• No results found

University of Groningen The use of organoids in the study of radiation response and therapeutic window Nagle, Peter

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen The use of organoids in the study of radiation response and therapeutic window Nagle, Peter"

Copied!
29
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

The use of organoids in the study of radiation response and therapeutic window

Nagle, Peter

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2018

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Nagle, P. (2018). The use of organoids in the study of radiation response and therapeutic window.

University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

General Introduction

CHAP

(3)
(4)

1

Cancer is a major health problem, with 14.1 million new cases and 8.2 million deaths due to cancer worldwide in 2012 alone (1). The number of new cases per year is estimated to rise to 22.2 million by the year 2030 (2). Therefore, it is necessary to continually advance and develop cancer treatments, to better cope with increasing numbers of patients and to attain a better level of patient care post-treatment. The three oldest and most common cancer treatment modalities are surgery (3), chemotherapy (4) and radiotherapy (5), with more modern treatments, such as immunotherapy (6-8), being developed.

Radiotherapy is used (either alone or in combination with other forms of treatment) in approximately 50% of all cancer patients (9, 10). The objective of radiotherapy as a cancer treatment is to administer the greatest dose possible to the tumor while limiting the dose to the surrounding normal tissue (5). This balance between the benefit of radiotherapy and the risk to normal tissue is known as the therapeutic window. However, there are many factors which limit the ability to optimally deliver radiation doses. Nowadays, developments are continuously taking place to further enhance the effects of radiotherapy. As techniques are developed to optimize the efficacy of radiotherapy, it is also necessary to develop new experimental models to better understand the mechanisms that determine the limitations of radiotherapy and how new developments impact these mechanisms.

Limitations of radiotherapy

The therapeutic outcome of radiotherapy is influenced by many factors. These factors include variables such as availability of oxygen, stage of tumor differentiation, presence of cancer stem cells, rates of cellular proliferation and initiation/capacity of DNA damage responses. While all these factors potentially contribute to the effects of radiotherapy on tumor cells, they may also play a role in the response of normal tissue to unavoidable co-irradiation and consequently limit the dose which can be delivered to the tumor.

Mechanisms of radioresistance

To enhance tumor elimination while minimizing the adverse side effects caused to normal tissue, removal of cancer stem cells (CSCs) is necessary, as these are the cells which are resistant to many anti-cancer therapies, including radiation therapy. Moreover, CSCs are capable of regenerating a tumor following treatment. Features of the tumor microenvironment, such as hypoxia (10), and intracellular signal transduction pathways (11) are all factors that may contribute to tumor radioresistance.

(5)

12

1

Furthermore, CSCs themselves are often found to be intrinsically more radiorestistant than non-stem cells of the tumor, often due the same factors listed above (11-13).

Regions of varying oxygen levels are often found within a tumor, and cells within hypoxic regions are frequently more resistant to anti-cancer therapy than cells in regions with higher oxygen levels. One mechanism that is suggested to contribute to the increased radioresistance in hypoxic regions compared to normoxic conditions is the ‘oxygen fixation hypothesis’, first proposed by Alexander

and Charlesby in 1954 (14). Under normoxic conditions DNA lesions generated by ionizing radiation

are ‘fixed’ by oxygen to form more ‘unrepairable’ DNA damage, which induce cell death. However, under hypoxic conditions DNA radicals are not fixed by oxygen (or to a lesser degree), leading to less permanent DNA damage, and thus the induction levels of cell death are reduced (10).

Furthermore, hypoxia stabilizes the transcription factor hypoxia inducible factor 1α (HIF-1α). Upon stabilization, HIF-1α translocates to the nucleus where it forms a dimer with HIF-1β. HIF-1 binds to DNA and induces its downstream targets which promote cell survival (15), and thus further counteracts the toxic effects of radiation treatment. Therefore, the two primary causes of radioresistance in hypoxic regions are due to the influence of hypoxia on DNA damage and cell survival mechanisms. Furthermore, hypoxia and HIF-1α may also contribute towards the inherent resistance of CSCs (16, 17). Hypoxic regulation of signalling downstream of mammalian target of rapamycin (mTOR) contributes to autophagy and thus induces stem-like properties, including treatment resistance (18). An increase in autophagy due to HIF-1α activation has been shown to induce non-CSCs to transform CSCs (or at least CSC-like cells) (19, 20).

Changes in many intracellular signalling pathways have also been shown to promote radioresistance in tumor cells. For example, insulin-like growth factor 1 (IGF-1) overexpression is associated with radioresistance and tumor recurrence in breast cancer (21), while epidermal growth factor (EGF) and its receptor (EGFR) have to been shown to be associated with radioresistance in prostate carcinoma cell lines (22), head and neck squamous cell carcinomas (23, 24) and glioma cells (25, 26), amongst others. Another cytokine, transforming growth factor beta (TGF-β), has also been shown to enhance radioresistance of esophageal cell lines (27). Furthermore, mutations in the p53 tumor suppressor pathway have long been known to induce radioresistance (28). These pathways play various physiological roles in inflammation, promoting cell survival, cell division, to promoting angiogenesis,

(6)

1

and thus may lead to resistance or repopulation after treatment. Targeting these pathways may offer a promising means to enhance the radiosensitivity of a tumor.

Further understanding of the mechanisms behind radioresistance may reveal novel cues for enhancing tumor response to (chemo-)radiotherapy. New models for studying radioresistance, especially CSC radioresistance, may allow for the identification of patients who may respond to radiotherapy prior to undergoing treatment.

Normal tissue side effects

Normal tissue side effects can be divided into early and late side effects based on when they occur and the severity of the consequences. Early (or acute) side effects occur during, immediately after or soon after (within weeks) radiotherapy treatment (29). These side effects occur due to DNA damage induced cell death of highly proliferative cells with a fast turnover, resulting in inflammation. Early side effects may be (partly) reversible and include pneumonitis (inflammation of the alveolar walls in the lungs) (30, 31), erythema (32), memory loss, fatigue (33), diarrhoea (29) and oral mucositis (34). Late normal tissue side effects occur several months, to even years, after radiation treatment (29). Late side effects are generally chronic (and often become worse over time) and are therefore often dose limiting for radiotherapy (29). It has been shown by numerous in vivo animal studies and clinical studies that normal tissue side effects show a dose-to-volume relationship. For example, following irradiation for head and neck cancers, xerostomia is the result of late and acute side effects of the submandibular and parotid salivary glands in particular and is well-known to impact greatly on the quality of life of patients (35-37).

Another well-known late side effect with a dose-to-volume relationship is ischemic heart disease (38, 39). Darby et al. showed the risk of ischemic heart disease increases proportionally with the mean dose to the heart during radiation treatment for breast cancer (40). Other chronic side effects that may impact on the quality of life following radiation treatment include radiation induced fibrosis (41). Radiation induced fibrosis can occur in the heart (42), lung (30, 43), liver (44), or any tissue which lies in the field of therapy.

(7)

14

1

side effects, the damage is frequently irreversible and, as stated, has dramatic effects on the quality of life of patients following treatment. Normal tissue stem cells are essential for the homeostasis of tissue which is affected by radiation therapy (45). Therefore, further insights into the mechanisms and consequences of normal tissue side effects, particularly with a focus on normal tissue stem cells, will allow for better patient care in the future. Although in vivo studies offer the most complete insights into the mechanisms behind normal tissue damage, the development of further in vitro models for studying radiotherapy will provide an invaluable source of evidence into mechanisms of normal tissue side effects that may not be easily discovered in vivo.

Low dose hyper-radiosensitivity and increased radioresistance

Optimal treatment planning for conventional radiotherapy (using X-rays) treatments targets the maximum irradiation dose (usually 1.8-2 Gy/fraction up to doses of 72 Gy or even more) to the tumor. This results in potentially large volumes of normal tissue surrounding the tumor receiving low doses, often below 1 Gy/fraction (as depicted in figure 1), which can potentially elicit a low dose response.

(8)

1

Figure 1: Standard IMRT treatment plan for head and neck tumor. The maximal dose (green) is delivered to the tumor, however, a large portion of normal healthy tissue surrounding the tumor still receives a lower dose. Image courtesy of Department of Radiation Oncology, University Medical Center Groningen.

Mathematical models, such as the linear quadratic model, for the prediction of cell survival following irradiation generally assume that survival (and thereby functional) responses will decrease continuously with increasing radiation doses (46-48). However, both in vitro and in vivo, there are many indications that responses deviate from these predictions (49). Deviation from the predicted response was first described by Marples and Joiner in Chinese hamster V79 cells in which they saw a steeper decrease in survival at low doses (below 0.6 Gy) than at fractionally higher doses (>0.6 Gy) (50). This phenomenon is frequently known as low dose hyper-radiosensitivity (LDHRS) and increased radioresistance (IRR) (49). Wouters et al. studied a panel of tumor cell lines and observed similar responses in the majority of these, showing that this was not a cell type specific phenomenon (51). Furthermore, Wouters et

al. suggested that a dose-dependent difference in the handling of DNA damage induced by photons

resulted in the subsequent resistance at higher doses (51).

Although the exact mechanism behind the increased survival at slightly higher doses is not known, many studies point towards the role of an induction of DNA damage responses at higher doses. Short

et al. first identified a role of the cell cycle phase in affecting the presence/absence of LDHRS in T98G

glioblastoma cells (52) and later identified this to be most prominent in G2 (53), with Krueger et al. suggesting this was due to evasion of the G2 checkpoint dependent of ATM (54). Despite a lack of a full understanding of the mechanism for LDHRS/IRR, this phenomenon may still be of significant importance for patient treatment planning.

Attempts have been made to harness the effects of LDHRS/IRR to enhance the efficacy of radiation of tumors, particularly tumors which are known to be radiorestistant, such as glioblastoma. Beauchesne

et al. performed a phase II trial of ultra-fractionation (3 daily fractions of 0.75 Gy each, 5 days a week

for 6 weeks) to treat unresectable glioblastoma (55). This study showed an increased benefit and prolonged survival compared to radiotherapy fractions of 2 Gy per day for 6 weeks (55). While this study shows the potential importance of LDHRS/IRR in clinical tumor treatment planning, very little has been done to study the role of LDHRS/IRR in normal tissue.

(9)

16

1

If present, LDHRS/IRR in normal tissue may have a large impact on normal tissue side effects, as normal tissue surrounding a tumor may receive multiple low doses (often below 0.5 Gy). Hamilton et

al. investigated erythema (a common early side effect of the skin) following doses of less than 2 Gy

(32). Below doses of 1.5 Gy, the linear quadratic model for predicting radiation response, significantly deviated from the predicted erythema response (32), indicating the presence of LDHRS/IRR in normal tissue. These results show that LDHRS/IRR may indeed be present in normal tissue, however there is still much left to understand of LDHRS/IRR in normal tissue.

Furthermore, the impact of LDHRS/IRR has not been investigated in normal tissue stem cells. Normal tissue stem cells are the most important cells for the regeneration and function of tissues co-irradiated during radiotherapy treatment (45, 56, 57). In the future, the role of LDHRS/IRR in normal tissue stem cells should be evaluated, to enable the most effective normal tissue sparing. Appropriate in vitro stem cell models may allow for this.

Stem cell models in radiotherapy

As mentioned above, stem cells are the most important cells required for regeneration of tissues damaged by radiation treatment (depicted in figure 2). Due to this integral role of stem cells, stem cell therapy in patients to alleviate normal tissue side effects is not a new concept (58). Transplantation of bone marrow cells to replace the hematopoietic system following total body irradiation has already been used as part of the treatment of leukaemia since it was first described in mice by Vos et al. in 1956 (59) and subsequently in humans by Thomas et al. in 1957 (60). However, stem cell transplantation other than bone marrow for normal tissues affected by radiation is still a relatively new field, but there are highly promising preclinical data to show the strength and potential of stem cell therapy.

(10)

1

Figure 2: Tissue homeostasis. In normal healthy tissues, there is a balance between cell proliferation driven by stem cells and cell death to keep tissues under homeostatic conditions. Following damage, normally an increase in cell death is balanced by an increase in stem cell driven proliferation. However, if this damage is induced by irradiation stem cells may be depleted, shifting the balance towards cell death, leading to irradiation-induced side effects and reduced tissue functionality.

(11)

18

1

Recent in vivo mouse studies have shown that transplantation of human embryonic stem cells and human neural stem cells can alleviate radiation-induced cognitive impairment (61, 62), while the potential of mesenchymal stem cells to protect against radiation-induced late side effects of the lung in mouse models has also been revealed (63, 64). Furthermore, our group has shown the capacity of cultured salivary gland organoid cells, derived from both mouse and human, to restore the function of radiation-damaged salivary glands in mice (56, 57, 65, 66). These studies hold significant promise for the treatment of radiation-induced side effects.

While regenerative medicine is a field with great potential for the amelioration of late radiotherapy side effects, in vitro stem cell cultures have also opened new and exciting avenues in the fields of drug discovery and disease development (67-69). Organoids (in vitro 3D cell culture models derived from organ specific stem cells, which resemble mini-organs of their origin) offer the previously unprecedented opportunity to study organ specific drug interactions and furthermore hold great potential in the field of personalized medicine (depicted in figure 3) (70, 71).

Figure 3: Organoids: a potential model to study radiation-induced side effects. Radiation-induced side effects observed at an organ and tissue scale are induced by damage in single cells, at a DNA level. Organoid models have been developed for many of the critical organs/tissues that are affected by radiation-induced side effects, and potentially offer a more physiologically relevant model to study radiation-induced side effects than adherent cell lines.

(12)

1

Since identifying Lgr5 as a gastro-intestinal stem cell marker (72), Clevers and colleagues have developed many stem cell ‘mini-gut’ models, including intestinal (73) and colon (74). These models enabled Dekkers et al. to very elegantly show restoration of a functional CFTR gene (a cystic fibrosis transmembrane conductance regulator) in rectal tissue-derived organoids from cystic fibrosis patients (75). Treatment with Forskolin resulted in rapid swelling in healthy donor derived organoids, whereas swelling was reduced in organoids from patients with a mutation in CFTR (75). Furthermore, using this organoid model it was possible to restore a functional CFTR gene by chemical, temperature-based (75) and gene-editing methods (76), and to show differential drug responses between patient-derived organoids (77). These studies show the potential of using organoids models for diagnosis, gene-cell therapy, and drug discovery.

Many other groups are also successfully using organoid models for drug and toxicity studies in both healthy and cancer derived models. Using kidney organoids derived from induced pluripotent stem cells, Takasato et al. showed the known nephrotoxicity of cisplatin (78) and thus showed the potential using organoid models to identify nephrotoxicants in the future. Using patient-derived organoids for pancreatic adenocarcinoma, Huang et al. showed patient-specific sensitivities to an inhibitor of EZH2 (an epigenetic regulator), which furthermore showed the possibilities of using organoid models for drug screening (79).

The results of these studies are of great significance for the field of drug discovery, but also open the possibility of using organoid cultures to study the effects of radiotherapy treatment on normal (or cancer-derived) tissue. Furthermore, 3D culturing models are closer to in vivo tissue response than 2D cell cultures in reaction to exogenous stimuli, such as drugs and irradiation (80, 81). For example, in response to irradiation 2D cell culture models have been shown to be far more sensitive than 3D models or in vivo models (80). Therefore, organoid studies could be of huge importance for studying mechanistic and tissue specific responses to radiation in the future, but may also offer the possibility of personalized treatment planning based on patient derived organoid cultures.

Developments to optimize radiotherapy

Current radiobiology studies primarily focus on means to optimize the efficacy of radiotherapy, by increasing the therapeutic window. However, there is no single parameter which can achieve this in the most successful way. Therefore, radiobiological studies are currently looking at many

(13)

20

1

different parameters that can be changed to achieve an enhanced therapeutic window. Physically, the development of high precision radiation techniques has enabled the delivery of more accurate beam delivery, and therefore a reduction of the dose to the normal tissue surrounding the tissue. Furthermore, a better understanding of both normal tissue responses and signalling pathways altered in cancer (stem cells) has enabled the possibilities of altering the biological responses to radiation, allowing for the specific targeting of tumor cells to yield a more radiosensitive phenotype and/or even protection of normal tissue.

High precision techniques to enhance radiotherapy

High precision techniques, such as intensity modulated radiotherapy (IMRT) and particle therapy, can be used to more accurately deliver radiation to the desired tumor target and spare more normal tissue from irradiation. As discussed above, dose to normal tissue results in side effects (both early and late), and thus the use of IMRT and particle therapy is increasing.

In recent decades, the use of IMRT to treat head and neck cancers has become increasingly more popular (82-85). This is due to the location of critical organs and structures in the head and neck region, such as the brainstem and spinal cord, optic nerves (86) or submandibular and parotid salivary glands (37, 87, 88) which can be spared better using IMRT than conventional radiotherapy practices. However, although IMRT is intended to reduce radiation-induced side effects by sparing healthy tissue, not all side effects are eliminated. Despite sparing the parotid gland, xerostomia is still a persistent problem even in patients who undergo IMRT (82, 89). Furthermore, somewhat counter-intuitively, IMRT can result in a smaller effect in tissue sparing than might be expected due to a lower dose tolerance of parts of organs adjacent to an area receiving a high dose (88). When critical regions of organs are within the irradiation field a relative increased detrimental effect has been shown at lower doses. This is known as the bath-and-shower effect, and preclinical data indicates that it may affect many important organs, including cervical spinal cord and salivary glands (88, 90-92). Therefore, other high precision techniques such as particle therapy (protons or carbon ions) may be of even greater benefit than IMRT, as exit dose is eliminated and bath doses can be prevented.

Since charged particles were first used in 1954 until the end of 2014 more than 135,000 patients have been treated worldwide, with a rapid increase in both the number of facilities and the number of treated patients per year over the years (93). The potential benefit of using charged particles based on

(14)

1

the Bragg peak in therapeutic radiation treatment was first suggested in 1946 by Robert Wilson (94). Physically, charged particles can be modulated to encompass the whole tumor in a ‘spread-out Bragg peak’ and this is highly advantageous because entrance dose is minimized and exit dose is eliminated, and thus more efficiently spare healthy tissue (95, 96).

Despite the physical advantages of particle therapy, preclinical studies focusing on particle irradiations have so far been limited and thus much of what we know comes from clinical findings. Charged particles are currently being used in ongoing trials for the treatment of liver cancer, breast cancer, brain cancer and head and neck cancers, among others (97). Initial data suggests that proton therapy is extremely promising for the treatment of liver cancer (98) and head and neck cancers (99) with less adverse side effects than IMRT. However, trials are still ongoing, and further preclinical research and clinical trials are required to investigate the long-term side effects of particle therapy.

Preclinical studies of particle radiation generally investigate the differential induction and expression of various pathways important for the biological responses to radiation therapy, such as DNA damage induction and repair (100) or induction of apoptosis (101). Particle irradiation induces a more complex form of DNA damage (known as clustered damage) than X-ray irradiations. Clustered damage consists of two or more lesions in close proximity to each other on a DNA helix, and is usually repaired slower or remains unrepaired, potentially resulting in increased cell death (100). Furthermore, in terms of apoptosis induction, our group has shown that particles with a high linear energy transfer (LET) induce a greater level of p53 serine-37 phosphorylation, a phosphorylation event is involved in the induction of cellular apoptosis (101). Increased cellular death due to unrepaired lesions and induction of apoptosis, combined with the ability to better target the tumor with particles, all support the idea that particle irradiation is of greater strength in tumor control and limiting unwanted side effects than photon irradiation.

Charged particle irradiation is generally compared to high-energy photon irradiation based on its relative biological effectiveness (RBE). The RBE is a ratio of the doses of differing radiation modalities to reach a defined biological endpoint. For proton irradiation, a RBE of 1.1 is generally accepted for comparison with photons (102). This value was calculated based on the analysis of multiple in vivo and in vitro studies (103, 104). However, while RBE is dependent on the differing endpoints that can be used to compare radiation modalities, it is also dependent on many other parameters including the

(15)

22

1

LET, position in the spread-out Bragg peak and even the tissue/cell lines being studied (102, 105-108), and these parameters were not considered in the calculation of the general RBE for protons. While the RBE is a useful concept for explaining different radiation modalities, it is perhaps not the best way to compare radiations types, as a single RBE can never be determined.

Drugs and other agents to enhance radiotherapy

To enhance the beneficial effects of treatment and minimize the detrimental side effects of radiotherapy, radiotherapy treatment is frequently combined with surgery, immunotherapy or chemotherapy. Combining radiotherapy with chemotherapy is not a new approach, and many attempts have been made to take advantage of the multitude of preclinical data acquired in order to enhance radiosensitivity of tumors (via DNA repair pathways or other signal transduction pathways) (109).

Targeting the DNA damage response (DDR) is one of the most common methods of enhancing tumor radiosensitivity (110). Within the DDR, one of the best understood and most common approaches to enhance radiosensitivity involves inhibition of poly (ADP-ribose) polymerase (PARP) (111). If a cell is unable to efficiently cope with DSBs, these lesions are lethal. PARP1 is essential for base excision repair and single strand break repair (112). Following inhibition of PARP, unrepaired radiation-induced single strand breaks will form double strand breaks (DSB) during replication and thus increase the load on DSB repair pathways. BRCA1 and BRCA2 are essential proteins of the homologous recombination pathway of DSB repair, which are frequently mutated in cancers. Therefore, targeting PARP in BRCA1/2 deficient tumors is extremely attractive, as the increased load of unrepairable DSBs induces “synthetic lethality”, and has been shown to be a successful means of increasing tumor radiosensitivity (109, 111, 113, 114).

While the use of drugs to enhance radiosensitivity is a promising means of improving radiotherapy, drugs are also being used and developed to reduce or protect against adverse side effects induced by radiotherapy (109). Both early and late radiation induced side effects can be minimized by drug treatment before, during or after radiotherapy. Radical scavengers have been shown to effectively reduce many early side effects, as they bind with free radicals generated by ionizing radiation and thus reduce competition for oxygen (115). Radical scavenger drugs, such as amifostine, have been shown to have a radioprotective effect in normal tissue and are used clinically (115). Reducing late side

(16)

1

effects by therapeutic agents is less advanced than early side effects. However, targeting inflammation responses, particularly inhibition of TGFβ (116), is of interest in preventing/reducing late side effects. Furthermore, pilocarpine (117, 118), IGF (119, 120) and keratinocyte growth factor (KGF) (121) have been demonstrated to reduce radiation-induced salivary gland loss of function in in vivo models for radiotherapy of head and neck cancers, while early results indicate that ACE inhibition by Captopril may reduce the side effects of irradiation of normal heart and lung during thoracic irradiation (122).

Concluding remarks

While the benefits of radiotherapy greatly outweigh the adverse effects, and our knowledge of radiation treatment is ever-growing, there is still much to be understood about radiotherapy and its effects on normal tissue. There still remain many unanswered radiobiological questions. For instance: How can we best enhance the therapeutic window of radiation treatment? If a particular pathway is modulated to induce death in a tumor, what is the effect in normal tissue? If cell death is also enhanced in the normal tissue, then the overall benefit of modulating this pathway is lost, as there is no increase in the therapeutic index. The development of new models to study radiation-induced side effects and obtaining new knowledge regarding possible drug targets for radiosensitization may further enhance our knowledge.

Aim and outline of the thesis

The overall aim of this thesis is to establish and use new in vitro models to gain more relevant and personalized insight into the response of stem cells (both normal tissue and cancer stem cells) to radiotherapy, which can be used to enhance the overall efficacy of radiotherapy in the future. As discussed above, these new models (organoids) can potentially offer a more patient-specific response, while their 3D nature may represent a response more similar to the real in vivo patient response than the more commonly in vitro models. In this thesis, mechanisms of radioresistance, normal tissue damage and the effects of low dose irradiations on normal tissue stem cells are investigated. In vitro models to study these limitations are developed and possible targets to increase radiosensitivity are identified.

TGF-β signalling has been implicated in radioresistance in many cell lines. Thus, in Chapter 2, we investigate the role of ΔNp73, a truncated form of p73 (a member of the p53 superfamily), in TGF-β signalling.

(17)

24

1

Targets for increasing tumor radiosensitivity via means of drugs are an invaluable tool in radiation research and therapy. DDR proteins offer an intriguing target for increasing radiosensitivity due the importance of DNA damage in cell death mechanisms. In Chapter 3, a novel role in the DDR of Hsp70 (a molecular chaperone) and DNAJB6 (a co-chaperone of Hsp70) is investigated, offering a potential new drug target for enhanced radiosensitivity.

Cancer stem cells frequently are associated with an increased resistance to radiotherapy than other cells. In Chapter 4, a previously defined CSC-like subpopulation (CD44+/CD24-) of esophageal cancer cell lines was targeted via the mTOR pathway to reduce radiosensitivity. Furthermore, stimulation of mTOR pathway was investigated in patient derived esophageal cancer material containing this subpopulation.

Radiation therapy using charged particles spares normal tissue to a much greater extent than conventional radiotherapy. However, the effects of irradiations with varying LETs on normal tissue stem cells is not yet known, as in vitro models are not readily available to investigate this. In Chapter

5, a newly developed in vitro model for studying the effects of irradiation on normal tissue stem cells

is used to compare conventional photon irradiations with carbon ion irradiations with varying LET in terms of survival post-irradiation.

LDHRS/IRR has been shown in multiple tumor cell lines in the past. However, the clinical relevance of this phenomenon has remained elusive. Chapter 6 focuses on LDHRS/IRR in normal tissue stem cells both in vivo and in vitro. The effect of low dose irradiations is investigated with regard to function and survival of these cells, also with a more clinical fractionated dose. 

(18)

1

REFERENCES

1. Ferlay J, Soerjomataram I, Dikshit R, Eser S, Mathers C, Rebelo M, et al. Cancer incidence and mortality worldwide: Sources, methods and major patterns in GLOBOCAN 2012. Int J Cancer 2015;136(5):E359-86.

2. Bray F, Jemal A, Grey N, Ferlay J, Forman D. Global cancer transitions according to the human development index (2008-2030): A population-based study. Lancet Oncol 2012;13(8):790-801. 3. Wyld L, Audisio RA, Poston GJ. The evolution of cancer surgery and future perspectives. Nat Rev Clin Oncol 2015;12(2):115-24.

4. Chabner BA, Roberts TG,Jr. Timeline: Chemotherapy and the war on cancer. Nat Rev Cancer 2005;5(1):65-72.

5. Thariat J, Hannoun-Levi JM, Sun Myint A, Vuong T, Gerard JP. Past, present, and future of radiotherapy for the benefit of patients. Nat Rev Clin Oncol 2013;10(1):52-60.

6. Luke JJ, Flaherty KT, Ribas A, Long GV. Targeted agents and immunotherapies: Optimizing outcomes in melanoma. Nat Rev Clin Oncol 2017.

7. Gotwals P, Cameron S, Cipolletta D, Cremasco V, Crystal A, Hewes B, et al. Prospects for combining targeted and conventional cancer therapy with immunotherapy. Nat Rev Cancer 2017;17(5):286-301. 8. Khalil DN, Smith EL, Brentjens RJ, Wolchok JD. The future of cancer treatment: Immunomodulation, CARs and combination immunotherapy. Nat Rev Clin Oncol 2016;13(5):273-90.

9. Moding EJ, Kastan MB, Kirsch DG. Strategies for optimizing the response of cancer and normal tissues to radiation. Nat Rev Drug Discov 2013;12(7):526-42.

10. Barker HE, Paget JT, Khan AA, Harrington KJ. The tumour microenvironment after radiotherapy: Mechanisms of resistance and recurrence. Nat Rev Cancer 2015;15(7):409-25.

11. Chang L, Graham P, Hao J, Ni J, Deng J, Bucci J, et al. Cancer stem cells and signaling pathways in radioresistance. Oncotarget 2016;7(10):11002-17.

12. Wang D, Plukker JT, Coppes RP. Cancer stem cells with increased metastatic potential as a therapeutic target for esophageal cancer. Semin Cancer Biol 2017.

13. Baumann M, Krause M, Hill R. Exploring the role of cancer stem cells in radioresistance. Nat Rev Cancer 2008;8(7):545-54.

14. Alexander P, Charlesby A. Energy transfer in macromolecules exposed to ionizing radiations. Nature 1954;173(4404):578-9.

15. Liu J, Zhang J, Wang X, Li Y, Chen Y, Li K, et al. HIF-1 and NDRG2 contribute to hypoxia-induced radioresistance of cervical cancer hela cells. Exp Cell Res 2010;316(12):1985-93.

(19)

26

1

16. Li Z, Bao S, Wu Q, Wang H, Eyler C, Sathornsumetee S, et al. Hypoxia-inducible factors regulate tumorigenic capacity of glioma stem cells. Cancer Cell. 2009 2 June 2009;15(6):501-13.

17. Wang Y, Liu Y, Malek S, Zheng P, Liu Y. Targeting HIF1α eliminates cancer stem cells in hematological malignancies. Cell Stem Cell. 2011 8 April 2011;8(4):399-411.

18. Wouters BG, Koritzinsky M. Hypoxia signalling through mTOR and the unfolded protein response in cancer. Nat Rev Cancer 2008;8(11):851-64.

19. Zhu H, Wang D, Liu Y, Su Z, Zhang L, Chen F, et al. Role of the hypoxia-inducible factor-1 alpha induced autophagy in the conversion of non-stem pancreatic cancer cells into CD133(+) pancreatic cancer stem-like cells. Cancer Cell International 2013;13:119-.

20. Hashimoto O, Shimizu K, Semba S, Chiba S, Ku Y, Yokozaki H, et al. Hypoxia induces tumor aggressiveness and the expansion of CD133-positive cells in a hypoxia-inducible factor-1alpha-dependent manner in pancreatic cancer cells. Pathobiology 2011;78(4):181-92.

21. Turner BC, Haffty BG, Narayanan L, Yuan J, Havre PA, Gumbs AA, et al. Insulin-like growth factor-I receptor overexpression mediates cellular radioresistance and local breast cancer recurrence after lumpectomy and radiation. Cancer Res 1997;57(15):3079-83.

22. Skvortsova I, Skvortsov S, Stasyk T, Raju U, Popper BA, Schiestl B, et al. Intracellular signaling pathways regulating radioresistance of human prostate carcinoma cells. Proteomics 2008;8(21):4521-33.

23. Bentzen SM, Atasoy BM, Daley FM, Dische S, Richman PI, Saunders MI, et al. Epidermal growth factor receptor expression in pretreatment biopsies from head and neck squamous cell carcinoma as a predictive factor for a benefit from accelerated radiation therapy in a randomized controlled trial. J Clin Oncol 2005;23(24):5560-7.

24. Eicheler W, Krause M, Hessel F, Zips D, Baumann M. Kinetics of EGFR expression during fractionated irradiation varies between different human squamous cell carcinoma lines in nude mice. Radiother Oncol 2005;76(2):151-6.

25. Hatanpaa KJ, Burma S, Zhao D, Habib AA. Epidermal growth factor receptor in glioma: Signal transduction, neuropathology, imaging, and radioresistance. Neoplasia 2010;12(9):675-84.

26. Chakravarti A, Dicker A, Mehta M. The contribution of epidermal growth factor receptor (EGFR) signaling pathway to radioresistance in human gliomas: A review of preclinical and correlative clinical data. Int J Radiat Oncol Biol Phys 2004;58(3):927-31.

27. Kim AH, Lebman DA, Dietz CM, Snyder SR, Eley KW, Chung TD. Transforming growth factor-beta is an endogenous radioresistance factor in the esophageal adenocarcinoma cell line OE-33. Int J Oncol 2003;23(6):1593-9.

28. Lee JM, Bernstein A. P53 mutations increase resistance to ionizing radiation. Proc Natl Acad Sci U S A 1993;90(12):5742-6.

(20)

1

29. Bentzen SM. Preventing or reducing late side effects of radiation therapy: Radiobiology meets molecular pathology. Nat Rev Cancer 2006;6(9):702-13.

30. McDonald S, Rubin P, Phillips TL, Marks LB. Injury to the lung from cancer therapy: Clinical syndromes, measurable endpoints, and potential scoring systems. Int J Radiat Oncol Biol Phys 1995;31(5):1187-203.

31. Bronova I, Smith B, Aydogan B, Weichselbaum RR, Vemuri K, Erdelyi K, et al. Protection from radiation-induced pulmonary fibrosis by peripheral targeting of cannabinoid receptor-1. Am J Respir Cell Mol Biol 2015;53(4):555-62.

32. Hamilton CS, Denham JW, O’Brien M, Ostwald P, Kron T, Wright S, et al. Underprediction of human skin erythema at low doses per fraction by the linear quadratic model. Radiother Oncol 1996;40(1):23-30.

33. Tofilon PJ, Fike JR. The radioresponse of the central nervous system: A dynamic process. Radiat Res 2000;153(4):357-70.

34. Gruber S, Dörr W. Tissue reactions to ionizing radiation-oral mucosa. Mutat Res 2016;770(Pt B):292-8.

35. Jellema AP, Slotman BJ, Doornaert P, Leemans CR, Langendijk JA. Impact of radiation-induced xerostomia on quality of life after primary radiotherapy among patients with head and neck cancer. Int J Radiat Oncol Biol Phys 2007;69(3):751-60.

36. Deasy JO, Moiseenko V, Marks L, Chao KS, Nam J, Eisbruch A. Radiotherapy dose-volume effects on salivary gland function. Int J Radiat Oncol Biol Phys 2010;76(3 Suppl):S58-63.

37. Vissink A, van Luijk P, Langendijk JA, Coppes RP. Current ideas to reduce or salvage radiation damage to salivary glands. Oral Dis 2015;21(1):e1-e10.

38. Clarke M, Collins R, Darby S, Davies C, Elphinstone P, Evans V, et al. Effects of radiotherapy and of differences in the extent of surgery for early breast cancer on local recurrence and 15-year survival: An overview of the randomised trials. Lancet 2005;366(9503):2087-106.

39. Carr ZA, Land CE, Kleinerman RA, Weinstock RW, Stovall M, Griem ML, et al. Coronary heart disease after radiotherapy for peptic ulcer disease. Int J Radiat Oncol Biol Phys 2005;61(3):842-50.

40. Darby SC, Ewertz M, McGale P, Bennet AM, Blom-Goldman U, Bronnum D, et al. Risk of ischemic heart disease in women after radiotherapy for breast cancer. N Engl J Med 2013;368(11):987-98. 41. Rodemann HP, Bamberg M. Cellular basis of radiation-induced fibrosis. Radiother Oncol 1995;35(2):83-90.

42. Groarke JD, Nguyen PL, Nohria A, Ferrari R, Cheng S, Moslehi J. Cardiovascular complications of radiation therapy for thoracic malignancies: The role for non-invasive imaging for detection of cardiovascular disease. Eur Heart J 2014;35(10):612-23.

(21)

28

1

43. Williams JP, Johnston CJ, Finkelstein JN. Treatment for radiation-induced pulmonary late effects: Spoiled for choice or looking in the wrong direction? Curr Drug Targets 2010;11(11):1386-94. 44. Guha C, Kavanagh BD. Hepatic radiation toxicity: Avoidance and amelioration. Semin Radiat Oncol 2011;21(4):256-63.

45. Metcalfe C, Kljavin NM, Ybarra R, de Sauvage FJ. Lgr5+ stem cells are indispensable for radiation-induced intestinal regeneration. Cell Stem Cell 2014;14(2):149-59.

46. Wang JZ, Huang Z, Lo SS, Yuh WT, Mayr NA. A generalized linear-quadratic model for radiosurgery, stereotactic body radiation therapy, and high-dose rate brachytherapy. Sci Transl Med 2010;2(39):39ra48.

47. Brenner DJ. The linear-quadratic model is an appropriate methodology for determining isoeffective doses at large doses per fraction. Semin Radiat Oncol 2008;18(4):234-9.

48. Sheu T, Molkentine J, Transtrum MK, Buchholz TA, Withers HR, Thames HD, et al. Use of the LQ model with large fraction sizes results in underestimation of isoeffect doses. Radiother Oncol 2013;109(1):21-5.

49. Martin LM, Marples B, Lynch TH, Hollywood D, Marignol L. Exposure to low dose ionising radiation: Molecular and clinical consequences. Cancer Lett 2014;349(1):98-106.

50. Marples B, Joiner MC. The response of chinese hamster V79 cells to low radiation doses: Evidence of enhanced sensitivity of the whole cell population. Radiat Res 1993;133(1):41-51.

51. Wouters BG, Sy AM, Skarsgard LD. Low-dose hypersensitivity and increased radioresistance in a panel of human tumor cell lines with different radiosensitivity. Radiat Res 1996;146(4):399-413. 52. Short S, Mayes C, Woodcock M, Johns H, Joiner MC. Low dose hypersensitivity in the T98G human glioblastoma cell line. Int J Radiat Biol 1999;75(7):847-55.

53. Short SC, Woodcock M, Marples B, Joiner MC. Effects of cell cycle phase on low-dose hyper-radiosensitivity. Int J Radiat Biol 2003;79(2):99-105.

54. Krueger SA, Joiner MC, Weinfeld M, Piasentin E, Marples B. Role of apoptosis in low-dose hyper-radiosensitivity. Radiat Res 2007;167(3):260-7.

55. Beauchesne P, Bernier V, Carnin C, Taillandier L, Djabri M, Martin L, et al. Prolonged survival for patients with newly diagnosed, inoperable glioblastoma with 3-times daily ultrafractionated radiation therapy. Neuro Oncol 2010;12(6):595-602.

56. Lombaert IM, Brunsting JF, Wierenga PK, Faber H, Stokman MA, Kok T, et al. Rescue of salivary gland function after stem cell transplantation in irradiated glands. PLoS One 2008;3(4):e2063. 57. Maimets M, Rocchi C, Bron R, Pringle S, Kuipers J, Giepmans BN, et al. Long-term in vitro expansion of salivary gland stem cells driven by wnt signals. Stem Cell Reports 2016;6(1):150-62.

(22)

1

58. Benderitter M, Caviggioli F, Chapel A, Coppes RP, Guha C, Klinger M, et al. Stem cell therapies for the treatment of radiation-induced normal tissue side effects. Antioxid Redox Signal 2014;21(2):338-55.

59. Vos O, Davids JA, Weyzen WW, van Bekkum DW. Evidence for the cellular hypothesis in radiation protection by bone marrow cells. Acta Physiol Pharmacol Neerl 1956;4(4):482-6.

60. Thomas ED, Lochte HL,Jr, Lu WC, Ferrebee JW. Intravenous infusion of bone marrow in patients receiving radiation and chemotherapy. N Engl J Med 1957;257(11):491-6.

61. Acharya MM, Christie LA, Lan ML, Giedzinski E, Fike JR, Rosi S, et al. Human neural stem cell transplantation ameliorates radiation-induced cognitive dysfunction. Cancer Res 2011;71(14):4834-45.

62. Acharya MM, Christie LA, Hazel TG, Johe KK, Limoli CL. Transplantation of human fetal-derived neural stem cells improves cognitive function following cranial irradiation. Cell Transplant 2014;23(10):1255-66.

63. Klein D, Schmetter A, Imsak R, Wirsdorfer F, Unger K, Jastrow H, et al. Therapy with multipotent mesenchymal stromal cells protects lungs from radiation-induced injury and reduces the risk of lung metastasis. Antioxid Redox Signal 2016;24(2):53-69.

64. Klein D, Steens J, Wiesemann A, Schulz F, Kaschani F, Rock K, et al. Mesenchymal stem cell therapy protects lungs from radiation-induced endothelial cell loss by restoring superoxide dismutase 1 expression. Antioxid Redox Signal 2017;26(11):563-82.

65. Nanduri LS, Baanstra M, Faber H, Rocchi C, Zwart E, de Haan G, et al. Purification and ex vivo expansion of fully functional salivary gland stem cells. Stem Cell Reports 2014;3(6):957-64.

66. Pringle S, Maimets M, van der Zwaag M, Stokman MA, van Gosliga D, Zwart E, et al. Human salivary gland stem cells functionally restore radiation damaged salivary glands. Stem Cells 2016;34(3):640-52. 67. Pouton CW, Haynes JM. Embryonic stem cells as a source of models for drug discovery. Nat Rev Drug Discov 2007;6(8):605-16.

68. Kitambi SS, Chandrasekar G. Stem cells: A model for screening, discovery and development of drugs. Stem Cells Cloning 2011;4:51-9.

69. Hung SS, Khan S, Lo CY, Hewitt AW, Wong RC. Drug discovery using induced pluripotent stem cell models of neurodegenerative and ocular diseases. Pharmacol Ther 2017.

70. Dutta D, Heo I, Clevers H. Disease modeling in stem cell-derived 3D organoid systems. Trends Mol Med 2017;23(5):393-410.

71. Clevers H. Modeling development and disease with organoids. Cell 2016;165(7):1586-97.

(23)

30

1

cells in small intestine and colon by marker gene Lgr5. Nature 2007;449(7165):1003-7.

73. Sato T, Vries RG, Snippert HJ, van de Wetering M, Barker N, Stange DE, et al. Single Lgr5 stem cells build crypt-villus structures in vitro without a mesenchymal niche. Nature 2009;459(7244):262-5. 74. Sato T, Stange DE, Ferrante M, Vries RG, Van Es JH, Van den Brink S, et al. Long-term expansion of epithelial organoids from human colon, adenoma, adenocarcinoma, and barrett’s epithelium. Gastroenterology 2011;141(5):1762-72.

75. Dekkers JF, Wiegerinck CL, de Jonge HR, Bronsveld I, Janssens HM, de Winter-de Groot KM, et al. A functional CFTR assay using primary cystic fibrosis intestinal organoids. Nat Med 2013;19(7):939-45. 76. Schwank G, Koo BK, Sasselli V, Dekkers JF, Heo I, Demircan T, et al. Functional repair of CFTR by CRISPR/Cas9 in intestinal stem cell organoids of cystic fibrosis patients. Cell Stem Cell 2013;13(6):653-8.

77. Dekkers JF, Berkers G, Kruisselbrink E, Vonk A, de Jonge HR, Janssens HM, et al. Characterizing responses to CFTR-modulating drugs using rectal organoids derived from subjects with cystic fibrosis. Sci Transl Med 2016;8(344):344ra84.

78. Takasato M, Er PX, Chiu HS, Maier B, Baillie GJ, Ferguson C, et al. Kidney organoids from human iPS cells contain multiple lineages and model human nephrogenesis. Nature 2015;526(7574):564-8. 79. Huang L, Holtzinger A, Jagan I, BeGora M, Lohse I, Ngai N, et al. Ductal pancreatic cancer modeling and drug screening using human pluripotent stem cell- and patient-derived tumor organoids. Nat Med 2015;21(11):1364-71.

80. Eke I, Cordes N. Radiobiology goes 3D: How ECM and cell morphology impact on cell survival after irradiation. Radiotherapy and Oncology 2011;99(3):271-8.

81. Eke I, Hehlgans S, Sandfort V, Cordes N. 3D matrix-based cell cultures: Automated analysis of tumor cell survival and proliferation. Int J Oncol 2016;48(1):313-21.

82. Nutting CM, Morden JP, Harrington KJ, Urbano TG, Bhide SA, Clark C, et al. Parotid-sparing intensity modulated versus conventional radiotherapy in head and neck cancer (PARSPORT): A phase 3 multicentre randomised controlled trial. Lancet Oncol 2011;12(2):127-36.

83. Daly ME, Le QT, Jain AK, Maxim PG, Hsu A, Loo BW,Jr, et al. Intensity-modulated radiotherapy for locally advanced cancers of the larynx and hypopharynx. Head Neck 2011;33(1):103-11.

84. Mok G, Gauthier I, Jiang H, Huang SH, Chan K, Witterick IJ, et al. Outcomes of intensity-modulated radiotherapy versus conventional radiotherapy for hypopharyngeal cancer. Head Neck 2015;37(5):655-61.

85. Suh YG, Lee CG, Kim H, Choi EC, Kim SH, Kim CH, et al. Treatment outcomes of intensity-modulated radiotherapy versus 3D conformal radiotherapy for patients with maxillary sinus cancer in the postoperative setting. Head Neck 2016;38 Suppl 1:E207-13.

(24)

1

86. Eisbruch A, Foote RL, O’Sullivan B, Beitler JJ, Vikram B. Intensity-modulated radiation therapy for head and neck cancer: Emphasis on the selection and delineation of the targets. Semin Radiat Oncol 2002;12(3):238-49.

87. van Luijk P, Pringle S, Deasy JO, Moiseenko VV, Faber H, Hovan A, et al. Sparing the region of the salivary gland containing stem cells preserves saliva production after radiotherapy for head and neck cancer. Sci Transl Med 2015;7(305):305ra147.

88. van Luijk P, Faber H, Schippers JM, Brandenburg S, Langendijk JA, Meertens H, et al. Bath and shower effects in the rat parotid gland explain increased relative risk of parotid gland dysfunction after intensity-modulated radiotherapy. Int J Radiat Oncol Biol Phys 2009;74(4):1002-5.

89. Pow EH, Kwong DL, McMillan AS, Wong MC, Sham JS, Leung LH, et al. Xerostomia and quality of life after intensity-modulated radiotherapy vs. conventional radiotherapy for early-stage nasopharyngeal carcinoma: Initial report on a randomized controlled clinical trial. Int J Radiat Oncol Biol Phys 2006;66(4):981-91.

90. Bijl HP, van Luijk P, Coppes RP, Schippers JM, Konings AW, van der Kogel AJ. Unexpected changes of rat cervical spinal cord tolerance caused by inhomogeneous dose distributions. Int J Radiat Oncol Biol Phys 2003;57(1):274-81.

91. Bijl HP, van Luijk P, Coppes RP, Schippers JM, Konings AW, van der Kogel AJ. Influence of adjacent low-dose fields on tolerance to high doses of protons in rat cervical spinal cord. Int J Radiat Oncol Biol Phys 2006;64(4):1204-10.

92. Philippens ME, Pop LA, Visser AG, Peeters WJ, van der Kogel AJ. Bath and shower effect in spinal cord: The effect of time interval. Int J Radiat Oncol Biol Phys 2009;73(2):514-22.

93. Jermann M. Particle therapy statistics in 2014. Int. J. Particle Ther. 2015;2(1):50-4. 94. Wilson RR. Radiological use of fast protons. Radiology 1946;47(5):487-91.

95. Lomax AJ. Charged particle therapy: The physics of interaction. Cancer J 2009;15(4):285-91. 96. Newhauser WD, Zhang R. The physics of proton therapy. Phys Med Biol 2015;60(8):R155-209. 97. Durante M, Orecchia R, Loeffler JS. Charged-particle therapy in cancer: Clinical uses and future perspectives. Nat Rev Clin Oncol 2017.

98. Granovetter M. Proton radiotherapy for primary liver cancers. Lancet Oncol 2016;17(2):e49,2045(15)00616-6. Epub 2015 Dec 24.

99. Gregoire V, Langendijk JA, Nuyts S. Advances in radiotherapy for head and neck cancer. J Clin Oncol 2015;33(29):3277-84.

100. Asaithamby A, Chen DJ. Mechanism of cluster DNA damage repair in response to high-atomic number and energy particles radiation. Mutat Res 2011;711(1-2):87-99.

(25)

32

1

101. Niemantsverdriet M, van Goethem MJ, Bron R, Hogewerf W, Brandenburg S, Langendijk JA, et al. High and low LET radiation differentially induce normal tissue damage signals. Int J Radiat Oncol Biol Phys 2012;83(4):1291-7.

102. Paganetti H. Relative biological effectiveness (RBE) values for proton beam therapy. variations as a function of biological endpoint, dose, and linear energy transfer. Physics in Medicine & Biology 2014;59(22):R419.

103. Paganetti H, Niemierko A, Ancukiewicz M, Gerweck LE, Goitein M, Loeffler JS, et al. Relative biological effectiveness (RBE) values for proton beam therapy. International Journal of Radiation Oncology*Biology*Physics. 2002 1 June 2002;53(2):407-21.

104. Gerweck LE, Kozin SV. Relative biological effectiveness of proton beams in clinical therapy. Radiotherapy and Oncology. 1999 1 February 1999;50(2):135-42.

105. Sørensen BS, Bassler N, Nielsen S, Horsman MR, Grzanka L, Spejlborg H, et al. Relative biological effectiveness (RBE) and distal edge effects of proton radiation on early damage in vivo. Acta Oncol 2017;56(11):1387-91.

106. Ilicic K, Combs SE, Schmid TE. New insights in the relative radiobiological effectiveness of proton irradiation. Radiation Oncology 2018;13(1):6.

107. Cuaron JJ, Chang C, Lovelock M, Higginson DS, Mah D, Cahlon O, et al. Exponential increase in relative biological effectiveness along distal edge of a proton bragg peak as measured by deoxyribonucleic acid double-strand breaks. International Journal of Radiation Oncology*Biology*Physics. 2016 1 May 2016;95(1):62-9.

108. Chaudhary P, Marshall TI, Perozziello FM, Manti L, Currell FJ, Hanton F, et al. Relative biological effectiveness variation along monoenergetic and modulated bragg peaks of a 62-MeV therapeutic proton beam: A preclinical assessment. International Journal of Radiation Oncology*Biology*Physics. 2014 1 September 2014;90(1):27-35.

109. Begg AC, Stewart FA, Vens C. Strategies to improve radiotherapy with targeted drugs. Nat Rev Cancer 2011;11(4):239-53.

110. O’Connor MJ. Targeting the DNA damage response in cancer. Mol Cell 2015;60(4):547-60. 111. Farmer H, McCabe N, Lord CJ, Tutt AN, Johnson DA, Richardson TB, et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 2005;434(7035):917-21.

112. De Vos M, Schreiber V, Dantzer F. The diverse roles and clinical relevance of PARPs in DNA damage repair: Current state of the art. Biochem Pharmacol 2012;84(2):137-46.

113. Kaufman B, Shapira-Frommer R, Schmutzler RK, Audeh MW, Friedlander M, Balmana J, et al. Olaparib monotherapy in patients with advanced cancer and a germline BRCA1/2 mutation. J Clin Oncol 2015;33(3):244-50.

(26)

1

114. Livraghi L, Garber JE. PARP inhibitors in the management of breast cancer: Current data and future prospects. BMC Med 2015;13:188,015-0425-1.

115. Andreassen CN, Grau C, Lindegaard JC. Chemical radioprotection: A critical review of amifostine as a cytoprotector in radiotherapy. Semin Radiat Oncol 2003;13(1):62-72.

116. Anscher MS. Targeting the TGF-beta1 pathway to prevent normal tissue injury after cancer therapy. Oncologist 2010;15(4):350-9.

117. Coppes RP, Vissink A, Zeilstra LJ, Konings AW. Muscarinic receptor stimulation increases tolerance of rat salivary gland function to radiation damage. Int J Radiat Biol 1997;72(5):615-25.

118. Roesink JM, Konings AW, Terhaard CH, Battermann JJ, Kampinga HH, Coppes RP. Preservation of the rat parotid gland function after radiation by prophylactic pilocarpine treatment: Radiation dose dependency and compensatory mechanisms. Int J Radiat Oncol Biol Phys 1999;45(2):483-9.

119. Limesand KH, Said S, Anderson SM. Suppression of radiation-induced salivary gland dysfunction by IGF-1. PLoS One 2009;4(3):e4663.

120. Meyer S, Chibly AM, Burd R, Limesand KH. Insulin-like growth factor-1-mediated DNA repair in irradiated salivary glands is sirtuin-1 dependent. J Dent Res 2017;96(2):225-32.

121. Lombaert IM, Brunsting JF, Wierenga PK, Kampinga HH, de Haan G, Coppes RP. Keratinocyte growth factor prevents radiation damage to salivary glands by expansion of the stem/progenitor pool. Stem Cells 2008;26(10):2595-601.

122. van der Veen SJ, Ghobadi G, de Boer RA, Faber H, Cannon MV, Nagle PW, et al. ACE inhibition attenuates radiation-induced cardiopulmonary damage. Radiother Oncol 2015;114(1):96-103. 123. Huch M, Dorrell C, Boj SF, van Es JH, Li VS, van de Wetering M, et al. In vitro expansion of single Lgr5+ liver stem cells induced by wnt-driven regeneration. Nature 2013;494(7436):247-50.

124. Broutier L, Andersson-Rolf A, Hindley CJ, Boj SF, Clevers H, Koo BK, et al. Culture and establishment of self-renewing human and mouse adult liver and pancreas 3D organoids and their genetic manipulation. Nat Protoc 2016;11(9):1724-43.

125. Pouget JP, Frelon S, Ravanat JL, Testard I, Odin F, Cadet J. Formation of modified DNA bases in cells exposed either to gamma radiation or to high-LET particles. Radiat Res 2002;157(5):589-95.

126. Dizdaroglu M, Jaruga P. Mechanisms of free radical-induced damage to DNA. Free Radic Res 2012;46(4):382-419.

127. Lomax ME, Folkes LK, O’Neill P. Biological consequences of radiation-induced DNA damage: Relevance to radiotherapy. Clinical Oncology. 2013 October 2013;25(10):578-85.

128. Vignard J, Mirey G, Salles B. Ionizing-radiation induced DNA double-strand breaks: A direct and indirect lighting up. Radiother Oncol 2013;108(3):362-9.

(27)

34

1

129. David SS, O’Shea VL, Kundu S. Base-excision repair of oxidative DNA damage. Nature 2007;447(7147):941-50.

130. Stewart FA, Te Poele JA, Van der Wal AF, Oussoren YG, Van Kleef EM, Kuin A, et al. Radiation nephropathy--the link between functional damage and vascular mediated inflammatory and thrombotic changes. Acta Oncol 2001;40(8):952-7.

131. Bartfeld S, Bayram T, van de Wetering M, Huch M, Begthel H, Kujala P, et al. In Vitro expansion of human gastric epithelial stem cells and their responses to bacterial infection. Gastroenterology. 2015 January 2015;148(1):126,136.e6.

132. Barkauskas CE, Chung MI, Fioret B, Gao X, Katsura H, Hogan BL. Lung organoids: Current uses and future promise. Development 2017;144(6):986-97.

133. Cruz NM, Song X, Czerniecki SM, Gulieva RE, Churchill AJ, Kim YK, et al. Organoid cystogenesis reveals a critical role of microenvironment in human polycystic kidney disease. Nat Mater 2017;16(11):1112-9.

(28)
(29)

Referenties

GERELATEERDE DOCUMENTEN

The use of organoids in the study of radiation response and therapeutic window Nagle, Peter.. IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if

Second, the Hek293 cells were used because they do express endogenous p53 and p73 including ∆Np73 (figures S1 and 2D) and express very high levels of transfected DNA (figure

Mouse salivary gland organoids were cultured and irradiated with 7 Gy photons (IR) at day 5 in culture (D5) and collected 7 days later (D12). b) Pie charts showing cell

These findings indicate that the relative biological effect of cellular senescence is similar to what was measured based on SGO survival (Fig.. Induction of cellular senescence

chimera (PROTAC) can be used to design new compounds which are highly specific to senescent cells while less toxic to the normal cells. 2) to define ‘the good and the bad’

Het hoofddoel van dit proefschrift is te onderzoeken wat de rol is van door straling geïnduceerde senescente speekselkliercellen en hoe door overlevende stamcellen te stimuleren

In conclusion, we have demonstrated that low and low LET radiation induce a similar senescence response at equivalent cell survival based doses, however, with

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright