Next Article in Journal
Striatal GDNF Neurons Chemoattract RET-Positive Dopamine Axons at Seven Times Farther Distance Than Medium Spiny Neurons
Previous Article in Journal
Tumour Microenvironment: The General Principles of Pathogenesis and Implications in Diffuse Large B Cell Lymphoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Potential Benefits of Combining Proton or Carbon Ion Therapy with DNA Damage Repair Inhibitors

1
Department of Radiation Biology, Institute for Cancer Research, Norwegian Radium Hospital, Oslo University Hospital, 0379 Oslo, Norway
2
Department of Life and Environmental Physics, Horia Hulubei National Institute of Physics and Nuclear Engineering, 077125 Magurele, Romania
3
UMR6252 CIMAP, Team Applications in Radiobiology with Accelerated Ions, CEA-CNRS-ENSICAEN-Université de Caen Normandie, 14000 Caen, France
*
Author to whom correspondence should be addressed.
Cells 2024, 13(12), 1058; https://doi.org/10.3390/cells13121058
Submission received: 29 April 2024 / Revised: 3 June 2024 / Accepted: 5 June 2024 / Published: 19 June 2024
(This article belongs to the Topic Innovative Radiation Therapies)

Abstract

:
The use of charged particle radiotherapy is currently increasing, but combination therapy with DNA repair inhibitors remains to be exploited in the clinic. The high-linear energy transfer (LET) radiation delivered by charged particles causes clustered DNA damage, which is particularly effective in destroying cancer cells. Whether the DNA damage response to this type of damage is different from that elicited in response to low-LET radiation, and if and how it can be targeted to increase treatment efficacy, is not fully understood. Although several preclinical studies have reported radiosensitizing effects when proton or carbon ion irradiation is combined with inhibitors of, e.g., PARP, ATR, ATM, or DNA-PKcs, further exploration is required to determine the most effective treatments. Here, we examine what is known about repair pathway choice in response to high- versus low-LET irradiation, and we discuss the effects of inhibitors of these pathways when combined with protons and carbon ions. Additionally, we explore the potential effects of DNA repair inhibitors on antitumor immune signaling upon proton and carbon ion irradiation. Due to the reduced effect on healthy tissue and better immune preservation, particle therapy may be particularly well suited for combination with DNA repair inhibitors.

1. Introduction

Radiotherapy is frequently used in cancer treatment. In traditional radiotherapy, high-energy X-rays are employed. However, particle radiotherapy with protons or carbon ions has become more widely available in recent years [1]. The main reasoning for introducing particle radiotherapy is the beneficial depth dose distribution of particles. Most of the particles’ energy is deposited at the end of their trajectory as they are slowed down, causing a sharp dose peak named the Bragg peak [2] (Figure 1A). By directing this peak to the tumor, particle irradiation can thus result in improved sparing of surrounding normal tissue, reducing treatment-related side effects [2,3]. The beam energy determines the depth of the Bragg peak, and by using beams with different energies in the so-called spread-out Bragg peak (SOBP), one can cover the whole tumor volume with a uniform dose (Figure 1B).
Another notable feature of particle radiotherapy is the elevated ionization density, leading to more clustered DNA damage. The ionization density reflects the absorbed radiation dose per unit length, also termed linear energy transfer (LET). As protons or carbon ions traverse tissue, they are slowed down, and the probability for ionizations increases. There is therefore a gradual increase in LET from the beam entrance towards the end of the Bragg peak. Clinical carbon ion beams produce high-LET radiation, with typical LET values within the spread-out Bragg peak (SOBP) of 40–80 keV/µm [4] and >200 keV/µm at the distal end [5,6], depending on the beam energy. For clinical proton beams, average LET values are approximately 2–3 keV/µm at the middle of the SOBP and 10–15 keV/µm at the distal end of the Bragg peak [7,8]. In comparison, the LET value for a clinical 6 MV X-ray beam is about 2 keV/µm [9]. The average LET values of proton beams are thus similar or modestly elevated compared to those of X-ray beams. However, these are average values, and proton beams exhibit heterogeneity due to the multiple proton energies present in the SOBP [10,11]. This can lead to a considerable amount of proton radiation with higher LET values, even at the middle of the SOBP [11].
The clustered DNA damage may consist of several DNA double-strand breaks and/or other types of DNA lesions in close proximity to each other (also termed complex DNA damage, see Figure 1B). This damage may be particularly hard to repair and often results in cancer cell death [12]. Importantly, the ionization density is highest at the very distal end of the Bragg peak, typically located in the surrounding margin outside the tumor volume. The clustered damage could thus in some cases affect nearby critical normal tissues, resulting in increased adverse effects. There is therefore a high need to better understand how cells respond to DNA damage caused by the high-LET particle irradiation, both in order to find ways to enhance tumor cell killing and to avoid radiation-induced side effects [1,10].
Previously, a somewhat common assumption has been that DNA repair is not very important following high-LET irradiation, as the clustered DNA damage has been considered almost non-reparable [13,14,15]. Upon high-LET irradiation, the shape of the survival curve is typically straight-lined at low doses, indicating lack of repair [15]. Furthermore, the cell cycle is decoupled from radiosensitivity as LET increases [16], further indicating that DNA repair is not contributing much to cell survival after high-LET irradiation. This view was, e.g., supported by previous experiments showing higher relative biological effectiveness (RBE) for repair-proficient than repair-deficient (Ku80 negative) Xrs5 Chinese hamster ovarian (CHO) cells after irradiation with carbon ions [17]. Consequently, it was proposed that the benefit of carbon ion therapy, over conventional X-ray therapy, would be largest in tumors with high repair capacity [17].
However, more recent studies indicate that DNA repair is highly important even after high-LET irradiation [18,19]. Combination treatment with DNA repair inhibitors is thus relevant for both carbon ions and protons. Here, we discuss potential differences in repair signaling after high- and low-LET irradiation that may impact such treatments (Figure 2). We also examine previous preclinical work involving DNA repair inhibitors combined with proton or carbon ion irradiation. Moreover, we discuss how such combination treatment may promote antitumor immune signaling. We propose that DNA repair inhibitors are useful in combination with particle radiotherapy, causing tumor radiosensitization and enhancing antitumor immune effects.

2. Differences in Repair Signaling for High- and Low-LET Irradiation

The question of whether high- and low-LET radiation induces different repair pathways has not yet been conclusively answered. Interestingly, many previous studies have shown a delayed resolution of DNA repair foci after high-LET irradiation, as assessed through immunofluorescence microscopy. For example, a higher number of persistent γH2AX and 53BP1 foci, two commonly used markers for DNA double-strand breaks, has been observed following irradiation with carbon ions compared to X-irradiation [20,21,22,23]. Additionally, a slower resolution of 53BP1 and γH2AX foci was found in cells irradiated at the distal end compared to the front or center position of the Bragg peak for clinical proton beams [24,25]. These studies suggest that DNA repair may be slower following high-LET compared to low-LET irradiation. Notably, slower repair could likely involve different use of repair pathways, meaning that specific pathways are involved in the repair of the clustered damage induced by high-LET particles.
On the other hand, the γH2AX and 53BP1 foci caused by high-LET irradiation are large and may not represent single DNA double-strand breaks [20,26]. Thus, repair may occur on a subset of the breaks without the foci disappearing, leading to the underestimation of repair rates. Interestingly, the assessment of DNA double-strand breaks by pulse-field gel electrophoresis has shown rapid rejoining for most of the breaks even after irradiation with helium and nitrogen ions with very high LET values ranging from 80 to 225 keV/µm [27]. The measurement of DNA double-strand breaks via the neutral comet assay has also revealed similar repair rates for protons with different LETs (1 and 12 keV/μm) [28]. It is noteworthy that in the latter study, when applying the alkaline comet assay, slower repair was observed for the high-LET protons. This suggests that high-LET irradiation may cause a slower resolution of single-strand breaks/alkali-labile sites [28].
Consistent with the notion that specific repair pathways may be involved in the repair of DNA double-strand breaks after high-LET irradiation, it has been proposed that clustered DNA damage generally induces more DNA end resection [12]. This is a process where nucleases such as MRE11, EXO1, and CtIP (RBBP8) generate single-stranded DNA at sites of DNA double-strand breaks, thereby facilitating the binding of specific repair proteins [29]. Accordingly, DNA double-strand break repair pathways requiring end resection, such as alternative non-homologous end joining (alt-NHEJ) and homologous recombination (HR), may be more important after high-LET than low-LET irradiation [12]. The classical NHEJ (c-NHEJ) repair pathway, by contrast, does not require end resection [29]. In support of increased end resection after high-LET irradiation, more foci of the single-strand binding protein RPA were observed in G2 phase cells following carbon ion compared to X-irradiation [23]. Furthermore, exposure to heavy ions with a broad spectrum of LET values ranging from 90 to 15,000 keV/µm showed a clear positive correlation between the number of RPA foci with increasing LET in both G2 and G1 phase cells [30]. The increased resection in the G1 phase likely promotes error-prone alt-NHEJ [30]. It has also been reported that high-LET irradiation leads to a more pronounced activation of the ATR-dependent G2 checkpoint [31,32,33,34,35]. As activation of the ATR kinase may depend on end resection mechanisms [36], this further supports that end resection is increased.
Several studies have demonstrated a higher dependency on HR repair after high-LET as compared to low-LET irradiation, for instance in comparisons of carbon ion and X-irradiation [37,38]. Increased dependency on HR repair has also been reported for proton irradiation at relatively low average LET values of less than 10–15 keV/µm [39,40,41,42]. However, other studies did not find such correlations between HR repair and LET [43], or the effects varied between different cell lines [21]. The link between LET and HR repair is thus not entirely clear [44,45]. In addition, the role of c-NHEJ after high-LET irradiation is incompletely understood. Intriguingly, it has been shown that small DNA fragments associated with clustered DNA damage suppress the binding of KU70/80 (XRCC6/5), the initial step of the c-NHEJ pathway [46,47]. This suggests a reduced involvement of c-NHEJ in repairing damage caused by high-LET irradiation. Notably, electron microscopy studies have shown that particle tracks induce the largest DNA damage clusters in densely packed chromatin [26,48]. As large clusters may yield the highest number of small fragments, the inhibition of NHEJ by small DNA fragments may likely occur specifically in regions of heterochromatin [48]. Conversely, other studies have shown that c-NHEJ remains the major repair pathway following both low- and high-LET irradiation. It was found that the loss of c-NHEJ gave bigger radiosensitizing effects compared to the loss of HR, even for carbon ion and other heavy ion irradiation [38,49,50].
As mentioned above, the clustered damage sites may contain different types of DNA damage. Therefore, the repair of base damage or single-strand breaks may also be important in addition to the DNA double-strand break repair pathways. Indeed, a recent study showed that the base excision repair protein 8-oxoguanine DNA glycosylase (OGG1) contributes to radioresistance after high-LET (12 keV/µm) proton irradiation but not low-LET (1 keV/µm) proton or X-irradiation [51]. Furthermore, repair focus analysis, with antibodies to OGG1 as a surrogate marker for base damage, showed that these foci were more persistent after high-LET as compared to low-LET proton irradiation [51]. In cells irradiated with high-LET iron ions, persistent OGG1 foci co-localized with XRCC1 and 53BP1 foci, in agreement with the repair of clustered damage being dependent on multiple repair pathways [52]. An interesting issue is whether the clustered damage sites may cause obstacles to the replication fork, potentially creating additional DNA double-strand breaks when cells undergo DNA replication several hours after irradiation. It has been demonstrated that the loss of SLX4-MUS81, which is involved in resolving stalled forks, leads to a bigger sensitization of cells irradiated with protons than with X-rays [53]. The loss of FANCD2, which binds SLX4, also sensitizes cells to protons [11]. Although these factors may also be involved in regulating HR repair of the radiation-induced breaks, one intriguing explanation could be that proton-induced clustered DNA damage causes more stalling of the DNA replication forks, resulting in increased DNA breakage when cells are lacking SLX4-MUS81 or FANCD2 [53]. The loss of PARG, which also plays roles in resolving stalled replication forks, has been shown to cause bigger sensitization to protons than to X-rays [51]. Moreover, DNA fiber assays have revealed a bigger increase in replication stalling following proton compared to X-irradiation of KRAS-mutated cells [54]. However, the study did not find similar effects in KRAS wild-type cells. More work is needed to understand the interaction between DNA replication and clustered damage sites following high-LET irradiation.

3. Previous Studies of DNA Repair Inhibitors Combined with Protons or Carbon Ions

Several inhibitors involved in DNA repair are under clinical investigation as radiosensitizers in combination with traditional X-ray radiotherapy. However, their effects in combination with particle radiotherapy have not been fully elucidated. Below we discuss the radiosensitizing effects observed in recent preclinical studies with carbon ions or protons in combination with inhibitors of various DNA damage response proteins.

3.1. PARP Inhibitors

It is well known that the loss of function of HR repair sensitizes cells to PARP inhibitors [55]. In cells defective in HR, DNA repair is performed by other processes such as NHEJ, alt-NHEJ, and single-strand break repair [56]. Several of these alternative repair pathways are regulated by PARP, leaving the cells sensitive to PARP inhibition [56]. Indeed, when PARP is inhibited, it leads to single-strand break accumulation, stalling of the replication fork, and the generation of many toxic double-strand breaks [57]. These double-strand breaks are poorly repaired, if at all, with error-prone repair mechanisms. This leads to genetic instability and eventually cell death, by synthetic lethality [57,58]. Combining irradiation with PARP inhibitors offers a promising approach for cancer treatment [57,58,59,60]. Radioresistant cancer cells can often have alterations in the HR or other DNA repair pathways [61], leaving them sensitive to PARP inhibitors which can help overcome this resistance [62,63]. The effect is even more pronounced using high-LET irradiation, such as carbon ions and protons [64]. The inhibition of PARP prevents the repair of single-strand breaks, while high-LET radiation induces more double-strand breaks, altogether overwhelming the compromised repair machinery of cancer cells with defective HR. Additionally, PARP is required for alt-NHEJ [12], which is thought to be particularly important for the repair of DNA damage following high-LET irradiation (as mentioned in Section 2).
The combination of radiation with various LETs and PARP inhibitors has been tested in several cellular models. A study comparing olaparib treatment combined with Bragg peak protons (LET: 4.3 keV/μm) versus entrance beam protons (LET: 0.5 keV/μm) indicated that the most effective radiosensitization occurred with the higher LET of Bragg peak protons in lung and pancreatic cancer cells [65]. Similar results were obtained with HeLa cells in another study, where olaparib treatment was combined with protons with LET values of 12 keV/μm versus 1 keV/μm [66]. Even for BRCA-proficient breast cancer cells, greater radiosensitization was observed with protons (LET within 3.0–8.1 keV/μm) than with X-rays when combined with olaparib [67]. Moreover, in 3D spheroid models of oropharyngeal squamous cell carcinoma, both olaparib and talazoparib were found to sensitize to both low-LET protons (~1 keV/µm) and X-rays. Talazoparib exhibited enhanced efficacy over olaparib specifically for proton irradiation, possibly due to talazoparib’s stronger PARP trapping ability [68]. Olaparib also sensitized pancreatic ductal adenocarcinoma xenograft models to chemoradiotherapy in vivo, with greater radiosensitizing effects observed with SOBP protons (LET within 2–5 keV/μm) compared to X-rays [69]. The combination of olaparib and proton irradiation was also effective in platinum- and radiation-resistant esophageal cancer [70]. Niraparib, another PARP inhibitor, increased the RBE of protons versus photons in human head and neck squamous cell carcinoma (HNSCC) cell lines [71]. (In the two latter reports the proton LET was not mentioned).
A comparison of high-LET (>100 keV/μm) carbon ion irradiation and low-LET gamma radiation reported greater radiosensitization with PARP inhibition for the carbon ions [72]. Additionally, triple-negative breast cancer cell lines irradiated with carbon ions (80 keV/μm) or X-rays in combination with olaparib showed radiosensitization mainly for BRCA-mutated cells, and the sensitizing effect was bigger for carbon ions than for X-rays [73]. Carbon ions (50 keV/μm) in combination with talazoparib also reduced the stem cell proportion in radioresistant glioma [74]. A comparison of X-rays, protons (11 keV/µm), and carbon ions (73 keV/µm) in combination with olaparib showed a greater impact of high-LET carbon ion irradiation on chondrosarcoma cell survival, but radiosensitization with olaparib was observed for all types of radiation. The sensitizing effect was biggest for protons, followed by carbon ions and X-rays [75]. In summary, numerous previous studies suggest that the radiosensitizing effects of PARP inhibitors are more pronounced for carbon ion and proton irradiation than for X-irradiation.

3.2. ATR, CHK1, and WEE1 Inhibitors

The ATR kinase and its downstream target CHK1 kinase mediate G2 checkpoint arrest and HR repair after irradiation and are also the main responders to replication stress [76,77,78,79,80,81]. Similarly, the WEE1 kinase is a crucial regulator of the G2 checkpoint, HR repair, and replication stress [77,78,82,83,84]. If HR is indeed particularly important for the repair of clustered DNA damage, and such damage potently stalls replication forks, inhibitors of ATR, CHK1, or WEE1 may be particularly good radiosensitizers for high-LET radiation. The stronger activation of the G2 checkpoint after high-LET irradiation may further prove a greater efficacy of these inhibitors in combination with high-LET particles.
However, studies comparing X-irradiation with proton and carbon ion irradiation in combination with ATR inhibitors have most often reported smaller or equal radiosensitizing effects for the highest LET irradiation. For instance, treatment of cancer cells with the ATR inhibitor AZD6738, in combination with protons of two LET values (1 and 7 keV/μm) or X-rays, revealed equal radiosensitization for all radiation types [85]. Furthermore, although cancer cells treated with VE-821, another ATR inhibitor, were sensitized to carbon ion irradiation (LET: 70 keV/µm) [86], the sensitization enhancement ratio (SER) was slightly lower than that seen for X-rays. ATR inhibition also led to the abrogation of the G2 checkpoint and increased formation of micronuclei. The data in this study are in good agreement with our own data showing G2 checkpoint abrogation and decreased clonogenic survival upon the co-treatment of glioblastoma cells with the ATR inhibitor VE-822 and carbon ion irradiation of two LET values (28 and 73 keV/μm), but with a slightly smaller effect for the higher LET value as compared to the lower LET value and X-irradiation [33]. A study of chondrosarcoma cell proliferation has also reported a sensitizing effect of VE-821 combined with carbon ion irradiation (LET: 55 keV/µm) [87]. The sensitizing effect was higher than that observed for relatively low-LET protons (2.9 keV/μm), but lower than for X-rays. In HPV-negative HNSCC cell lines in 2D culture, the sensitizing effect of VE-821 was diminished even by low-LET proton irradiation (~1 keV/µm) as compared to that observed in X-irradiated samples [88]. However, the same study showed that growth of tumor spheroids was delayed by ATR inhibition after exposure to both radiation types. Another study comparing protons of a relatively high LET (9.9 keV/µm) with X-rays in combination with AZD6738, found that the RBE measured by clonogenic survival was increased in two out of four cancer cell lines [89].
Several studies have also addressed the effects of CHK1 inhibitors in combination with X-irradiation versus proton or carbon ion irradiation. In breast cancer cell lines irradiated at the SOBP of a 230 MeV proton beam, clonogenic survival seemed to be equally well reduced in the proton and X-irradiated cells upon co-treatment with the CHK1 inhibitor PF-477736 [90]. Similar results were observed in 3D pancreatic tumor models treated with a different inhibitor (prexasertib) in combination with low-LET protons (3.7 keV/µm) and X-rays [91]. Another study reported that clonogenic survival in one out of two lung cancer cell lines was decreased by co-treatment with the CHK1 inhibitor (AZD7762) and carbon ions (50 keV/µm), but not X-irradiation [92]. Furthermore, a synergistic effect between high-LET carbon ion (184 keV/µm) irradiation and the CHK1 inhibitor UCN-01 was observed in cancer stem-like cells of HNSCC [93]. Synergy was observed for comparatively more doses of carbon ion than of X-radiation. However, in both studies with carbon ions, the radiosensitizing effects of CHK1 inhibition were very small. Reports on WEE1 inhibition in combination with high-LET irradiation are sparse. However, one study showed that the WEE1 inhibitor AZD1775 caused the abrogation of G2 arrest and reduced survival in lung cancer cells irradiated with carbon ions (50 keV/μm) [94]. The radiosensitizing effect was, however, quite low, and lower than that observed for X-irradiated cells in combination with the WEE1 inhibitor. Taken together, based on the literature to date, the radiosensitizing effects of ATR, CHK1, and WEE1 inhibitors do not appear to be bigger for high-LET than for low-LET radiation.

3.3. DNA-PKcs Inhibitors

The DNA-dependent protein kinase (DNA-PK) catalytic subunit (DNA-PKcs) is the key driver of the c-NHEJ pathway. The DNA-PKcs is also implicated in mitosis, telomere maintenance, and other cellular processes [95,96]. As mentioned above, it has been suggested that c-NHEJ may be less involved in the repair of DNA damage caused by high-LET irradiation, due to the inhibiting effect of small DNA fragments (see Section 2). One might thus expect that DNA-PKcs inhibitors would not likely work well in combination with high-LET carbon ion irradiation. However, preclinical studies have reported that DNA-PKcs inhibitors can act as radiosensitizing agents to protons as well as carbon ions. The potent DNA-PKcs inhibitor NU7026 significantly enhanced H2AX phosphorylation and reduced clonogenic survival in non-small cell lung cancer H1299 cells exposed to carbon ion (50 keV/µm) and X-radiation [49], although with a slightly reduced SER for carbon ions. The same inhibitor increased the sensitivity of lung cancer A549 cells to carbon ion radiation (49 keV/µm), enhancing cell death, reducing DNA damage repair, exacerbating cell cycle G2/M phase arrest, and increasing apoptosis [97]. In comparison with X-rays, apoptosis seemed to be more pronounced with carbon ions. In a study by Bright et al. [89], the effects of NU7441, another DNA-PKcs inhibitor, in parallel with other inhibitors targeting the DNA damage response (e.g., ATM, ATR, RAD51, PARP), was investigated in combination with proton (9.9 keV/µm) or X-irradiation in lung cancer (H460 and H1299), pancreatic cancer (PANC-1 and Panc 10.05), and normal endothelial (HUVEC) cells. NU7441 was the only inhibitor that caused a significant radiosensitization of all cell lines, but with either higher or lower SER for protons than for X-rays depending on the cell line. In studies comparing treatment of 3D cancer cell cultures with low-LET protons (3.7 keV/μm) versus X-rays in combination with NU7026, similar radiosensitization was observed for protons and X-rays [91,98]. The radiosensitizing effect of the DNA-PKcs inhibitor M3814 to carbon ion irradiation has also been studied under hypoxic (1% O2) compared to normoxic (21% O2) conditions. While the oxygen enhancement ratio for the X-irradiation of A549 cells was 1.4, no significant oxygen effect was found after carbon ion irradiation (LET value not reported, but irradiation was done at SOBP). Interestingly, the radiosensitizing effect of M3814 was found to be stronger at hypoxia than normoxia when combined with carbon ion irradiation [99]. These reports highlight the ability of DNA-PKcs inhibitors to enhance cell death by particle irradiation, likely through the inhibition of c-NHEJ. However, the underlying mechanism is not completely clear. There is evidence that both NU7026 and NU7441 increase the radiosensitivity of cancer cells to carbon ion therapy independently of DNA DSB repair [100,101]. DNA-PKcs inhibition by NU7026 led to rapid telomere length shortening together with extensive senescence in MCF-7 breast cancer cells exposed to carbon ion radiation [100]. Furthermore, NU7441 radiosensitized H1299 lung cancer cells via a significant G2/M cell cycle arrest and a higher level of senescence, with no apparent inhibition of DNA repair [101]. Thus, more studies using protons and carbon ions are needed to uncover the detailed mechanisms of radiosensitization by DNA-PKcs inhibitors for therapeutic optimization.

3.4. ATM Inhibitors

The ATM kinase is involved in the repair of DNA double-strand breaks, being considered a key protein in the HR pathway [102]. ATM also promotes resection-dependent repair of DNA double-strand breaks during the G1 phase and serves as a major regulator of cell cycle checkpoints [103]. While many previous studies have demonstrated radiosensitization by ATM inhibitors in combination with X-rays in a variety of preclinical tumor models, only a few studies have been performed with a focus on the LET dependency of the effects.
In one study, the ATM inhibitor AZD0156 exhibited greater radiosensitization of cancer cells exposed to proton radiation at the Bragg peak (7 keV/μm) of a 76.8 MeV spot scanning proton beam than to X-rays or low-LET protons (2.2 keV/µm; irradiated at the proton beam entrance). This radiosensitization was found to depend on the NHEJ repair proteins XRCC4 and LIG4, suggesting that ATM inhibition in combination with Bragg peak irradiation may lead to the increased use of NHEJ with toxic consequences [85]. Interestingly, this study noted increased phosphorylation of CHK1 and RPA32, but not of CHK2, following Bragg peak proton irradiation as compared to X-irradiation, indicating heightened activation of ATR signaling rather than ATM signaling. However, only ATM inhibition selectively sensitized cells to Bragg peak protons. In the abovementioned study by Bright et al., with protons of LET 9.9 keV/μm in comparison to X-rays, it was observed that KU55933, another ATM inhibitor, sensitized H460 lung cancer cells more to protons than X-rays, but H1299 lung cancer cells were sensitized more to X-rays than protons [89]. The radiosensitizing effects seen in the aforementioned studies in 3D cancer cell cultures [91,98] were largely similar for treatment with KU55933 in combination with low-LET protons (3.7 keV/μm) and X-rays. Similarly, a study on colorectal cancer organoids with the two ATM inhibitors AZ32 and KU55933 showed roughly similar sensitization to both proton and X-irradiation. Although the LET of protons was not specified in this study, it was likely relatively low, as irradiation was conducted in the middle of the SOBP [104]. Additionally, a study treating uveal cancer cells with KU55933 in combination with X-rays and low-LET protons (1 keV/µm) also demonstrated comparable sensitization to both types of radiation [105]. Recently, we showed that AZD1390, another potent ATM inhibitor, could radiosensitize glioblastoma cells to both X-rays and carbon ions (28 and 72 keV/μm), albeit with slightly lower sensitization observed for the higher-LET carbon ions as compared to X-rays [33]. ATM inhibitors, with their low single-agent toxicity, may show clinical success as sensitizers for both X-rays and particles. It is worth mentioning, however, that the ATM inhibitor KU55933 significantly reduced the clonogenic survival of normal human fibroblasts irradiated with carbon ions (LET: 70 keV/µm) in vitro [106].

4. Antitumor Immune Signaling Induced by DNA Repair Inhibitors

Recent studies have shown that combination treatment of DNA repair inhibitors and X-rays can yield increased antitumor immune effects relative to those seen after irradiation alone. This can occur through several different mechanisms. A type 1 interferon (IFN) response may be elevated due to enhanced micronucleus formation through the activation of the cGAS/STING/TBK1/IRF3 pathway [107,108]. Micronuclei are formed when cells with unrepaired or misrepaired DNA damage undergo mitosis [80,109], and the abrogation of cell cycle checkpoints or defective repair can enhance micronucleus formation after irradiation [80,109,110,111]. Multiple studies have shown that ATR, PARP, and CHK1 inhibitors can increase the IFN response through this mechanism after irradiation [110,111,112,113,114,115]. Interestingly, after treatment with ATR or ATM inhibitors and X-rays, the cytosolic RNA sensor RIG-1 may also initiate IFN signaling [116,117]. Cytosolic RNA may arise from the activation of transposable elements [118,119], resulting from the opening of compacted chromatin and transcription of DNA that is normally silent [120]. Cytosolic RNA or DNA can also be released from mitochondria, as has been shown in cells irradiated with X-rays [121]. Inhibition of ATM in non-irradiated cells has also been shown to facilitate the release of mitochondrial DNA [122]. Another way of increasing immune responses is by inducing immunogenic cell death, in which dying cells present damage-associated molecular patterns (DAMPs) that stimulate dendritic cells and other immune cells [123]. Some studies report the increased presentation of immunogenic cell death markers HMGB1, ATP, and/or calreticulin upon treatment with X-rays and either ATR inhibitors in vitro [124] or PARP inhibitors in vitro and in vivo [113,125]. DNA repair inhibitors have also been shown to regulate immunosuppressive effects after X-irradiation. For example, ATR inhibition impairs the radiation-induced upregulation of the immune checkpoint molecules PD-L1 [126,127,128] and CD47 [126].
As of now, very few studies have addressed how combined treatment with DNA repair inhibitors and high-LET particle radiation affects immune responses, but the antitumor immune signaling could likely be enhanced (Figure 3). We recently observed higher levels of secreted IFN-β from cells treated with an ATR inhibitor in combination with carbon ions and high-LET protons (38 keV/µm) as compared to low-LET protons and X-rays [33]. The mechanism underlying the increased IFN response when the ATR inhibitor was combined with high-LET irradiation is not known. However, higher-LET radiation alone has been shown to induce higher levels of micronuclei as compared to lower-LET radiation [129,130]. Combined treatment with DNA repair inhibitors and proton or carbon ion radiation may therefore induce the formation of even more micronuclei than the combination of such inhibitors with X-rays. Indeed, one study observed increased radiosensitization and micronucleus formation when an inhibitor of ATR was combined with high-LET carbon ion (70 keV/µm) as compared to X-irradiation [86]. Increased micronucleus formation following DNA repair inhibition and high-LET irradiation could thus likely result in higher IFN-β levels. Additionally, in a recent study, higher IFN-β signaling after proton than photon irradiation was associated with proton-induced activation of transposable elements, likely due to the aforementioned opening of compacted chromatin [131]. As non-repaired clustered DNA damage can cause changes in chromatin structure [132], it is thus tempting to speculate that co-treatment with DNA repair inhibitors and high-LET radiation could result in the further opening of compacted chromatin, and thereby heighten the transcription of retrotransposons or other transposable elements. This could likely result in cytosolic dsRNA that stimulates RIG-1-dependent IFN transcription [131]. Moreover, another potential mechanism that could give an elevated IFN response after treatment with DNA repair inhibitors and high-LET radiation is the leakage of small, radiation-induced DNA fragments from the nucleus into the cytosol [107]. Due to the high ionization density of high-LET irradiation, small DNA fragments are produced at clustered damage sites. We speculate that the inhibition of repair may potentially enhance and prolong this effect, resulting in cGAS/STING-dependent IFN production after leakage of these fragments into the cytosol.
The ability of proton or carbon ion radiation to induce immunogenic cell death, as measured by the hallmark factors calreticulin, HMGB1, and ATP, has been demonstrated in several model systems [129,133,134]. Although multiple studies show similar induction of immunogenic cell death after low- and high-LET irradiation, certain studies have found that high-LET carbon ion (50–70 keV/µm) irradiation increases the secretion of HMGB1 to a greater level than lower-LET irradiation (13 keV/µm or lower) [134,135]. However, repair inhibitors were not addressed in these studies. As ATR and PARP inhibitors can increase X-ray-induced immunogenic cell death, they likely also induce immunogenic cell death to at least the same extent when combined with protons or carbon ions. Moreover, several studies have demonstrated the enhanced expression of the immune checkpoint protein PD-L1 also after proton and carbon ion irradiation [136,137,138]. The increase in PD-L1 was shown to be dependent on ATR and CHK1 after carbon ion irradiation [137,138], and the inhibitors of these kinases were able to impair the PD-L1 upregulation. This gives another rationale for combining high-LET irradiation with ATR or CHK1 inhibition.
Of note is that proton and carbon ion radiotherapy may induce less lymphopenia compared to classical radiotherapy with X-rays [139,140]. Lymphopenia leads to immunosuppression and is considered a major limiting factor for combination treatments of radiotherapy and immunotherapy. Irradiation with particles induces less lymphopenia because of the reduced radiation dose to the surrounding volumes of normal tissue. This will result in the sparing of circulating lymphocytes, lymph nodes, and/or bone marrow [141]. The reduced lymphopenia with carbon ions or protons will likely also be highly important upon co-treatment with DNA repair inhibitors, making it possible to achieve higher antitumor immune responses than for DNA repair inhibitors in combination with X-rays.

5. Conclusions and Future Perspectives

Recent studies have shown that low- and high-LET irradiation may cause different responses in the cells. The increased clustered DNA damage after high-LET irradiation may induce a different spectrum of DNA damage signaling events. Furthermore, the use of DNA repair pathways and the induction of antitumor immune signaling may be different for low- and high-LET irradiation. Based on the current knowledge, DNA resection and activation of the HR repair pathway seem to be increased upon high-LET compared to low-LET irradiation. As normal cells are more often non-cycling than cancer cells, targeting the repair pathways of cycling cells (e.g., HR repair) may partly provide tumor-specific radiosensitization. However, some of the preclinical studies conducted so far do not find that HR deficiency is associated with increased sensitivity to high-LET irradiation, and there are also conflicting results regarding the involvement of the NHEJ repair pathway. Of note is that HR repair occurs mainly in the S and G2 phases, and differences in cell cycle progression between the various cancer cell lines may thus be important. High-LET irradiation typically causes a stronger G2 checkpoint arrest compared to low-LET irradiation, and this could likely lead to the detection of more resection and HR repair, but the extent of arrest could vary between cell lines. It is also possible that high statistical uncertainty in some of the experimental assays (e.g., the clonogenic survival assay), or in the radiation dose measurements or LET values, may explain some of the discrepancies.
Better knowledge about the differences between responses at low- and high-LET could potentially provide a rationale for trials with so-called “LET painting”, a proton beam delivery technique to achieve a higher ionization density not only towards the distal end of the irradiation field but also in the center of the tumor [142]. Additionally, a better understanding of the relationship between cancer-associated genetic alterations and responses to low- and high-LET particle irradiation may help to guide patient selection for particle radiotherapy in the future.
It will be highly important to further investigate combination treatments of particle therapy with DNA repair inhibitors or other radiosensitizing drugs. Even for inhibitors showing similar sensitization of the tumor cells to low- and high-LET radiation, the tumor-specific radiosensitizing effects will likely be improved for particle radiotherapy. Due to the reduced radiation dose to the surrounding healthy tissue, there will be less radiosensitization of normal cells. High-throughput screening, directly comparing the effects of drugs in combination with X-rays versus protons or carbon ions, will likely be useful to identify the most efficient combinations. Notably, a joint activity of several repair pathways may be required to survive clustered DNA damage. Hence, a stronger and more specific sensitization to high-LET particle radiation may be obtained through drug combinations. Emerging preclinical work with traditional X-ray-based radiotherapy shows that DNA repair inhibitors, e.g., inhibitors of ATR or PARP, do not only lead to cellular radiosensitization, but can also enhance antitumor immune signaling in many cases. However, very little is known about antitumor immune signaling upon combination treatment with such inhibitors and carbon ion or proton radiation, and this should be explored in future studies evaluating both immunogenic and potential immunosuppressive effects. Finally, to promote clinical translation, further preclinical assessment of the combination treatments will be needed in immunocompetent mouse models. The evaluation of triple combinations of particle radiotherapy, radiosensitizing drugs, and immune checkpoint inhibition will also be valuable.

Author Contributions

Conceptualization and writing of manuscript draft, R.G.S. and G.E.R.; writing parts of original draft, all authors; finding references, all authors; figures, G.E.R., A.E.M., S.H. and R.G.S.; review and revision of manuscript text, all authors; funding acquisition, D.I.S. and R.G.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Norway Romania Grant RO-NO-2019-0510, contract no. 41/2021 and The Norwegian Cancer Society (projects no. 198018 and 245570).

Acknowledgments

Figure 3 contains elements from SMART Servier Medical Art by Laboratoires Servier (CC BY 4.0).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dosanjh, M.; Jones, B.; Pawelke, J.; Pruschy, M.; Sorensen, B.S. Overview of research and therapy facilities for radiobiological experimental work in particle therapy. Report from the European Particle Therapy Network radiobiology group. Radiother. Oncol. 2018, 128, 14–18. [Google Scholar] [CrossRef]
  2. Grau, C.; Durante, M.; Georg, D.; Langendijk, J.A.; Weber, D.C. Particle therapy in Europe. Mol. Oncol. 2020, 14, 1492–1499. [Google Scholar] [CrossRef]
  3. Spiotto, M.T.; McGovern, S.L.; Gunn, G.B.; Grosshans, D.; McAleer, M.F.; Frank, S.J.; Paulino, A.C. Proton Radiotherapy to Reduce Late Complications in Childhood Head and Neck Cancers. Int. J. Part. Ther. 2021, 8, 155–167. [Google Scholar] [CrossRef]
  4. Tinganelli, W.; Durante, M. Carbon Ion Radiobiology. Cancers 2020, 12, 3022. [Google Scholar] [CrossRef]
  5. Chailapakul, P.; Maloney, O.; Hirakawa, H.; Fujimori, A.; Kitamura, H.; Kato, T.A. The contribution of high-LET track to DNA damage formation and cell death for Monoenergy and SOBP carbon ion irradiation. Biochem. Biophys. Res. Commun. 2024, 696, 149500. [Google Scholar] [CrossRef]
  6. Kantemiris, I.; Karaiskos, P.; Papagiannis, P.; Angelopoulos, A. Dose and dose averaged LET comparison of 1H, 4He, 6Li, 8Be, 10B, 12C, 14N, and 16O ion beams forming a spread-out Bragg peak. Med. Phys. 2011, 38, 6585–6591. [Google Scholar] [CrossRef]
  7. Grassberger, C.; Paganetti, H. Elevated LET components in clinical proton beams. Phys. Med. Biol. 2011, 56, 6677–6691. [Google Scholar] [CrossRef]
  8. Paganetti, H. Relative biological effectiveness (RBE) values for proton beam therapy. Variations as a function of biological endpoint, dose, and linear energy transfer. Phys. Med. Biol. 2014, 59, R419–R472. [Google Scholar] [CrossRef]
  9. Okamoto, H.; Kanai, T.; Kase, Y.; Matsumoto, Y.; Furusawa, Y.; Fujita, Y.; Saitoh, H.; Itami, J.; Kohno, T. Relation between lineal energy distribution and relative biological effectiveness for photon beams according to the microdosimetric kinetic model. J. Radiat. Res. 2011, 52, 75–81. [Google Scholar] [CrossRef]
  10. Horendeck, D.; Walsh, K.D.; Hirakawa, H.; Fujimori, A.; Kitamura, H.; Kato, T.A. High LET-Like Radiation Tracks at the Distal Side of Accelerated Proton Bragg Peak. Front. Oncol. 2021, 11, 690042. [Google Scholar] [CrossRef]
  11. Liu, Q.; Ghosh, P.; Magpayo, N.; Testa, M.; Tang, S.; Gheorghiu, L.; Biggs, P.; Paganetti, H.; Efstathiou, J.A.; Lu, H.M.; et al. Lung cancer cell line screen links fanconi anemia/BRCA pathway defects to increased relative biological effectiveness of proton radiation. Int. J. Radiat. Oncol. Biol. Phys. 2015, 91, 1081–1089. [Google Scholar] [CrossRef]
  12. Mladenova, V.; Mladenov, E.; Stuschke, M.; Iliakis, G. DNA Damage Clustering after Ionizing Radiation and Consequences in the Processing of Chromatin Breaks. Molecules 2022, 27, 1540. [Google Scholar] [CrossRef]
  13. Chen, G.T.; Castro, J.R.; Quivey, J.M. Heavy charged particle radiotherapy. Annu. Rev. Biophys. Bioeng. 1981, 10, 499–529. [Google Scholar] [CrossRef]
  14. Eguchi-Kasai, K.; Murakami, M.; Itsukaichi, H.; Fukutsu, K.; Kanai, T.; Furusawa, Y.; Sato, K.; Ohara, H.; Yatagai, F. The role of DNA repair on cell killing by charged particles. Adv. Space Res. 1996, 18, 109–118. [Google Scholar] [CrossRef]
  15. Hall, E.J.; Giaccia, A.J. Radiobiology for the Radiologist, 6th ed.; Lippincott William & Wilkins: Philadelphia, PA, USA, 2006. [Google Scholar]
  16. Bird, R.P.; Burki, H.J. Survival of synchronized Chinese hamster cells exposed to radiation of different linear-energy transfer. Int. J. Radiat. Biol. Relat. Stud. Phys. Chem. Med. 1975, 27, 105–120. [Google Scholar] [CrossRef]
  17. Weyrather, W.K.; Ritter, S.; Scholz, M.; Kraft, G. RBE for carbon track-segment irradiation in cell lines of differing repair capacity. Int. J. Radiat. Biol. 1999, 75, 1357–1364. [Google Scholar] [CrossRef]
  18. Flint, D.B.; Bright, S.J.; McFadden, C.H.; Konishi, T.; Ohsawa, D.; Turner, B.; Lin, S.H.; Grosshans, D.R.; Chiu, H.S.; Sumazin, P.; et al. Cell lines of the same anatomic site and histologic type show large variability in intrinsic radiosensitivity and relative biological effectiveness to protons and carbon ions. Med. Phys. 2021, 48, 3243–3261. [Google Scholar] [CrossRef]
  19. Furusawa, Y.; Nakano-Aoki, M.; Matsumoto, Y.; Hirayama, R.; Kobayashi, A.; Konishi, T. Equivalency of the quality of sublethal lesions after photons and high-linear energy transfer ion beams. J. Radiat. Res. 2017, 58, 803–808. [Google Scholar] [CrossRef]
  20. Averbeck, N.B.; Topsch, J.; Scholz, M.; Kraft-Weyrather, W.; Durante, M.; Taucher-Scholz, G. Efficient Rejoining of DNA Double-Strand Breaks despite Increased Cell-Killing Effectiveness following Spread-Out Bragg Peak Carbon-Ion Irradiation. Front. Oncol. 2016, 6, 28. [Google Scholar] [CrossRef]
  21. Lopez Perez, R.; Nicolay, N.H.; Wolf, J.C.; Frister, M.; Schmezer, P.; Weber, K.J.; Huber, P.E. DNA damage response of clinical carbon ion versus photon radiation in human glioblastoma cells. Radiother. Oncol. 2019, 133, 77–86. [Google Scholar] [CrossRef]
  22. Oonishi, K.; Cui, X.; Hirakawa, H.; Fujimori, A.; Kamijo, T.; Yamada, S.; Yokosuka, O.; Kamada, T. Different effects of carbon ion beams and X-rays on clonogenic survival and DNA repair in human pancreatic cancer stem-like cells. Radiother. Oncol. 2012, 105, 258–265. [Google Scholar] [CrossRef]
  23. Shibata, A.; Conrad, S.; Birraux, J.; Geuting, V.; Barton, O.; Ismail, A.; Kakarougkas, A.; Meek, K.; Taucher-Scholz, G.; Lobrich, M.; et al. Factors determining DNA double-strand break repair pathway choice in G2 phase. EMBO J. 2011, 30, 1079–1092. [Google Scholar] [CrossRef]
  24. Chaudhary, P.; Marshall, T.I.; Currell, F.J.; Kacperek, A.; Schettino, G.; Prise, K.M. Variations in the Processing of DNA Double-Strand Breaks Along 60-MeV Therapeutic Proton Beams. Int. J. Radiat. Oncol. Biol. Phys. 2016, 95, 86–94. [Google Scholar] [CrossRef]
  25. Hojo, H.; Dohmae, T.; Hotta, K.; Kohno, R.; Motegi, A.; Yagishita, A.; Makinoshima, H.; Tsuchihara, K.; Akimoto, T. Difference in the relative biological effectiveness and DNA damage repair processes in response to proton beam therapy according to the positions of the spread out Bragg peak. Radiat. Oncol. 2017, 12, 111. [Google Scholar] [CrossRef]
  26. Lorat, Y.; Brunner, C.U.; Schanz, S.; Jakob, B.; Taucher-Scholz, G.; Rube, C.E. Nanoscale analysis of clustered DNA damage after high-LET irradiation by quantitative electron microscopy—The heavy burden to repair. DNA Repair 2015, 28, 93–106. [Google Scholar] [CrossRef]
  27. Stenerlow, B.; Hoglund, E.; Carlsson, J.; Blomquist, E. Rejoining of DNA fragments produced by radiations of different linear energy transfer. Int. J. Radiat. Biol. 2000, 76, 549–557. [Google Scholar] [CrossRef]
  28. Carter, R.J.; Nickson, C.M.; Thompson, J.M.; Kacperek, A.; Hill, M.A.; Parsons, J.L. Complex DNA Damage Induced by High Linear Energy Transfer Alpha-Particles and Protons Triggers a Specific Cellular DNA Damage Response. Int. J. Radiat. Oncol. Biol. Phys. 2018, 100, 776–784. [Google Scholar] [CrossRef]
  29. Huertas, P. DNA resection in eukaryotes: Deciding how to fix the break. Nat. Struct. Mol. Biol. 2010, 17, 11–16. [Google Scholar] [CrossRef]
  30. Averbeck, N.B.; Ringel, O.; Herrlitz, M.; Jakob, B.; Durante, M.; Taucher-Scholz, G. DNA end resection is needed for the repair of complex lesions in G1-phase human cells. Cell Cycle 2014, 13, 2509–2516. [Google Scholar] [CrossRef]
  31. Mladenova, V.; Mladenov, E.; Scholz, M.; Stuschke, M.; Iliakis, G. Strong Shift to ATR-Dependent Regulation of the G2-Checkpoint after Exposure to High-LET Radiation. Life 2021, 11, 560. [Google Scholar] [CrossRef]
  32. Moertel, H.; Georgi, J.C.; Distel, L.; Eyrich, W.; Fritsch, M.; Grabenbauer, G.; Sauer, R. Effects of low energy protons on clonogenic survival, DSB repair and cell cycle in human glioblastoma cells and B14 fibroblasts. Radiother. Oncol. 2004, 73 (Suppl. S2), S115–S118. [Google Scholar] [CrossRef]
  33. Rødland, G.E.; Temelie, M.; Serban, A.M.; Eek Mariampillai, A.; Edin, N.F.J.; Malinen, E.; Gilbert, A.; Chevailer, F.; Savu, D.I.; Syljuåsen, R.G. Interferon signaling is enhanced by ATR inhibition in glioblastoma cells irradiated with X-rays, protons or carbon ions. bioRxiv 2024. [Google Scholar] [CrossRef]
  34. Suetens, A.; Konings, K.; Moreels, M.; Quintens, R.; Verslegers, M.; Soors, E.; Tabury, K.; Gregoire, V.; Baatout, S. Higher Initial DNA Damage and Persistent Cell Cycle Arrest after Carbon Ion Irradiation Compared to X-irradiation in Prostate and Colon Cancer Cells. Front. Oncol. 2016, 6, 87. [Google Scholar] [CrossRef]
  35. Tu, W.; Dong, C.; Fu, J.; Pan, Y.; Kobayashi, A.; Furusawa, Y.; Konishi, T.; Shao, C. Both irradiated and bystander effects link with DNA repair capacity and the linear energy transfer. Life Sci. 2019, 222, 228–234. [Google Scholar] [CrossRef]
  36. Zou, L.; Elledge, S.J. Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 2003, 300, 1542–1548. [Google Scholar] [CrossRef]
  37. Cartwright, I.M.; Su, C.; Haskins, J.S.; Salinas, V.A.; Sunada, S.; Yu, H.; Uesaka, M.; Hirakawa, H.; Chen, D.J.; Fujimori, A.; et al. DNA Repair Deficient Chinese Hamster Ovary Cells Exhibiting Differential Sensitivity to Charged Particle Radiation under Aerobic and Hypoxic Conditions. Int. J. Mol. Sci. 2018, 19, 2228. [Google Scholar] [CrossRef]
  38. Gerelchuluun, A.; Manabe, E.; Ishikawa, T.; Sun, L.; Itoh, K.; Sakae, T.; Suzuki, K.; Hirayama, R.; Asaithamby, A.; Chen, D.J.; et al. The major DNA repair pathway after both proton and carbon-ion radiation is NHEJ, but the HR pathway is more relevant in carbon ions. Radiat. Res. 2015, 183, 345–356. [Google Scholar] [CrossRef]
  39. Bright, S.J.; Flint, D.B.; Chakraborty, S.; McFadden, C.H.; Yoon, D.S.; Bronk, L.; Titt, U.; Mohan, R.; Grosshans, D.R.; Sumazin, P.; et al. Nonhomologous End Joining Is More Important Than Proton Linear Energy Transfer in Dictating Cell Death. Int. J. Radiat. Oncol. Biol. Phys. 2019, 105, 1119–1125. [Google Scholar] [CrossRef]
  40. Fontana, A.O.; Augsburger, M.A.; Grosse, N.; Guckenberger, M.; Lomax, A.J.; Sartori, A.A.; Pruschy, M.N. Differential DNA repair pathway choice in cancer cells after proton- and photon-irradiation. Radiother. Oncol. 2015, 116, 374–380. [Google Scholar] [CrossRef]
  41. Grosse, N.; Fontana, A.O.; Hug, E.B.; Lomax, A.; Coray, A.; Augsburger, M.; Paganetti, H.; Sartori, A.A.; Pruschy, M. Deficiency in homologous recombination renders Mammalian cells more sensitive to proton versus photon irradiation. Int. J. Radiat. Oncol. Biol. Phys. 2014, 88, 175–181. [Google Scholar] [CrossRef]
  42. Szymonowicz, K.; Krysztofiak, A.; Linden, J.V.; Kern, A.; Deycmar, S.; Oeck, S.; Squire, A.; Koska, B.; Hlouschek, J.; Vullings, M.; et al. Proton Irradiation Increases the Necessity for Homologous Recombination Repair Along with the Indispensability of Non-Homologous End Joining. Cells 2020, 9, 889. [Google Scholar] [CrossRef]
  43. Guerra Liberal, F.D.C.; Parsons, J.L.; McMahon, S.J. Most DNA repair defects do not modify the relationship between relative biological effectiveness and linear energy transfer in CRISPR-edited cells. Med. Phys. 2024, 51, 591–600. [Google Scholar] [CrossRef]
  44. Nikitaki, Z.; Velalopoulou, A.; Zanni, V.; Tremi, I.; Havaki, S.; Kokkoris, M.; Gorgoulis, V.G.; Koumenis, C.; Georgakilas, A.G. Key biological mechanisms involved in high-LET radiation therapies with a focus on DNA damage and repair. Expert Rev. Mol. Med. 2022, 24, e15. [Google Scholar] [CrossRef]
  45. Vitti, E.T.; Parsons, J.L. The Radiobiological Effects of Proton Beam Therapy: Impact on DNA Damage and Repair. Cancers 2019, 11, 946. [Google Scholar] [CrossRef]
  46. Pang, D.; Winters, T.A.; Jung, M.; Purkayastha, S.; Cavalli, L.R.; Chasovkikh, S.; Haddad, B.R.; Dritschilo, A. Radiation-generated short DNA fragments may perturb non-homologous end-joining and induce genomic instability. J. Radiat. Res. 2011, 52, 309–319. [Google Scholar] [CrossRef]
  47. Wang, H.; Wang, X.; Zhang, P.; Wang, Y. The Ku-dependent non-homologous end-joining but not other repair pathway is inhibited by high linear energy transfer ionizing radiation. DNA Repair 2008, 7, 725–733. [Google Scholar] [CrossRef]
  48. Lorat, Y.; Timm, S.; Jakob, B.; Taucher-Scholz, G.; Rube, C.E. Clustered double-strand breaks in heterochromatin perturb DNA repair after high linear energy transfer irradiation. Radiother. Oncol. 2016, 121, 154–161. [Google Scholar] [CrossRef]
  49. Ma, H.; Takahashi, A.; Yoshida, Y.; Adachi, A.; Kanai, T.; Ohno, T.; Nakano, T. Combining carbon ion irradiation and non-homologous end-joining repair inhibitor NU7026 efficiently kills cancer cells. Radiat. Oncol. 2015, 10, 225. [Google Scholar] [CrossRef]
  50. Takahashi, A.; Kubo, M.; Ma, H.; Nakagawa, A.; Yoshida, Y.; Isono, M.; Kanai, T.; Ohno, T.; Furusawa, Y.; Funayama, T.; et al. Nonhomologous end-joining repair plays a more important role than homologous recombination repair in defining radiosensitivity after exposure to high-LET radiation. Radiat. Res. 2014, 182, 338–344. [Google Scholar] [CrossRef]
  51. Fabbrizi, M.R.; Nickson, C.M.; Hughes, J.R.; Robinson, E.A.; Vaidya, K.; Rubbi, C.P.; Kacperek, A.; Bryant, H.E.; Helleday, T.; Parsons, J.L. Targeting OGG1 and PARG radiosensitises head and neck cancer cells to high-LET protons through complex DNA damage persistence. Cell Death Dis. 2024, 15, 150. [Google Scholar] [CrossRef]
  52. Asaithamby, A.; Hu, B.; Chen, D.J. Unrepaired clustered DNA lesions induce chromosome breakage in human cells. Proc. Natl. Acad. Sci. USA 2011, 108, 8293–8298. [Google Scholar] [CrossRef]
  53. Liu, Q.; Underwood, T.S.A.; Kung, J.; Wang, M.; Lu, H.M.; Paganetti, H.; Held, K.D.; Hong, T.S.; Efstathiou, J.A.; Willers, H. Disruption of SLX4-MUS81 Function Increases the Relative Biological Effectiveness of Proton Radiation. Int. J. Radiat. Oncol. Biol. Phys. 2016, 95, 78–85. [Google Scholar] [CrossRef]
  54. Al Zubaidi, T.; Gehrisch, O.H.F.; Genois, M.M.; Liu, Q.; Lu, S.; Kung, J.; Xie, Y.; Schuemann, J.; Lu, H.M.; Hata, A.N.; et al. Targeting the DNA replication stress phenotype of KRAS mutant cancer cells. Sci. Rep. 2021, 11, 3656. [Google Scholar] [CrossRef]
  55. Hutchinson, L. Targeted therapies: PARP inhibitor olaparib is safe and effective in patients with BRCA1 and BRCA2 mutations. Nat. Rev. Clin. Oncol. 2010, 7, 549. [Google Scholar] [CrossRef]
  56. Heitz, F.; Harter, P.; Ewald-Riegler, N.; Papsdorf, M.; Kommoss, S.; du Bois, A. Poly(ADP-ribosyl)ation polymerases: Mechanism and new target of anticancer therapy. Expert Rev. Anticancer Ther. 2010, 10, 1125–1136. [Google Scholar] [CrossRef]
  57. Farmer, H.; McCabe, N.; Lord, C.J.; Tutt, A.N.; Johnson, D.A.; Richardson, T.B.; Santarosa, M.; Dillon, K.J.; Hickson, I.; Knights, C.; et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 2005, 434, 917–921. [Google Scholar] [CrossRef]
  58. Cesaire, M.; Thariat, J.; Candeias, S.M.; Stefan, D.; Saintigny, Y.; Chevalier, F. Combining PARP inhibition, radiation, and immunotherapy: A possible strategy to improve the treatment of cancer? Int. J. Mol. Sci. 2018, 19, 3793. [Google Scholar] [CrossRef]
  59. Gilbert, A.; Tudor, M.; Montanari, J.; Commenchail, K.; Savu, D.I.; Lesueur, P.; Chevalier, F. Chondrosarcoma Resistance to Radiation Therapy: Origins and Potential Therapeutic Solutions. Cancers 2023, 15, 1962. [Google Scholar] [CrossRef]
  60. Calabrese, C.R.; Almassy, R.; Barton, S.; Batey, M.A.; Calvert, A.H.; Canan-Koch, S.; Durkacz, B.W.; Hostomsky, Z.; Kumpf, R.A.; Kyle, S.; et al. Anticancer chemosensitization and radiosensitization by the novel poly(ADP-ribose) polymerase-1 inhibitor AG14361. J. Natl. Cancer Inst. 2004, 96, 56–67. [Google Scholar] [CrossRef]
  61. Hu, Y.; Guo, M. Synthetic lethality strategies: Beyond BRCA1/2 mutations in pancreatic cancer. Cancer Sci. 2020, 111, 3111–3121. [Google Scholar] [CrossRef]
  62. Chevalier, F. Counteracting Radio-Resistance Using the Optimization of Radiotherapy. Int. J. Mol. Sci. 2020, 21, 1767. [Google Scholar] [CrossRef]
  63. Thariat, J.; Valable, S.; Laurent, C.; Haghdoost, S.; Peres, E.A.; Bernaudin, M.; Sichel, F.; Lesueur, P.; Cesaire, M.; Petit, E.; et al. Hadrontherapy Interactions in Molecular and Cellular Biology. Int. J. Mol. Sci. 2019, 21, 133. [Google Scholar] [CrossRef]
  64. Lesueur, P.; Chevalier, F.; Austry, J.B.; Waissi, W.; Burckel, H.; Noel, G.; Habrand, J.L.; Saintigny, Y.; Joly, F. Poly-(ADP-ribose)-polymerase inhibitors as radiosensitizers: A systematic review of pre-clinical and clinical human studies. Oncotarget 2017, 8, 69105–69124. [Google Scholar] [CrossRef]
  65. Hirai, T.; Saito, S.; Fujimori, H.; Matsushita, K.; Nishio, T.; Okayasu, R.; Masutani, M. Radiosensitization by PARP inhibition to proton beam irradiation in cancer cells. Biochem. Biophys. Res. Commun. 2016, 478, 234–240. [Google Scholar] [CrossRef]
  66. Carter, R.J.; Nickson, C.M.; Thompson, J.M.; Kacperek, A.; Hill, M.A.; Parsons, J.L. Characterisation of Deubiquitylating Enzymes in the Cellular Response to High-LET Ionizing Radiation and Complex DNA Damage. Int. J. Radiat. Oncol. Biol. Phys. 2019, 104, 656–665. [Google Scholar] [CrossRef]
  67. Park, S.; Choi, C.; Kim, H.; Shin, Y.J.; Oh, Y.; Park, W.; Cho, W.K.; Kim, N. Olaparib enhances sensitization of BRCA-proficient breast cancer cells to x-rays and protons. Breast Cancer Res. Treat. 2024, 203, 449–461. [Google Scholar] [CrossRef]
  68. Zhou, C.; Fabbrizi, M.R.; Hughes, J.R.; Grundy, G.J.; Parsons, J.L. Effectiveness of PARP inhibition in enhancing the radiosensitivity of 3D spheroids of head and neck squamous cell carcinoma. Front. Oncol. 2022, 12, 940377. [Google Scholar] [CrossRef]
  69. Waissi, W.; Nicol, A.; Jung, M.; Rousseau, M.; Jarnet, D.; Noel, G.; Burckel, H. Radiosensitizing Pancreatic Cancer with PARP Inhibitor and Gemcitabine: An In Vivo and a Whole-Transcriptome Analysis after Proton or Photon Irradiation. Cancers 2021, 13, 527. [Google Scholar] [CrossRef]
  70. Kageyama, S.I.; Du, J.; Hojo, H.; Motegi, A.; Nakamura, M.; Tsuchihara, K.; Akimoto, T. PARP inhibitor olaparib sensitizes esophageal carcinoma cells to fractionated proton irradiation. J. Radiat. Res. 2020, 61, 177–186. [Google Scholar] [CrossRef]
  71. Wang, L.; Cao, J.; Wang, X.; Lin, E.; Wang, Z.; Li, Y.; Li, Y.; Chen, M.; Wang, X.; Jiang, B.; et al. Proton and photon radiosensitization effects of niraparib, a PARP-1/-2 inhibitor, on human head and neck cancer cells. Head Neck 2020, 42, 2244–2256. [Google Scholar] [CrossRef]
  72. Ghorai, A.; Bhattacharyya, N.P.; Sarma, A.; Ghosh, U. Radiosensitivity and Induction of Apoptosis by High LET Carbon Ion Beam and Low LET Gamma Radiation: A Comparative Study. Scientifica 2014, 2014, 438030. [Google Scholar] [CrossRef]
  73. Kawanishi, M.; Fujita, M.; Karasawa, K. Combining Carbon-Ion Irradiation and PARP Inhibitor, Olaparib Efficiently Kills BRCA1-Mutated Triple-Negative Breast Cancer Cells. Breast Cancer 2022, 16, 11782234221080553. [Google Scholar] [CrossRef]
  74. Lesueur, P.; Chevalier, F.; El-Habr, E.A.; Junier, M.P.; Chneiweiss, H.; Castera, L.; Muller, E.; Stefan, D.; Saintigny, Y. Radiosensitization Effect of Talazoparib, a Parp Inhibitor, on Glioblastoma Stem Cells Exposed to Low and High Linear Energy Transfer Radiation. Sci. Rep. 2018, 8, 3664. [Google Scholar] [CrossRef]
  75. Cesaire, M.; Ghosh, U.; Austry, J.B.; Muller, E.; Cammarata, F.P.; Guillamin, M.; Caruso, M.; Castera, L.; Petringa, G.; Cirrone, G.A.P.; et al. Sensitization of chondrosarcoma cells with PARP inhibitor and high-LET radiation. J. Bone Oncol. 2019, 17, 100246. [Google Scholar] [CrossRef]
  76. Buisson, R.; Niraj, J.; Rodrigue, A.; Ho, C.K.; Kreuzer, J.; Foo, T.K.; Hardy, E.J.; Dellaire, G.; Haas, W.; Xia, B.; et al. Coupling of Homologous Recombination and the Checkpoint by ATR. Mol. Cell 2017, 65, 336–346. [Google Scholar] [CrossRef]
  77. Dobbelstein, M.; Sorensen, C.S. Exploiting replicative stress to treat cancer. Nat. Rev. Drug Discov. 2015, 14, 405–423. [Google Scholar] [CrossRef]
  78. Hauge, S.; Eek Mariampillai, A.; Rødland, G.E.; Bay, L.T.E.; Landsverk, H.B.; Syljuåsen, R.G. Expanding roles of cell cycle checkpoint inhibitors in radiation oncology. Int. J. Radiat. Biol. 2023, 99, 941–950. [Google Scholar] [CrossRef]
  79. Morgan, M.A.; Parsels, L.A.; Zhao, L.; Parsels, J.D.; Davis, M.A.; Hassan, M.C.; Arumugarajah, S.; Hylander-Gans, L.; Morosini, D.; Simeone, D.M.; et al. Mechanism of radiosensitization by the Chk1/2 inhibitor AZD7762 involves abrogation of the G2 checkpoint and inhibition of homologous recombinational DNA repair. Cancer Res. 2010, 70, 4972–4981. [Google Scholar] [CrossRef]
  80. Syljuåsen, R.G.; Sørensen, C.S.; Nylandsted, J.; Lukas, C.; Lukas, J.; Bartek, J. Inhibition of Chk1 by CEP-3891 accelerates mitotic nuclear fragmentation in response to ionizing Radiation. Cancer Res. 2004, 64, 9035–9040. [Google Scholar] [CrossRef]
  81. Sørensen, C.S.; Hansen, L.T.; Dziegielewski, J.; Syljuåsen, R.G.; Lundin, C.; Bartek, J.; Helleday, T. The cell-cycle checkpoint kinase Chk1 is required for mammalian homologous recombination repair. Nat. Cell Biol. 2005, 7, 195–201. [Google Scholar] [CrossRef]
  82. Beck, H.; Nahse, V.; Larsen, M.S.; Groth, P.; Clancy, T.; Lees, M.; Jorgensen, M.; Helleday, T.; Syljuasen, R.G.; Sorensen, C.S. Regulators of cyclin-dependent kinases are crucial for maintaining genome integrity in S phase. J. Cell Biol. 2010, 188, 629–638. [Google Scholar] [CrossRef]
  83. Krajewska, M.; Heijink, A.M.; Bisselink, Y.J.; Seinstra, R.I.; Sillje, H.H.; de Vries, E.G.; van Vugt, M.A. Forced activation of Cdk1 via wee1 inhibition impairs homologous recombination. Oncogene 2013, 32, 3001–3008. [Google Scholar] [CrossRef]
  84. Wang, Y.; Li, J.; Booher, R.N.; Kraker, A.; Lawrence, T.; Leopold, W.R.; Sun, Y. Radiosensitization of p53 mutant cells by PD0166285, a novel G2 checkpoint abrogator. Cancer Res. 2001, 61, 8211–8217. [Google Scholar]
  85. Zhou, Q.; Howard, M.E.; Tu, X.; Zhu, Q.; Denbeigh, J.M.; Remmes, N.B.; Herman, M.G.; Beltran, C.J.; Yuan, J.; Greipp, P.T.; et al. Inhibition of ATM Induces Hypersensitivity to Proton Irradiation by Upregulating Toxic End Joining. Cancer Res. 2021, 81, 3333–3346. [Google Scholar] [CrossRef]
  86. Fujisawa, H.; Nakajima, N.I.; Sunada, S.; Lee, Y.; Hirakawa, H.; Yajima, H.; Fujimori, A.; Uesaka, M.; Okayasu, R. VE-821, an ATR inhibitor, causes radiosensitization in human tumor cells irradiated with high LET radiation. Radiat. Oncol. 2015, 10, 175. [Google Scholar] [CrossRef]
  87. Lohberger, B.; Glanzer, D.; Eck, N.; Stasny, K.; Falkner, A.; Leithner, A.; Georg, D. The ATR Inhibitor VE-821 Enhances the Radiosensitivity and Suppresses DNA Repair Mechanisms of Human Chondrosarcoma Cells. Int. J. Mol. Sci. 2023, 24, 2315. [Google Scholar] [CrossRef]
  88. Vitti, E.T.; Kacperek, A.; Parsons, J.L. Targeting DNA Double-Strand Break Repair Enhances Radiosensitivity of HPV-Positive and HPV-Negative Head and Neck Squamous Cell Carcinoma to Photons and Protons. Cancers 2020, 12, 1490. [Google Scholar] [CrossRef]
  89. Bright, S.J.; Flint, D.B.; Martinus, D.K.J.; Turner, B.X.; Manandhar, M.; Ben Kacem, M.; McFadden, C.H.; Yap, T.A.; Shaitelman, S.F.; Sawakuchi, G.O. Targeted Inhibition of DNA-PKcs, ATM, ATR, PARP, and Rad51 Modulate Response to X rays and Protons. Radiat. Res. 2022, 198, 336–346. [Google Scholar] [CrossRef]
  90. Choi, C.; Cho, W.K.; Park, S.; Shin, S.W.; Park, W.; Kim, H.; Choi, D.H. Checkpoint Kinase 1 (CHK1) Inhibition Enhances the Sensitivity of Triple-Negative Breast Cancer Cells to Proton Irradiation via Rad51 Downregulation. Int. J. Mol. Sci. 2020, 21, 2691. [Google Scholar] [CrossRef]
  91. Gorte, J.; Beyreuther, E.; Danen, E.H.J.; Cordes, N. Comparative Proton and Photon Irradiation Combined with Pharmacological Inhibitors in 3D Pancreatic Cancer Cultures. Cancers 2020, 12, 3216. [Google Scholar] [CrossRef]
  92. Liu, B.; Chen, W.; Li, H.; Li, F.; Jin, X.; Li, Q. Radiosensitization of NSCLC cells to X-rays and carbon ions by the CHK1/CHK2 inhibitor AZD7762, Honokiol and Tunicamycin. Radiat. Environ. Biophys. 2020, 59, 723–732. [Google Scholar] [CrossRef]
  93. Bertrand, G.; Maalouf, M.; Boivin, A.; Battiston-Montagne, P.; Beuve, M.; Levy, A.; Jalade, P.; Fournier, C.; Ardail, D.; Magne, N.; et al. Targeting head and neck cancer stem cells to overcome resistance to photon and carbon ion radiation. Stem Cell Rev. Rep. 2014, 10, 114–126. [Google Scholar] [CrossRef]
  94. Ma, H.; Takahashi, A.; Sejimo, Y.; Adachi, A.; Kubo, N.; Isono, M.; Yoshida, Y.; Kanai, T.; Ohno, T.; Nakano, T. Targeting of Carbon Ion-Induced G2 Checkpoint Activation in Lung Cancer Cells Using Wee-1 Inhibitor MK-1775. Radiat. Res. 2015, 184, 660–669. [Google Scholar] [CrossRef]
  95. Jette, N.; Lees-Miller, S.P. The DNA-dependent protein kinase: A multifunctional protein kinase with roles in DNA double strand break repair and mitosis. Prog. Biophys. Mol. Biol. 2015, 117, 194–205. [Google Scholar] [CrossRef]
  96. Gilley, D.; Tanaka, H.; Hande, M.P.; Kurimasa, A.; Li, G.C.; Oshimura, M.; Chen, D.J. DNA-PKcs is Critical for Telomere Capping. Proc. Natl. Acad. Sci. USA 2001, 98, 15084–15088. [Google Scholar] [CrossRef]
  97. Yang, L.; Liu, Y.; Sun, C.; Yang, X.; Yang, Z.; Ran, J.; Zhang, Q.; Zhang, H.; Wang, X.; Wang, X. Inhibition of DNA-PKcs enhances radiosensitivity and increases the levels of ATM and ATR in NSCLC cells exposed to carbon ion irradiation. Oncol. Lett. 2015, 10, 2856–2864. [Google Scholar] [CrossRef]
  98. Meerz, A.; Deville, S.S.; Muller, J.; Cordes, N. Comparative Therapeutic Exploitability of Acute Adaptation Mechanisms to Photon and Proton Irradiation in 3D Head and Neck Squamous Cell Carcinoma Cell Cultures. Cancers 2021, 13, 1190. [Google Scholar] [CrossRef]
  99. Klein, C.; Dokic, I.; Mairani, A.; Mein, S.; Brons, S.; Haring, P.; Haberer, T.; Jakel, O.; Zimmermann, A.; Zenke, F.; et al. Overcoming hypoxia-induced tumor radioresistance in non-small cell lung cancer by targeting DNA-dependent protein kinase in combination with carbon ion irradiation. Radiat. Oncol. 2017, 12, 208. [Google Scholar] [CrossRef]
  100. Zhou, X.; Zhang, X.; Xie, Y.; Tanaka, K.; Wang, B.; Zhang, H. DNA-PKcs inhibition sensitizes cancer cells to carbon-ion irradiation via telomere capping disruption. PLoS ONE 2013, 8, e72641. [Google Scholar] [CrossRef]
  101. Sunada, S.; Kanai, H.; Lee, Y.; Yasuda, T.; Hirakawa, H.; Liu, C.; Fujimori, A.; Uesaka, M.; Okayasu, R. Nontoxic concentration of DNA-PK inhibitor NU7441 radio-sensitizes lung tumor cells with little effect on double strand break repair. Cancer Sci. 2016, 107, 1250–1255. [Google Scholar] [CrossRef]
  102. Lee, J.H.; Paull, T.T. Cellular functions of the protein kinase ATM and their relevance to human disease. Nat. Rev. Mol. Cell Biol. 2021, 22, 796–814. [Google Scholar] [CrossRef]
  103. Shibata, A.; Jeggo, P.A. ATM’s Role in the Repair of DNA Double-Strand Breaks. Genes 2021, 12, 1370. [Google Scholar] [CrossRef]
  104. Pape, K.; Lossner, A.J.; William, D.; Czempiel, T.; Beyreuther, E.; Klimova, A.; Lehmann, C.; Schmache, T.; Merker, S.R.; Naumann, M.; et al. Sensitization of Patient-Derived Colorectal Cancer Organoids to Photon and Proton Radiation by Targeting DNA Damage Response Mechanisms. Cancers 2022, 14, 4984. [Google Scholar] [CrossRef]
  105. Hussain, R.N.; Coupland, S.E.; Khzouz, J.; Kalirai, H.; Parsons, J.L. Inhibition of ATM Increases the Radiosensitivity of Uveal Melanoma Cells to Photons and Protons. Cancers 2020, 12, 1388. [Google Scholar] [CrossRef]
  106. Xue, L.; Yu, D.; Furusawa, Y.; Cao, J.; Okayasu, R.; Fan, S. ATM-dependent hyper-radiosensitivity in mammalian cells irradiated by heavy ions. Int. J. Radiat. Oncol. Biol. Phys. 2009, 75, 235–243. [Google Scholar] [CrossRef]
  107. Durante, M.; Formenti, S.C. Radiation-Induced Chromosomal Aberrations and Immunotherapy: Micronuclei, Cytosolic DNA, and Interferon-Production Pathway. Front. Oncol. 2018, 8, 192. [Google Scholar] [CrossRef]
  108. MacDonald, K.M.; Benguerfi, S.; Harding, S.M. Alerting the immune system to DNA damage: Micronuclei as mediators. Essays Biochem. 2020, 64, 753–764. [Google Scholar] [CrossRef]
  109. Dillon, M.T.; Barker, H.E.; Pedersen, M.; Hafsi, H.; Bhide, S.A.; Newbold, K.L.; Nutting, C.M.; McLaughlin, M.; Harrington, K.J. Radiosensitization by the ATR Inhibitor AZD6738 through Generation of Acentric Micronuclei. Mol. Cancer Ther. 2017, 16, 25–34. [Google Scholar] [CrossRef]
  110. Chen, J.; Harding, S.M.; Natesan, R.; Tian, L.; Benci, J.L.; Li, W.; Minn, A.J.; Asangani, I.A.; Greenberg, R.A. Cell Cycle Checkpoints Cooperate to Suppress DNA- and RNA-Associated Molecular Pattern Recognition and Anti-Tumor Immune Responses. Cell Rep. 2020, 32, 108080. [Google Scholar] [CrossRef]
  111. Eek Mariampillai, A.; Hauge, S.; Øynebraten, I.; Rødland, G.E.; Corthay, A.; Syljuåsen, R.G. Caspase activation counteracts interferon signaling after G2 checkpoint abrogation by ATR inhibition in irradiated human cancer cells. Front. Oncol. 2022, 12, 981332. [Google Scholar] [CrossRef]
  112. Chao, H.H.; Karagounis, I.V.; Thomas, C.; François, N.B.; Facciabene, A.; Koumenis, C.; Maity, A. Combination of CHEK1/2 inhibition and ionizing radiation results in abscopal tumor response through increased micronuclei formation. Oncogene 2020, 39, 4344–4357. [Google Scholar] [CrossRef]
  113. Chen, G.; Zheng, D.; Zhou, Y.; Du, S.; Zeng, Z. Olaparib enhances radiation-induced systemic anti-tumor effects via activating STING-chemokine signaling in hepatocellular carcinoma. Cancer Lett. 2024, 582, 216507. [Google Scholar] [CrossRef]
  114. Harding, S.M.; Benci, J.L.; Irianto, J.; Discher, D.E.; Minn, A.J.; Greenberg, R.A. Mitotic progression following DNA damage enables pattern recognition within micronuclei. Nature 2017, 548, 466–470. [Google Scholar] [CrossRef]
  115. Zhang, N.; Gao, Y.; Zeng, Z.; Luo, Y.; Jiang, X.; Zhang, J.; Li, J.; Zhang, J.; Gong, Y.; Xie, C. PARP inhibitor niraparib as a radiosensitizer promotes antitumor immunity of radiotherapy in EGFR-mutated non-small cell lung cancer. Clin. Transl. Oncol. 2021, 23, 1827–1837. [Google Scholar] [CrossRef]
  116. Feng, X.; Tubbs, A.; Zhang, C.; Tang, M.; Sridharan, S.; Wang, C.; Jiang, D.; Su, D.; Zhang, H.; Chen, Z.; et al. ATR inhibition potentiates ionizing radiation-induced interferon response via cytosolic nucleic acid-sensing pathways. EMBO J. 2020, 39, e104036. [Google Scholar] [CrossRef]
  117. Zhang, Q.; Jiang, L.; Wang, W.; Huber, A.K.; Valvo, V.M.; Jungles, K.M.; Holcomb, E.A.; Pearson, A.N.; The, S.; Wang, Z.; et al. Potentiating the radiation-induced type I interferon antitumoral immune response by ATM inhibition in pancreatic cancer. JCI Insight 2024, 9, e168824. [Google Scholar] [CrossRef]
  118. Du, J.; Kageyama, S.I.; Yamashita, R.; Tanaka, K.; Okumura, M.; Motegi, A.; Hojo, H.; Nakamura, M.; Hirata, H.; Sunakawa, H.; et al. Transposable elements potentiate radiotherapy-induced cellular immune reactions via RIG-I-mediated virus-sensing pathways. Commun. Biol. 2023, 6, 818. [Google Scholar] [CrossRef]
  119. Lee, A.K.; Pan, D.; Bao, X.; Hu, M.; Li, F.; Li, C.Y. Endogenous Retrovirus Activation as a Key Mechanism of Anti-Tumor Immune Response in Radiotherapy. Radiat. Res. 2020, 193, 305–317. [Google Scholar] [CrossRef]
  120. Pan, D.; Bao, X.; Hu, M.; Jiao, M.; Li, F.; Li, C.Y. SETDB1 Restrains Endogenous Retrovirus Expression and Antitumor Immunity during Radiotherapy. Cancer Res. 2022, 82, 2748–2760. [Google Scholar] [CrossRef]
  121. Tigano, M.; Vargas, D.C.; Tremblay-Belzile, S.; Fu, Y.; Sfeir, A. Nuclear sensing of breaks in mitochondrial DNA enhances immune surveillance. Nature 2021, 591, 477–481. [Google Scholar] [CrossRef]
  122. Hu, M.; Zhou, M.; Bao, X.; Pan, D.; Jiao, M.; Liu, X.; Li, F.; Li, C.Y. ATM inhibition enhances cancer immunotherapy by promoting mtDNA leakage and cGAS/STING activation. J. Clin. Investig. 2021, 131, e139333. [Google Scholar] [CrossRef]
  123. Galluzzi, L.; Guilbaud, E.; Schmidt, D.; Kroemer, G.; Marincola, F.M. Targeting immunogenic cell stress and death for cancer therapy. Nat. Rev. Drug Discov. 2024, 23, 445–460. [Google Scholar] [CrossRef]
  124. Eek Mariampillai, A.; Hauge, S.; Kongsrud, K.; Syljuåsen, R.G. Immunogenic cell death after combined treatment with radiation and ATR inhibitors is dually regulated by apoptotic caspases. Front. Immunol. 2023, 14, 1138920. [Google Scholar] [CrossRef]
  125. Zhang, N.; Gao, Y.; Huang, Z.; Dai, P.; Luo, Y.; Wu, Q.; Jiang, X.; Sun, W.; Zhang, J.; Han, L.; et al. PARP inhibitor plus radiotherapy reshapes an inflamed tumor microenvironment that sensitizes small cell lung cancer to the anti-PD-1 immunotherapy. Cancer Lett. 2022, 545, 215852. [Google Scholar] [CrossRef]
  126. Hsieh, R.C.; Krishnan, S.; Wu, R.C.; Boda, A.R.; Liu, A.; Winkler, M.; Hsu, W.H.; Lin, S.H.; Hung, M.C.; Chan, L.C.; et al. ATR-mediated CD47 and PD-L1 up-regulation restricts radiotherapy-induced immune priming and abscopal responses in colorectal cancer. Sci. Immunol. 2022, 7, eabl9330. [Google Scholar] [CrossRef]
  127. Sato, H.; Niimi, A.; Yasuhara, T.; Permata, T.B.M.; Hagiwara, Y.; Isono, M.; Nuryadi, E.; Sekine, R.; Oike, T.; Kakoti, S.; et al. DNA double-strand break repair pathway regulates PD-L1 expression in cancer cells. Nat. Commun. 2017, 8, 1751. [Google Scholar] [CrossRef]
  128. Sun, L.L.; Yang, R.Y.; Li, C.W.; Chen, M.K.; Shao, B.; Hsu, J.M.; Chan, L.C.; Yang, Y.; Hsu, J.L.; Lai, Y.J.; et al. Inhibition of ATR downregulates PD-L1 and sensitizes tumor cells to T cell-mediated killing. Am. J. Cancer Res. 2018, 8, 1307–1316. [Google Scholar]
  129. Marcus, D.; Lieverse, R.I.Y.; Klein, C.; Abdollahi, A.; Lambin, P.; Dubois, L.J.; Yaromina, A. Charged Particle and Conventional Radiotherapy: Current Implications as Partner for Immunotherapy. Cancers 2021, 13, 1468. [Google Scholar] [CrossRef]
  130. Ahmed, S.U.; Carruthers, R.; Gilmour, L.; Yildirim, S.; Watts, C.; Chalmers, A.J. Selective Inhibition of Parallel DNA Damage Response Pathways Optimizes Radiosensitization of Glioblastoma Stem-like Cells. Cancer Res. 2015, 75, 4416–4428. [Google Scholar] [CrossRef]
  131. Cinat, D.; van der Wal, R.; Baanstra, M.; Soto-Gamez, A.; Jellema-de Bruin, A.L.; van Goethem, M.-J.; van Vugt, M.A.T.M.; Barazzuol, L.; Coppes, R.P. Derepression of transposable elements enhances interferon beta signaling and stem/progenitor cell activity after proton irradiation. bioRxiv 2024. [Google Scholar] [CrossRef]
  132. Timm, S.; Lorat, Y.; Jakob, B.; Taucher-Scholz, G.; Rube, C.E. Clustered DNA damage concentrated in particle trajectories causes persistent large-scale rearrangements in chromatin architecture. Radiother. Oncol. 2018, 129, 600–610. [Google Scholar] [CrossRef]
  133. Durante, M.; Formenti, S. Harnessing radiation to improve immunotherapy: Better with particles? Br. J. Radiol. 2020, 93, 20190224. [Google Scholar] [CrossRef]
  134. Zhou, H.; Tu, C.; Yang, P.; Li, J.; Kepp, O.; Li, H.; Zhang, L.; Zhang, L.; Zhao, Y.; Zhang, T.; et al. Carbon ion radiotherapy triggers immunogenic cell death and sensitizes melanoma to anti-PD-1 therapy in mice. Oncoimmunology 2022, 11, 2057892. [Google Scholar] [CrossRef]
  135. Onishi, M.; Okonogi, N.; Oike, T.; Yoshimoto, Y.; Sato, H.; Suzuki, Y.; Kamada, T.; Nakano, T. High linear energy transfer carbon-ion irradiation increases the release of the immune mediator high mobility group box 1 from human cancer cells. J. Radiat. Res. 2018, 59, 541–546. [Google Scholar] [CrossRef]
  136. Chen, M.F.; Chen, P.T.; Hsieh, C.C.; Wang, C.C. Effect of Proton Therapy on Tumor Cell Killing and Immune Microenvironment for Hepatocellular Carcinoma. Cells 2023, 12, 332. [Google Scholar] [CrossRef]
  137. Iijima, M.; Okonogi, N.; Nakajima, N.I.; Morokoshi, Y.; Kanda, H.; Yamada, T.; Kobayashi, Y.; Banno, K.; Wakatsuki, M.; Yamada, S.; et al. Significance of PD-L1 expression in carbon-ion radiotherapy for uterine cervical adeno/adenosquamous carcinoma. J. Gynecol. Oncol. 2020, 31, e19. [Google Scholar] [CrossRef]
  138. Permata, T.B.M.; Sato, H.; Gu, W.; Kakoti, S.; Uchihara, Y.; Yoshimatsu, Y.; Sato, I.; Kato, R.; Yamauchi, M.; Suzuki, K.; et al. High linear energy transfer carbon-ion irradiation upregulates PD-L1 expression more significantly than X-rays in human osteosarcoma U2OS cells. J. Radiat. Res. 2021, 62, 773–781. [Google Scholar] [CrossRef]
  139. Li, Y.; Fan, X.; Yu, Q.; Zhai, H.; Mo, M.; Sun, J.; Mi, J.; Lu, R.; Mao, J.; Chen, J.; et al. Proton and Carbon Ion Radiation Therapy Decreased Severe Lymphopenia by Reducing Thoracic Vertebra and Aortic Doses in Non-Small Cell Lung Cancer Versus Intensity Modulated Radiation Therapy. Int. J. Radiat. Oncol. Biol. Phys. 2023, 116, 579–589. [Google Scholar] [CrossRef]
  140. Wang, X.; van Rossum, P.S.N.; Chu, Y.; Hobbs, B.P.; Grassberger, C.; Hong, T.S.; Liao, Z.; Yang, J.; Zhang, X.; Netherton, T.; et al. Severe Lymphopenia During Chemoradiation Therapy for Esophageal Cancer: Comprehensive Analysis of Randomized Phase 2B Trial of Proton Beam Therapy Versus Intensity Modulated Radiation Therapy. Int. J. Radiat. Oncol. Biol. Phys. 2024, 118, 368–377. [Google Scholar] [CrossRef]
  141. Durante, M. Kaplan lecture 2023: Lymphopenia in particle therapy. Int J Radiat Biol 2024, 100, 669–677. [Google Scholar] [CrossRef]
  142. Guan, F.; Geng, C.; Ma, D.; Bronk, L.; Kerr, M.; Li, Y.; Gates, D.; Kroger, B.; Sahoo, N.; Titt, U.; et al. RBE Model-Based Biological Dose Optimization for Proton Radiobiology Studies. Int. J. Part. Ther. 2018, 5, 160–171. [Google Scholar] [CrossRef]
Figure 1. Illustration of depth dose and LET distribution of particle beams. (A) Pristine Bragg peak (blue curve) with increasing linear energy transfer (LET; orange curve) at the distal dose fall-off. (B) Spread-out Bragg peak (SOBP; dotted curve) with typical LET profiles for clinical carbon ion and proton beams (orange curves). Examples of DNA damage caused by high- and low-LET irradiation are shown to the right.
Figure 1. Illustration of depth dose and LET distribution of particle beams. (A) Pristine Bragg peak (blue curve) with increasing linear energy transfer (LET; orange curve) at the distal dose fall-off. (B) Spread-out Bragg peak (SOBP; dotted curve) with typical LET profiles for clinical carbon ion and proton beams (orange curves). Examples of DNA damage caused by high- and low-LET irradiation are shown to the right.
Cells 13 01058 g001
Figure 2. DNA repair pathway choice may be different for damage induced by low- versus high-LET irradiation. Consequently, the effects of DNA repair inhibitors on radiation-induced cell death and antitumor immune signaling may be different.
Figure 2. DNA repair pathway choice may be different for damage induced by low- versus high-LET irradiation. Consequently, the effects of DNA repair inhibitors on radiation-induced cell death and antitumor immune signaling may be different.
Cells 13 01058 g002
Figure 3. Potential mechanisms of increased antitumor immune signaling following treatment with DNA repair inhibitors and high-LET radiation. DNA repair inhibitors include inhibitors of cell cycle checkpoints. IFN: interferon; ICD: immunogenic cell death. Small, vertical errors indicate changes in occurrence of proteins.
Figure 3. Potential mechanisms of increased antitumor immune signaling following treatment with DNA repair inhibitors and high-LET radiation. DNA repair inhibitors include inhibitors of cell cycle checkpoints. IFN: interferon; ICD: immunogenic cell death. Small, vertical errors indicate changes in occurrence of proteins.
Cells 13 01058 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rødland, G.E.; Temelie, M.; Eek Mariampillai, A.; Hauge, S.; Gilbert, A.; Chevalier, F.; Savu, D.I.; Syljuåsen, R.G. Potential Benefits of Combining Proton or Carbon Ion Therapy with DNA Damage Repair Inhibitors. Cells 2024, 13, 1058. https://doi.org/10.3390/cells13121058

AMA Style

Rødland GE, Temelie M, Eek Mariampillai A, Hauge S, Gilbert A, Chevalier F, Savu DI, Syljuåsen RG. Potential Benefits of Combining Proton or Carbon Ion Therapy with DNA Damage Repair Inhibitors. Cells. 2024; 13(12):1058. https://doi.org/10.3390/cells13121058

Chicago/Turabian Style

Rødland, Gro Elise, Mihaela Temelie, Adrian Eek Mariampillai, Sissel Hauge, Antoine Gilbert, François Chevalier, Diana I. Savu, and Randi G. Syljuåsen. 2024. "Potential Benefits of Combining Proton or Carbon Ion Therapy with DNA Damage Repair Inhibitors" Cells 13, no. 12: 1058. https://doi.org/10.3390/cells13121058

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

Article Metrics

Back to TopTop