cancer. Biochem Pharmacol. 2008;75(3):660–7. [DOI] [PubMed] [Google Scholar]
- 116.Orlowski RZ, Baldwin AS Jr., NF-kappaB as a therapeutic target in cancer. Trends Mol Med. 2002;8(8):385–9. [DOI] [PubMed] [Google Scholar]
- 117.Jung M, Zhang Y, Lee S, Dritschilo A. Correction of radiation sensitivity in ataxia telangiectasia cells by a truncated I kappa B-alpha. Science. 1995;268(5217):1619–21. [DOI] [PubMed] [Google Scholar]
- 118.Bergmann M, Hart L, Lindsay M, Barnes PJ, Newton R. IkappaBalpha degradation and nuclear factor-kappaB DNA binding are insufficient for interleukin-1 beta and tumor necrosis factor-alpha-induced kappaB-dependent transcription: Requirement for an additional activation pathway. J Biol Chem. 1998;273(12):6607–10. [DOI] [PubMed] [Google Scholar]
- 119.Ahmed KM, Li JJ. ATM-NF-kappaB connection as a target for tumor radiosensitization. Curr Cancer Drug Targets. 2007;7(4):335–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Flynn V, Ramanitharan A, Moparty K, Davis R, Sikka S, Agrawal KC, Abdel-Mageed AB. Adenovirus-mediated inhibition of NF-kappaB confers chemo-sensitization and apoptosis in prostate cancer cells. Int J Oncol. 2003;23(2):317–23. [PubMed] [Google Scholar]
- 121.Pajonk F, Pajonk K, McBride WH. Inhibition of NF-kappaB, clonogenicity, and radiosensitivity of human cancer cells. J Natl Cancer Inst. 1999;91(22):1956–60. [DOI] [PubMed] [Google Scholar]
- 122.Bai M, Ma X, Li X, Wang X, Mei Q, Li X, Wu Z, Han W. The accomplices of NF-kappaB lead to radioresistance. Curr Protein Pept Sci. 2015;16(4):279–94. [DOI] [PubMed] [Google Scholar]
- 123.Kumar S, Singh RK, Meena R. Emerging targets for radioprotection and radiosensitization in radiotherapy. Tumour Biol. 2016;37(9):11589–609. [DOI] [PubMed] [Google Scholar]
- 124.Zou X, Ratti BA, O’Brien JG, Lautenschlager SO, Gius DR, Bonini MG, Zhu Y. Manganese superoxide dismutase (SOD2): Is there a center in the universe of mitochondrial redox signaling? J Bioenerg Biomembr. 2017;49(4):325–33. [DOI] [PubMed] [Google Scholar]
- 125.Hardmeier R, Hoeger H, Fang-Kircher S, Khoschsorur A, Lubec G. Transcription and activity of antioxidant enzymes after ionizing irradiation in radiation-resistant and radiation-sensitive mice. Proc Natl Acad Sci U S A. 1997;94(14):7572–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Kashiwai T, Ichihara K, Tamaki H, Endo Y, Kimura M, Takeoka K, Amino N, Miyai K. The stability of immunological and biological activity of human thyrotropin in buffer: Its temperature-dependent dissociation into subunits during freezing. Scand J Clin Lab Invest. 1991;51(5):417–23. [DOI] [PubMed] [Google Scholar]
- 127.Hosoki A, Yonekura S, Zhao QL, Wei ZL, Takasaki I, Tabuchi Y, Wang LL, Hasuike S, Nomura T, Tachibana A, Hashiguchi K, Yonei S, Kondo T, Zhang-Akiyama QM. Mitochondria-targeted superoxide dismutase (SOD2) regulates radiation resistance and radiation stress response in HeLa cells. J Radiat Res. 2012;53(1):58–71. [DOI] [PubMed] [Google Scholar]
- 128.Xu Y, Fang F, Miriyala S, Crooks PA, Oberley TD, Chaiswing L, Noel T, Holley AK, Zhao Y, Kiningham KK, Clair DK, Clair WH. KEAP1 is a redox sensitive target that arbitrates the opposing radiosensitive effects of parthenolide in normal and cancer cells. Cancer Res. 2013;73(14):4406–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Xu Y, Josson S, Fang F, Oberley TD, St Clair DK, Wan XS, Sun Y, Bakthavatchalu V, Muthuswamy A, St Clair WH. RelB enhances prostate cancer growth: Implications for the role of the nuclear factor-kappaB alternative pathway in tumorigenicity. Cancer Res. 2009;69(8):3267–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Josson S, Xu Y, Fang F, Dhar SK, St Clair DK, St Clair WH. RelB regulates manganese superoxide dismutase gene and resistance to ionizing radiation of prostate cancer cells. Oncogene. 2006;25(10):1554–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Wei X, Xu Y, Xu FF, Chaiswing L, Schnell D, Noel T, Wang C, Chen J, St Clair DK, St Clair WH. RelB expression determines the differential effects of ascorbic acid in normal and cancer cells. Cancer Res. 2017;77(6):1345–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Lee SY, Jeong EK, Ju MK, Jeon HM, Kim MY, Kim CH, Park HG, Han SI, Kang HS. Induction of metastasis, cancer stem cell phenotype, and oncogenic metabolism in cancer cells by ionizing radiation. Mol Cancer. 2017;16(1):10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Deng X, Elzey BD, Poulson JM, Morrison WB, Ko SC, Hahn NM, Ratliff TL, Hu CD. Ionizing radiation induces neuroendocrine differentiation of prostate cancer cells in vitro, in vivo and in prostate cancer patients. Am J Cancer Res. 2011;1(7):834–44. [PMC free article] [PubMed] [Google Scholar]
- 134.Theys J, Jutten B, Habets R, Paesmans K, Groot A J, Lambin P, Wouters BG, Lammering G, Vooijs M. E-Cadherin loss associated with EMT promotes radioresistance in human tumor cells. Radiother Oncol. 2011;99(3):392–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Stark TW, Hensley PJ, Spear A, Pu H, Strup SS, Kyprianou N. Predictive value of epithelial-mesenchymal-transition (EMT) signature and PARP-1 in prostate cancer radioresistance. Prostate. 2017;77(16):1583–91. [DOI] [PubMed] [Google Scholar]
- 136.Nantajit D, Lin D, Li JJ. The network of epithelial-mesenchymal transition: Potential new targets for tumor resistance. J Cancer Res Clin Oncol. 2015;141(10):1697–713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Radisky DC, Levy DD, Littlepage LE, Liu H, Nelso Fata JE, Leake D, Godden EL, Albertson DG, Nieto MA, Werb Z, Bissell MJ. Rac1b and reactive oxygen species mediate MMP-3-induced EMT and genomic instability. Nature. 2005;436(7047):123–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Rhyu DY, Yang Y, Ha H, Lee GT, Song JS, Uh ST, Lee HB. Role of reactive oxygen species in TGF-beta1-induced mitogen-activated protein kinase activation and epithelial-mesenchymal transition in renal tubular epithelial cells. J Am Soc Nephrol. 2005;16(3):667–75. [DOI] [PubMed] [Google Scholar]
- 139.Mahabir R, Tanino M, Elmansuri A, Wang L, Kimura T, Itoh T, Ohba Y, Nishihara H, Shirato H, Tsuda M, Tanaka S. Sustained elevation of Snail promotes glial-mesenchymal transition after irradiation in malignant glioma. Neuro Oncol. 2014;16(5):671–85. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Lee SY, Jeon HM, Ju MK, Jeong EK, Kim CH, Yoo MA, Park HG, Han SI, Kang HS. Dlx-2 is implicated in TGF-beta- and Wnt-induced epithelial-mesenchymal, glycolytic switch, and mitochondrial repression by Snail activation. Int J Oncol. 2015;46(4):1768–80. [DOI] [PubMed] [Google Scholar]
- 141.Lee SY, Jeon HM, Ju MK, Kim CH, Yoon G, Han SI, Park HG, Kang HS. Wnt/Snail signaling regulates cytochrome C oxidase and glucose metabolism. Cancer Res. 2012;72(14):3607–17. [DOI] [PubMed] [Google Scholar]
- 142.Bastos LG, de Marcondes PG, de-Freitas-Junior JC, Leve F, Mencalha AL, de Souza WF, de Araujo WM, Tanaka MN, Abdelhay ES, Morgado-Diaz JA. Progeny from irradiated colorectal cancer cells acquire an EMT-like phenotype and activate Wnt/beta-catenin pathway. J Cell Biochem. 2014;115(12):2175–87. [DOI] [PubMed] [Google Scholar]
- 143.Kim RK, Kaushik N, Suh Y, Yoo KC, Cui YH, Kim MJ, Lee HJ, Kim IG, Lee SJ. Radiation driven epithelial-mesenchymal transition is mediated by Notch signaling in breast cancer. Oncotarget. 2016;7(33):53430–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Kang J, Kim E, Kim W, Seong KM, Youn H, Kim JW, Kim J, Youn B. Rhamnetin and cirsiliol induce radiosensitization and inhibition of epithelial-mesenchymal transition (EMT) by miR-34a-mediated suppression of Notch-1 expression in non-small cell lung cancer cell lines. J Biol Chem. 2013;288(38):27343–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Dent P, Yacoub A, Fisher PB, Hagan MP, Grant S. MAPK pathways in radiation responses. Oncogene. 2003;22(37):5885–96. [DOI] [PubMed] [Google Scholar]
- 146.Wu CT, Chang YH, Lin WY, Chen WC, Chen MF. TGF beta1 expression correlates with survival and tumor aggressiveness of prostate cancer. Ann Surg Oncol. 2015;22(Suppl 3):S1587–93. [DOI] [PubMed] [Google Scholar]
- 147.de Brot S, Ntekim A, Cardenas R, James V, Allegrucci C, Heery DM, Bates DO, Odum N, Persson JL, Mongan NP. Regulation of vascular endothelial growth factor in prostate cancer. Endocr Relat Cancer. 2015;22(3):R107–23. [DOI] [PubMed] [Google Scholar]
- 148.Chiba N, Comaills V, Shiotani B, Takahashi F, Shimada T, Tajima K, Winokur D, Hayashida T, Willers H, Brachtel E, Vivanco MD, Haber DA, Zou L, Maheswaran S. Homeobox B9 induces epithelial-to-mesenchymal transition-associated radioresistance by accelerating DNA damage re CM, sponses. Proc Natl Acad Sci U S A. 2012;109(8):2760–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Yoon YS, Lee JH, Hwang SC, Choi KS, Yoon G. TGF beta1 induces prolonged mitochondrial ROS generation through decreased complex IV activity with senescent arrest in Mv1Lu cells. Oncogene. 2005;24(11):1895–903. [DOI] [PubMed] [Google Scholar]
- 150.Quiros-Gonzalez I, Sainz RM, Hevia D, Mayo JC. MnSOD drives neuroendocrine differentiation, androgen independence, and cell survival in prostate cancer cells. Free Radic Biol Med. 2011;50(4):525–36. [DOI] [PubMed] [Google Scholar]
- 151.Cheng H, Lee SH, Wu S. Effects of N-acetyl-L-cysteine on adhesive strength between breast cancer cell and extracellular matrix proteins after ionizing radiation. Life Sci. 2013;93(21):798–803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Tobar N, Villar V, Santibanez JF. ROS-NFkappaB mediates TGF-beta1-induced expression of urokinase-type plasminogen activator, matrix metalloproteinase-9 and cell invasion. Mol Cell Biochem. 2010;340(1–2):195–202. [DOI] [PubMed] [Google Scholar]
- 153.Bentzen SM, Overgaard J. Patient-to-patient variability in the expression of radiation-induced normal tissue injury. Semin Radiat Oncol. 1994;4(2):68–80. [DOI] [PubMed] [Google Scholar]
- 154.Kozin SV, Niemierko A, Huang P, Silva J, Doppke KP, Suit HD. Inter- and intramouse heterogeneity of radiation response for a growing paired organ. Radiat Res. 2008;170(2):264–7. [DOI] [PubMed] [Google Scholar]
- 155.Krause M, Gurtner K, Deuse Y, Baumann M. Heterogeneity of tumour response to combined radiotherapy and EGFR inhibitors: Differences between antibodies and TK inhibitors. Int J Radiat Biol. 2009;85(11):943–54. [DOI] [PubMed] [Google Scholar]
- 156.Blaylock RL. Cancer microenvironment, inflammation and cancer stem cells: A hypothesis for a paradigm change and new targets in cancer control. Surg Neurol Int. 2015;6:92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Murray L, Henry A, Hoskin P, Siebert FA, Venselaar J, ESTRO PgoG. Second primary cancers after radiation for prostate cancer: A systematic review of the clinical data and impact of treatment technique. Radiother Oncol. 2014;110(2):213–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Wang YJ, Huang CY, Hou WH, Wang CC, Lan KH, Yu HJ, Lai MK, Liu SP, Pu YS, Cheng JC. Dual-timing PSA as a biomarker for patients with salvage intensity modulated radiation therapy for biochemical failure after radical prostatectomy. Oncotarget. 2016;7(28):44224–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Lacombe J, Azria D, Mange A, Solassol J. Proteomic approaches to identify biomarkers predictive of radiotherapy outcomes. Expert Rev Proteom. 2013;10(1):33–42. [DOI] [PubMed] [Google Scholar]
- 160.Amundson SA, Do KT, Vinikoor LC, Lee RA, Koch-Paiz CA, Ahn J, Reimers M, Chen Y, Scudiero DA, Weinstein JN, Trent JM, Bittner ML, Meltzer PS, Fornace AJ Jr. Integrating global gene expression and radiation survival parameters across the 60 cell lines of the National Cancer Institute Anticancer Drug Screen. Cancer Res. 2008;68(2):415–24. [DOI] [PubMed] [Google Scholar]
- 161.Takahashi Y, Rayman JB, Dynlacht BD. Analysis of promoter binding by the E2F and pRB families in vivo: Distinct E2F proteins mediate activation and repression. Genes Dev. 2000;14(7):804–16. [PMC free article] [PubMed] [Google Scholar]
- 162.DuPree EL, Mazumder S, Almasan A. Genotoxic stress induces expression of E2F4, leading to its association with p130 in prostate carcinoma cells. Cancer Res. 2004;64(13):4390–3. [DOI] [PubMed] [Google Scholar]
- 163.Verona R, Moberg K, Estes S, Starz M, Vernon JP, Lees JA. E2F activity is regulated by cell cycle-dependent changes in subcellular localization. Mol Cell Biol. 1997;17(12):7268–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Crosby ME, Jacobberger J, Gupta D, Macklis RM, Almasan A. E2F4 regulates a stable G2 arrest response to genotoxic stress in prostate carcinoma. Oncogene. 2007;26(13):1897–909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Shen MM, Abate-Shen C. Molecular genetics of prostate cancer: New prospects for old challenges. Genes Dev. 2010;24(18): 1967–2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Locke JA, Zafarana G, Malloff CA, Lam WL, Sykes J, Pintilie M, Ramnarine VR, Meng A, Ahmed O, Jurisica I, Guns ET, van der Kwast T, Milosevic M, Bristow RG. Allelic loss of the loci containing the androgen synthesis gene, StAR, is prognostic for relapse in intermediate-risk prostate cancer. Prostate. 2012;72(12):1295–305. [DOI] [PubMed] [Google Scholar]
- 167.Wu CL, Schroeder BE, Ma XJ, Cutie CJ, Wu S, Salunga R, Zhang Y, Kattan MW, Schnabel CA, Erlander MG, McDougal WS. Development and validation of a 32-gene prognostic index for prostate cancer progression. Proc Natl Acad Sci U S A. 2013;110(15):6121–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Erho N, Crisan A, Vergara IA, Mitra AP, Ghadessi M, Buerki C, Bergstralh EJ, Kollmeyer T, Fink S, Haddad Z, Zimmermann B, Sierocinski T, Ballman KV, Triche TJ, Black PC, Karnes RJ, Klee G, Davicioni E, Jenkins RB. Discovery and validation of a prostate cancer genomic classifier that predicts early metastasis following radical prostatectomy. PLoS One. 2013;8(6):e66855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Dal Pra A, Lalonde E, Sykes J, Warde F, Ishkanian A, Meng A, Maloff C, Srigley J, Joshua AM, Petrovics G, van der Kwast T, Evans A, Milosevic M, Saad F, Collins C, Squire J, Lam W, Bismar TA, Boutros PC, Bristow RG. TMPRSS2-ERG status is not prognostic following prostate cancer radiotherapy: Implications for fusion status and DSB repair. Clin Cancer Res. 2013;19(18):5202–9. [DOI] [PubMed] [Google Scholar]
- 170.Berg KD. The prognostic and predictive value of TMPRSS2-ERG gene fusion and ERG protein expression in prostate cancer biopsies. Dan Med J. 2016;63(12):B5319. [PubMed] [Google Scholar]
- 171.Zhao SG, Chang SL, Spratt DE, Erho N, Yu M, Ashab HA, Alshalalfa M, Speers C, Tomlins SA, Davicioni E, Dicker AP, Carroll PR, Cooperberg MR, Freedland SJ, Karnes RJ, Ross AE, Schaeffer EM, Den RB, Nguyen PL, Feng FY. Development and validation of a 24-gene predictor of response to postoperative radiotherapy in prostate cancer: A matched, retrospective analysis. Lancet Oncol. 2016;17(11):1612–20. [DOI] [PubMed] [Google Scholar]
- 172.Thoma C Prostate cancer: Postoperative RT response prediction. Nat Rev Urol. 2016;13(12):694. [DOI] [PubMed] [Google Scholar]
- 173.Il’yasova D, Mixon G, Wang F, Marcom PK, Marks J, Spasojevich I, Craft N, Arredondof, DiGiulio R. Markers of oxidative status in a clinical model of oxidative assault: A pilot study in human blood following doxorubicin administration. Biomarkers. 2009;14(5):321–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Frazier ME, Reese JA, Morris JE, Jostes RF, Miller DL. Exposure of mammalian cells to 60-Hz magnetic or electrie fields: Analysis ofDNArepair ofinduced, single-strand breaks. Bioelectromag. 1990;11(3):229–34. [DOI] [PubMed] [Google Scholar]
- 175.Ouyang X, DeWeese TL, Nelson WG, Abate-Shen C. Loss-of-function of Nkx3.1 promotes increased oxidative damage in prostate carcinogenesis. Cancer Res. 2005;65(15):6773–9. [DOI] [PubMed] [Google Scholar]
- 176.Kitagishi Y, Matsuda S. Redox regulation of tumor suppressor PTEN in cancer and aging (Review). Int J Mol Med. 2013;31 (3):511—5. [DOI] [PubMed] [Google Scholar]
- 177.Sonveaux P ROS and radiotherapy: More we care. Oncotarget. 2017;8(22):35482–3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Dal Pra A, Locke JA, Borst G, Supiot S, Bristow RG. Mechanistic insights into molecular targeting and combined modality therapy for aggressive, localized prostate cancer. Front Oncol. 2016;6:24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Brown JM. Tumor hypoxia, drug resistance, and metastases. J Natl Cancer Inst. 1990;82(5):338–9. [DOI] [PubMed] [Google Scholar]
- 180.Paul-Gilloteaux P, Potiron V, Delpon G, Supiot S, Chiavassa S, Paris F, Costes SV. Optimizing radiotherapy protocols using computer automata to model tumour cell death as a function of oxygen diffusion processes. Sci Rep. 2017;7(1):2280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Elahimanesh F, Shabestani Monfared A, Khosravifarsani M, Akhavan Niaki H, Abedian Z, Hajian-Tilaki K, Borzouisileh S, Seyfizadeh N, Amiri M. Is radiosensitivity associated to different types of blood groups? (A cytogenetic study). Int J Mol Cell Med. 2013;2(3):131–5. [PMC free article] [PubMed] [Google Scholar]
- 182.Challapalli A, Carroll L, Aboagye EO. Molecular mechanisms of hypoxia in cancer. Clin Transl Imaging. 2017;5(3):225–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Yaromina A, Thames H, Zhou X, Hering S, Eicheler W, Dorfler A, Leichtner T, Zips D, Baumann M. Radiobiological hypoxia, histological parameters of tumour microenvironment and local tumour control after fractionated irradiation. Radiother Oncol. 2010;96(1):116–22. [DOI] [PubMed] [Google Scholar]
- 184.Bristow RG, Hill RP Hypoxia and metabolism. Hypoxia, DNA repair and genetic instability. Nat Rev Cancer. 2008;8(3):180–92. [DOI] [PubMed] [Google Scholar]
- 185.Vergis R, Corbishley CM, Norman AR, Bartlett J, Jhavar S, Borre M, Heeboll S, Horwich A, Huddart R, Khoo V, Eeles R, Cooper C, Sydes M, Dearnaley D, Parker C. Intrinsic markers of tumour hypoxia and angiogenesis in localised prostate cancer and outcome of radical treatment: A retrospective analysis of two randomised radiotherapy trials and one surgical cohort study. Lancet Oncol. 2008;9(4):342–51. [DOI] [PubMed] [Google Scholar]
- 186.Deep G, Panigrahi GK. Hypoxia-induced signaling promotes prostate cancer progression: Exosomes role as messenger of hypoxic response in tumor microenvironment. Crit Rev Oncog. 2015;20(5–6):419–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Ramteke A, Ting H, Agarwal C, Mateen S, Somasagara R, Hussain A, Graner M, Frederick B, Agarwal R, Deep G. Exosomes secreted under hypoxia enhance invasiveness and stemness of prostate cancer cells by targeting adherens junction molecules. Mol Carcinog. 2015;54(7):554–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Lalonde E, Ishkanian AS, Sykes J, Fraser M, Ross-Adams H, Erho N, Dunning MJ, Halim S, Lamb AD, Moon NC, Zafarana G, Warren AY, Meng X, Thoms J, Grzadkowski MR, Berlin A, Have CL, Ramnarine VR, Yao CQ, Malloff CA, Lam LL, Xie H, Harding NJ, Mak DY, Chu KC, Chong LC, Sendorek DH, P’ng C, Collins CC, Squire JA, Jurisica I, Cooper C, Eeles R, Pintilie M, Dal Pra A, Davicioni E, Lam WL, Milosevic M, Neal DE, van der Kwast T, Boutros PC, Bristow RG. Tumour genomic and microenvironmental heterogeneity for integrated prediction of 5-year biochemical recurrence of prostate cancer: A retrospective cohort study. Lancet Oncol. 2014;15(13):1521–32. [DOI] [PubMed] [Google Scholar]
- 189.Milosevic M, Warde P, Menard C, Chung P Toi A, Ishkanian A, McLean M, Pintilie M, Sykes J, Gospodarowicz M, Catton C, Hill RP, Bristow R. Tumor hypoxia predicts biochemical failure following radiotherapy for clinically localized prostate cancer. Clin Cancer Res. 2012;18(7):2108–14. [DOI] [PubMed] [Google Scholar]
- 190.Zips D, Boke S, Kroeber T, Meinzer A, Bruchner K, Thames HD, Baumann M, Yaromina A. Prognostic value of radiobiological hypoxia during fractionated irradiation for local tumor control. Strahlenther Onkol. 2011;187(5):306–10. [DOI] [PubMed] [Google Scholar]
- 191.Lin A, Maity A. Molecular pathways: A novel approach to targeting hypoxia and improving radiotherapy efficacy via reduction in oxygen demand. Clin Cancer Res. 2015;21(9):1995–2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Goel S, Duda DG, Xu L, Munn LL, Boucher Y, Fukumura D, Jain RK. Normalization of the vasculature for treatment of cancer and other diseases. Physiol Rev. 2011;91(3):1071–121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Yip K, Alonzi R. Carbogen gas and radiotherapy outcomes in prostate cancer. Ther Adv Urol. 2013;5(1):25–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Chughtai B, Lee R, Te A, Kaplan S. Role of inflammation in benign prostatic hyperplasia. Rev Urol. 2011; 13(3):147–50. [PMC free article] [PubMed] [Google Scholar]
- 195.Holmes WE, Lee J, Kuang WJ, Rice GC, Wood WI. Structure and functional expression of a human interleukin-8 receptor. Science. 1991;253(5025):1278–80. [DOI] [PubMed] [Google Scholar]
- 196.Caruso DJ, Carmack AJ, Lokeshwar VB, Duncan RC, Soloway MS, Lokeshwar BL. Osteopontin and interleukin-8 expression is independently associated with prostate cancer recurrence. Clin Cancer Res. 2008;14(13): 4111–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Seaton A, Scullin P, Maxwell PJ, Wilson C, Pettigrew J, Gallagher R, O’Sullivan JM, Johnston PG, Waugh DJ. Interleukin-8 signaling promotes androgen-independent proliferation of prostate cancer cells via induction of androgen receptor expression and activation. Carcinogen. 2008;29(6): 1148–56. [DOI] [PubMed] [Google Scholar]
- 198.Medici D, Hay ED, Olsen BR. Snail and Slug promote epithelial-mesenchymal transition through beta-catenin-T-cell factor-4-dependent expression of transforming growth factor-beta3. Mol Biol Cell. 2008;19(11):4875–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Liu J, Uygur B, Zhang Z, Shao L, Romero D, Vary C, Ding Q, Wu WS. Slug inhibits proliferation of human prostate cancer cells via downregulation of cyclin D1 expression. Prostate. 2010;70(16):1768–77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Zhang P, Wei Y, Wang L, Debeb BG, Yuan Y, Zhang J, Yuan J, Wang M, Chen D, Sun Y, Woodward WA, Liu Y, Dean DC, Liang H, Hu Y, Ang KK, Hung MC, Chen J, Ma L. ATM-mediated stabilization of ZEB1 promotes DNA damage response and radioresistance through CHK1. Nat Cell Biol. 2014;16(9):864–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Zhan T, Rindtorff N, Boutros M. Wnt signaling in cancer. Oncogene. 2017;36(11):1461–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Wong SC, Lo ES, Chan AK, Lee KC, Hsiao WL. Nuclear beta catenin as a potential prognostic and diagnostic marker in patients with colorectal cancer from Hong Kong. Mol Pathol. 2003;56(6):347–52. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Kel’tsev VA, Amirov RZ, Ugnich KA. Integral electrocar-diotopography in children and its diagnostic value [Article in Russian], Pediatriia. 1989;4:46–50. [PubMed] [Google Scholar]
- 204.Pu H, Collazo J, Jones E, Gayheart D, Sakamoto S, Vogt A, Mitchell B, Kyprianou N. Dysfunctional transforming growth factor-beta receptor II accelerates prostate tumorigenesis in the TRAMP mouse model. Cancer Res. 2009;69(18):7366–74. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Drabsch Y, ten Dijke P. TGF-beta signalling and its role in cancer progression and metastasis. Cancer Meta Rev. 2012;31(3–4):553–68. [DOI] [PubMed] [Google Scholar]
- 206.Tang B, Vu M, Booker T, Santner SJ, Miller FR, Anver MR, Wakefield LM. TGF-beta switches from tumor suppressor to prometastatic factor in a model of breast cancer progression. J Clin Invest. 2003;112(7):1116–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Dancea HC, Shareef MM, Ahmed MM. Role of radiation-induced TGF-beta signaling in cancer therapy. Mol Cell Pharmacol. 2009;1(1):44–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Barcellos-Hoff MH, Dix TA. Redox-mediated activation of latent transforming growth factor-beta 1. Mol Endocrinol. 1996;10(9): 1077–83. [DOI] [PubMed] [Google Scholar]
- 209.Overstreet JM, Samarakoon R, Cardona-Grau D, Goldschmeding R, Higgins PJ. Tumor suppressor ataxia telangiectasia mutated functions downstream of TGF-betal in orchestrating profibrotic responses. FASEB J. 2015;29(4): 1258–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Michaeloudes C, Sukkar MB, Khorasani NM, Bhavsar PK, Chung KF. TGF-beta regulates Nox4, MnSOD and catalase expression, and IL-6 release in airway smooth muscle cells. Am J Physiol Lung Cell Mol Physiol. 2011;300(2):L295–304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Shariat SF, Semjonow A, Lilja H, Savage C, Vickers AJ, Bjartell A. Tumor markers in prostate cancer I: Blood-based markers. Acta Oncol. 2011;50(Suppl 1):61–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Batra JS, Girdhani S, Hlatky L. A Quest to identify prostate cancer circulating biomarkers with a bench-to-bedside potential. J Biomark. 2014;2014:321680. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Reis ST, Pontes-Junior J, Antunes AA, Sousa-Canavez JM, Abe DK, Cruz JA, Dall’oglio MF, Crippa A, Passerotti CC, Ribeiro-Filho LA, Viana NI, Srougi M, Leite KR. TGF-beta1 expression as a biomarker of poor prognosis in prostate cancer. Clinics (Sao Paulo). 2011;66(7):1143–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Khan MI, Hamid A, Adhami VM, Lall RK, Mukhtar H. Role of epithelial mesenchymal transition in prostate tumorigenesis. Curr Pharm Des. 2015;21(10):1240–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 215.Grant CM, Kyprianou N. Epithelial mesenchymal transition (EMT) in prostate growth and tumor progression. Transl Androl Urol. 2013;2(3):202–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Rycaj K, Li H, Zhou J, Chen X, Tang DG. Cellular determinants and microenvironmental regulation of prostate cancer metastasis. Semin Cancer Biol. 2017;44:83–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Shiota M, Itsumi M, Takeuchi A, Imada K, Yokomizo A, Kuruma H, Inokuchi J, Tatsugami K, Uchiumi T, Oda Y, Naito S. Crosstalk between epithelial-mesenchymal transition and castration resistance mediated by Twist1/AR signaling in prostate cancer. Endocr Relat Cancer. 2015;22(6):889–900. [DOI] [PubMed] [Google Scholar]
- 218.Munroe M, Kolesar J. Olaparib for the treatment of BRCA-mutated advanced ovarian cancer. Am J Health Syst Pharm. 2016;73(14):1037–41. [DOI] [PubMed] [Google Scholar]
- 219.Whitesell L, Lindquist SL. HSP90 and the chaperoning of cancer. Nat Rev Cancer. 2005;5(10):761–72. [DOI] [PubMed] [Google Scholar]
- 220.Camphausen K, Tofllon PJ. Inhibition of Hsp90: A multitarget approach to radiosensitization. Clin Cancer Res. 2007;13(15 Part 1): 4326–30. [DOI] [PubMed] [Google Scholar]
- 221.Beck R, Dejeans N, Glorieux C, Pedrosa RC, Vasquez D, Valderrama JA, Calderon PB, Verrax J. Molecular chaperone Hsp90 as a target for oxidant-based anticancer therapies. Curr Med Chem. 2011;18(18):2816–25. [DOI] [PubMed] [Google Scholar]
- 222.Alarcon SV, Mollapour M, Lee MJ, Tsutsumi S, Lee S, Kim YS, Prince T, Apolo AB, Giaccone G, Xu W, Neckers LM, Trepel JB. Tumor-intrinsic and tumor-extrinsic factors impacting hsp90- targeted therapy. Curr Mol Med. 2012;12(9):1125–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Gao J, Aksoy BA, Dogrusoz U, Dresdner G, Gross B, Sumer SO, Sun Y, Jacobsen A, Sinha R, Larsson E, Cerami E, Sander C, Schultz N. Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Sci Signal. 2013;6(269): pl1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Skvortsova I, Skvortsov S, Stasyk T, Raju U, Popper BA, Schiestl B, von Guggenberg E, Neher A, Bonn GK, Huber LA, Lukas P. Intracellular signaling pathways regulating radioresistance of human prostate carcinoma cells. Proteom. 2008;8(21):4521–33. [DOI] [PubMed] [Google Scholar]
- 225.Quanz M, Herbette A, Sayarath M, de Koning L, Dubois T, Sun JS, Dutreix M. Heat shock protein 90alpha (Hsp90alpha) is phosphorylated in response to DNA damage and accumulates in repair foci. J Biol Chem. 2012;287(12):8803–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Palacios DA, Miyake M, Rosser CJ. Radiosensitization in prostate cancer: Mechanisms and targets. BMC Urol. 2013;13:4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Roach M 3rd, Waldman F, Pollack A Predictive models in external beam radiotherapy for clinically localized prostate cancer. Cancer. 2009;115(13 Suppl): 3112–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 228.Gavande NS, VanderVere-Carozza PS, Hinshaw HD, Jalal SI, Sears CR, Pawelczak KS, Turchi JJ. DNA repair targeted therapy: The past or future of cancer treatment? Pharmacol Ther. 2016;160:65–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.Polkinghorn WR, Parker JS, Lee MX, Kass EM, Spratt DE, Iaquinta PJ, Arora VK, Yen WF, Cai L, Zheng D, Carver BS, Chen Y, Watson PA, Shah NP, Fujisawa S, Goglia AG, Gopalan A, Hieronymus H, Wongvipat J, Scardino PT, Zelefsky MJ, Jasin M, Chaudhuri J, Powell SN, Sawyers CL. Androgen receptor signaling regulates DNA repair in prostate cancers. Cancer Discov. 2013;3(11):1245–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Vance S, Liu E, Zhao L, Parsels JD, Parsels LA, Brown JL, Maybaum J, Lawrence TS, Morgan MA. Selective radiosensitization of p53 mutant pancreatic cancer cells by combined inhibition of Chk1 and PARP1. Cell Cycle. 2011;10(24):4321–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231.Dungey FA, Caldecott KW, Chalmers AJ. Enhanced radiosensitization of human glioma cells by combining inhibition of poly(ADP-ribose) polymerase with inhibition of heat shock protein 90. Mol Cancer Ther. 2009;8(8):2243–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Karnak D, Engelke CG, Parsels LA, Kausar T, Wei D, Robertson JR, Marsh KB, Davis MA, Zhao L, Maybaum J, Lawrence TS, Morgan MA. Combined inhibition of Wee1 and PARP1/2 for radiosensitization in pancreatic cancer. Clin Cancer Res. 2014;20(19):5085–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Morgan MA, Parsels LA, Maybaum J, Lawrence TS. Improving the efficacy of chemoradiation with targeted agents. Cancer Discov. 2014;4(3):280–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 234.Beltran H, Yelensky R, Frampton GM, Park K, Downing SR, MacDonald TY, Jarosz M, Lipson D, Tagawa ST, Nanus DM, Stephens PJ, Mosquera JM, Cronin MT, Rubin MA. Targeted next-generation sequencing of advanced prostate cancer identifies potential therapeutic targets and disease heterogeneity. Eur Urol. 2013;63(5):920–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Robinson D, Van Allen EM, Wu YM, Schultz N, Lonigro RJ, Mosquera JM, Montgomery B, Taplin ME, Pritchard CC, Attard G, Beltran H, Abida W, Bradley RK, Vinson J, Cao X, Vats P, Kunju LP, Hussain M, Feng FY, Tomlins SA, Cooney KA, Smith DC, Brennan C, Siddiqui J, Mehra R, Chen Y, Rathkopf DE, Morris MJ, Solomon SB, Durack JC, Reuter VE, Gopalan A, Gao J, Loda M, Lis RT, Bowden M, Balk SP, Gaviola G, Sougnez C, Gupta M, Yu EY, Mostaghel EA, Cheng HH, Mulcahy H, True LD, Plymate SR, Dvinge H, Ferraldeschi R, Flohr P, Miranda S, Zafeiriou Z, Tunariu N, Mateo J, Perez-Lopez R, Demichelis F, Robinson BD, Schiffman M, Nanus DM, Tagawa ST, Sigaras A, Eng KW, Elemento O, Sboner A, Heath EI, Scher HI, Pienta KJ, Kantoff P, de Bono JS, Rubin MA, Nelson PS, Garraway LA, Sawyers CL, Chinnaiyan AM. Integrative clinical genomics of advanced prostate cancer. Cell. 2015;161(5):1215–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Benafif S, Hall M. An update on PARP inhibitors for the treatment of cancer. Onco Targets Ther. 2015;8:519–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Fong PC, Boss DS, Yap TA, Tutt A, Wu P, MerguiRoelvink M, Mortimer P, Swaisland H, Lau A, O’Connor MJ, Ashworth A, Carmichael J, Kaye SB, Schellens JH, de Bono JS. Inhibition of poly(ADP-ribose) polymerase in tumors from BRCA mutation carriers. N Engl J Med. 2009;361(2): 123–34. [DOI] [PubMed] [Google Scholar]
- 238.Senftleben U, Cao Y, Xiao G, Greten FR, Krahn G, Bonizzi G, Chen Y, Hu Y, Fong A, Sun SC, Karin M. Activation by IKKalpha of a second, evolutionary conserved, NF-kappa B signaling pathway. Science. 2001;293(5534):1495–9. [DOI] [PubMed] [Google Scholar]
- 239.Cancer Genome Atlas Research Network. The molecular taxonomy of primary prostate cancer. Cell. 2015;163(4): 1011–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Dedes KJ, Wetterskog D, Mendes-Pereira AM, Natrajan R, Lambros MB, Geyer FC, Vatcheva R, Savage K, Mackay A, Lord CJ, Ashworth A, Reis-Filho JS. PTEN deficiency in endometrioid endometrial adenocarcinomas predicts sensitivity to PARP inhibitors. Sci Transl Med. 2010;2(53): 53ra75. [DOI] [PubMed] [Google Scholar]
- 241.Mendes-Pereira AM, Martin SA, Brough R, McCarthy A, Taylor JR, Kim JS, Waldman T, Lord CJ, Ashworth A. Synthetic lethal targeting of PTEN mutant cells with PARP inhibitors. EMBO Mol Med. 2009;1(6–7):315–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 242.Schiewer MJ, Goodwin JF, Han S, Brenner JC, Augello MA, Dean JL, Liu F, Planck JL, Ravindranathan P, Chinnaiyan AM, McCue P, Gomella LG, Raj GV, Dicker AP, Brody JR, Pascal JM, Centenera MM, Butler LM, Tilley WD, Feng FY, Knudsen KE. Dual roles of PARP-1 promote cancer growth and progression. Cancer Discov. 2012;2(12):1134–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 243.Hussain M, Daignault-Newton S, Twardowski PW, Albany C, Stein MN, Kunju LP, Siddiqui J, Wu YM, Robinson D, Lonigro RJ, Cao X, Tomlins SA, Mehra R, Cooney KA, Montgomery B, Antonarakis ES, Shevrin DH, Corn PG, Whang YE, Smith DC, Caram MV, Knudsen KE, Stadler WM, Feng FY, Chinnaiyan AM. Targeting androgen receptor and DNA repair in metastatic castration-resistant prostate cancer: Results from NCI 9012. J Clin Oncol. 2018;36(10):991–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244.Candas D, Li JJ. MnSOD in oxidative stress response-potential regulation via mitochondrial protein influx. Antioxid Redox Signal. 2014;20(10):1599–617. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245.Epperly MW, Bray JA, Krager S, Berry LM, Gooding W, Engelhardt JF, Zwacka R, Travis EL, Greenberger JS. Intratracheal injection of adenovirus containing the human MnSOD transgene protects athymic nude mice from irradiation-induced organizing alveolitis. Int J Radiat Oncol Biol Phys. 1999;43(1):169–81. [DOI] [PubMed] [Google Scholar]
- 246.Everett WH, Curiel DT. Gene therapy for Cancer Gene Ther. 2015;22(4):172–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Borrelli A, Schiattarella A, Mancini R, Morrica B, Cerciello V, Mormile M, d’Alesio V, Bottalico L, Morelli F, D ‘Armiento M, D ‘Armiento FP, Mancini A A recombinant MnSOD is radioprotective for normal cells and radiosensitizing for tumor cells. Free Radic Biol Med. 2009;46(1):110–6. [DOI] [PubMed] [Google Scholar]
- 248.Das KC, Lewis-Molock Y, White CW. Thiol modulation of TNF alpha and IL-1 induced MnSOD gene expression and activation of NF-kappa B. Mol Cell Biochem. 1995;148(1):45–57. [DOI] [PubMed] [Google Scholar]
- 249.Hussain SP, Amstad P, He P, Robles A, Lupold S, Kaneko I, Ichimiya M, Sengupta S, Mechanic L, Okamura S, Hofseth LJ, Moake M, Nagashima M, Forrester KS, Harris CC. p53-induced up-regulation of MnSOD and GPx but not catalase increases oxidative stress and apoptosis. Cancer Res. 2004;64(7):2350–6. [DOI] [PubMed] [Google Scholar]
- 250.Wei X, Xu Y, Xu FF, Chaiswing L, Schnell D, Noel T, Wang C, Chen J, St Clair DK, St Clair WH. RelB expression determines the differential effects of ascorbic acid in normal and cancer cells. Cancer Res. 2017;77(6):1345–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Xu Y, Fang F, Sun Y, St Clair DK, St Clair WH. RelB-dependent differential radiosensitization effect of STI571 on prostate cancer cells. Mol Cancer Ther. 2010;9(4): 803–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Xu Y, Fang F, St Clair DK, Josson S, Sompol P, Spasojevic I, St Clair WH. Suppression of RelB-mediated manganese superoxide dismutase expression reveals a primary mechanism for radiosensitization effect of 1alpha,25-dihydroxyvitamin D(3) in prostate cancer cells. Mol Cancer Ther. 2007;6(7):2048–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Batinic-Haberle I, Tovmasyan A, Roberts ER, Vujaskovic Z, Leong KW, Spasojevic I. SOD therapeutics: Latest insights into their structure-activity relationships and impact on the cellular redox-based signaling pathways. Antioxid Redox Signal. 2014;20(15):2372–415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Moeller BJ, Batinic-Haberle I, Spasojevic I, Rabbani ZN, Anscher MS, Vujaskovic Z, Dewhirst MW. A manganese porphyrin superoxide dismutase mimetic enhances tumor radioresponsiveness. Int J Radiat Oncol Biol Phys. 2005;63(2): 545–52. [DOI] [PubMed] [Google Scholar]
- 255.Weitner T, Kos I, Sheng H, Tovmasyan A, Reboucas JS, Fan P, Warner DS, Vujaskovic Z, Batinic-Haberle I, Spasojevic I. Comprehensive pharmacokinetic studies and oral bioavailability of two Mn porphyrin-based SOD mimics, MnTE-2-PyP5+ and MnTnHex-2-PyP5+. Free Radic Biol Med. 2013;58:73–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Shin SW, Choi C, Lee GH, Son A, Kim SH, Park HC, Batinic-Haberle I, Park W. Mechanism of the antitumor and radiosensitizing effects of a manganese porphyrin, MnHex-2-PyP. Antioxid Redox Signal. 2017;27(14):1067–82. [DOI] [PubMed] [Google Scholar]
- 257.Zhao Y, Carroll DW, You Y, Chaiswing L, Wen R, Batinic Haberle I, Bondada S, Liang Y, St Clair DK. A novel redox regulator, MnTnBuOE-2-PyP5+, enhances normal hematopoietic stem/progenitor cell function. Redox Biol. 2017;12:129–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Oberley-Deegan RE, Steffan JJ, Rove KO, Pate KM, Weaver MW, Spasojevic I, Frederick B, Raben D, Meacham RB, Crapo JD, Koul HK. The antioxidant, MnTE-2-PyP, prevents side-effects incurred by prostate cancer irradiation. PLoS One. 2012;7(9):e44178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 259.Guo L, Zhang Y, Zhang L, Huang F, Li J, Wang S. MicroRNAs, TGF-beta signaling, and the inflammatory microenvironment in cancer. Tumour Biol. 2016;37(1): 115–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 260.Bouquet F, Pal A, Pilones KA, Demaria S, Hann B, Akhurst RJ, Babb JS, Lonning SM, DeWyngaert JK, Formenti SC, Barcellos-Hoff MH. TGFbeta1 inhibition increases the radiosensitivity of breast cancer cells in vitro and promotes tumor control by radiation in vivo. Clin Cancer Res. 2011;17(21):6754–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Zhang M, Lahn M, Huber PE. Translating the combination of TGFbeta blockade and radiotherapy into clinical development in glioblastoma. Oncoimmunol. 2012;1(6):943–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 262.Biswas S, Guix M, Rinehart C, Dugger TC, Chytil A, Moses HL, Freeman ML, Arteaga CL. Inhibition of TGF-beta with neutralizing antibodies prevents radiation-induced acceleration of metastatic cancer progression. J Clin Invest. 2007;117(5): 1305–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Muraoka-Cook RS, Kurokawa H, Koh Y, Forbes JT, Roebuck LR, Barcellos-Hoff MH, Moody SE, Chodosh LA, Arteaga CL. Conditional overexpression of active transforming growth factor beta1 in vivo accelerates metastases of transgenic mammary tumors. Cancer Res. 2004; 64(24):9002–11. [DOI] [PubMed] [Google Scholar]
- 264.de Gramont A, Faivre S, Raymond E. Novel TGF-beta inhibitors ready for prime time in onco-immunology. Oncoimmunol. 2017;6(1):e1257453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 265.Schirmer MA, Brockmoller J, Rave-Frank M, Virsik P, Wilken B, Kuhnle E, Campean R, Hoffmann AO, Muller K, Goetze RG, Neumann M, Janke JH, Nasser F, Wolff HA, Ghadimi BM, Schmidberger H, Hess CF, Christiansen H, Hille A. A putatively functional haplotype in the gene encoding transforming growth factor beta-1 as a potential biomarker for radiosensitivity. Int J Radiat Oncol Biol Phys. 2011;79(3):866–74. [DOI] [PubMed] [Google Scholar]
- 266.Hinz S, Pagerols-Raluy L, Oberg HH, Ammerpohl O, Grussel S, Sipos B, Grutzmann R, Pilarsky C, Ungefroren H, Saeger HD, Kloppel G, Kabelitz D, Kalthoff H. Foxp3 expression in pancreatic carcinoma cells as a novel mechanism of immune evasion in cancer. Cancer Res. 2007;67(17):8344–50. [DOI] [PubMed] [Google Scholar]
- 267.Zhao L, Ji W, Zhang L, Ou G, Feng Q, Zhou Z, Lei M, Yang W, Wang L. Changes of circulating transforming growth factor-beta1 level during radiation therapy are correlated with the prognosis of locally advanced non-small cell lung cancer. J Thorac Oncol. 2010;5(4):521–5. [DOI] [PubMed] [Google Scholar]
- 268.Hall S, Rudrawar S, Zunk M, Bernaitis N, Arora D, McDermott CM, Anoopkumar-Dukie S. Protection against radiotherapy-induced toxicity. Antiox (Basel). 2016;5(3):E22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269.Herbertz S, Sawyer JS, Stauber AJ, Gueorguieva I, Driscoll KE, Estrem ST, Cleverly AL, Desaiah D, Guba SC, Benhadji KA, Slapak CA, Lahn MM. Clinical development of galunisertib (LY2157299 monohydrate), a small molecule inhibitor of transforming growth factor-beta signaling pathway. Drug Des Dev Ther. 2015;9:4479–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Kabakov AE, Kudryavtsev VA, Gabai VL. Hsp90 inhibitors as promising agents for radiotherapy. J Mol Med (Berl). 2010;88(3):241–7. [DOI] [PubMed] [Google Scholar]
- 271.Stingl L, Stuhmer T, Chatterjee M, Jensen MR, Flentje M, Djuzenova CS. Novel HSP90 inhibitors, NVP-AUY922 and NVP-BEP800, radiosensitise tumour cells through cell-cycle impairment, increased DNA damage and repair protraction. Br J Cancer. 2010;102(11):1578–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 272.Garcia-Morales P, Carrasco-Garcia E, Ruiz-Rico P, Martinez-Mira R, Menendez-Gutierrez MP, Ferragut JA, Saceda M, Martinez-Lacaci I. Inhibition of Hsp90 function by ansamycins causes downregulation of cdc2 and cdc25c and G(2)/M arrest in glioblastoma cell lines. Oncogene. 2007;26(51):7185–93. [DOI] [PubMed] [Google Scholar]
- 273.Proia DA, Bates RC. Ganetespib and HSP90: Translating preclinical hypotheses into clinical promise. Cancer Res. 2014;74(5):1294–300. [DOI] [PubMed] [Google Scholar]
- 274.Walton-Diaz A, Khan S, Bourboulia D, Trepel JB, Neckers L, Mollapour M. Contributions of co-chaperones and post-translational modifications towards Hsp90 drug sensitivity. Future Med Chem. 2013;5(9):1059–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 275.Dote H, Cema D, Burgan WE, Camphausen K, Tofilon PJ. ErbB3 expression predicts tumor cell radiosensitization induced by Hsp90 inhibition. Cancer Res. 2005;65(15):6967–75. [DOI] [PubMed] [Google Scholar]
- 276.Kang BH, Altieri DC. Compartmentalized cancer drug discovery targeting mitochondrial Hsp90 chaperones. Oncogene. 2009;28(42):3681–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277.Chae YC, Angelin A, Lisanti S, Kossenkov AV, Speicher KD, Wang H, Powers JF, Tischler AS, Pacak K, Fliedner S, Michalek RD, Karoly ED, Wallace DC, Languino LR, Speicher DW, Altieri DC. Landscape of the mitochondrial Hsp90 metabolome in tumours. Nat Commun. 2013;4:2139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278.Kang BH, Siegelin MD, Plescia J, Raskett CM, Garlick DS, Dohi T, Lian JB, Stein GS, Languino LR, Altieri DC.Preclinical characterization of mitochondria-targeted small molecule hsp90 inhibitors, gamitrinibs, in advanced prostate cancer. Clin Cancer Res. 2010;16(19):4779–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279.Kang BH, Tavecchio M, Goel HL, Hsieh CC, Garlick DS, Raskett CM, Lian JB, Stein GS, Languino LR, Altieri DC. Targeted inhibition of mitochondrial Hsp90 suppresses localised and metastatic prostate cancer growth in a genetic mouse model of disease. Br J Cancer. 2011;104(4):629–34. [DOI] [PMC free article] [PubMed] [Google Scholar]





