Next Article in Journal
Current Understanding of Microbiomes in Cancer Metastasis
Previous Article in Journal
Immune Checkpoint Inhibitors and the Kidney: A Focus on Diagnosis and Management for Personalised Medicine
Previous Article in Special Issue
Oncologic Impact and Safety of Pre-Operative Radiotherapy in Localized Prostate and Bladder Cancer: A Comprehensive Review from the Cancerology Committee of the Association Française d’Urologie
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Promise of Nanoparticles-Based Radiotherapy in Cancer Treatment

1
Department of Mathematics and Natural Sciences, BRAC University, Dhaka 1212, Bangladesh
2
Department of Biochemistry and Molecular Biology, Jahangirnagar University, Savar, Dhaka 1342, Bangladesh
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(6), 1892; https://doi.org/10.3390/cancers15061892
Submission received: 6 February 2023 / Revised: 13 March 2023 / Accepted: 15 March 2023 / Published: 22 March 2023
(This article belongs to the Special Issue Preoperative Radiotherapy in Cancers)

Abstract

:

Simple Summary

Radiotherapy (RT) is used worldwide as a gold standard treatment approach for cancer management. However, the RT treatment modality contains limitations along with numerous side effects. Nanoparticles (NPs) have unique properties that can be utilized in the field of cancer treatment. Therefore, the combination of NPs with RT opens a new arena in cancer treatment. Their synergistic effect strengthens ionizing radiation sensitivity and allows for tumor-selective treatment while reducing side effects. More importantly, NP-based RT offers greater control over RT alone and has shown higher selectivity. In addition, the combined treatment also helps to overcome radioresistance and drug-resistance phenomena. The main mechanism through which NP-based RT destroys cancer cells includes production of ROS, which damage DNA, inhibiting the DNA-repair system, perturbing the cell cycle, and controlling the tumor microenvironment. NP-based RT has been reported to destroy cancer stem cells and has shown good results in clinical trials. Moreover, the addition of phototherapy to NP-based RT reduces the limitations of phototherapy and has shown excellent cancer cell-killing potentiality.

Abstract

Radiation has been utilized for a long time for the treatment of cancer patients. However, radiotherapy (RT) has many constraints, among which non-selectivity is the primary one. The implementation of nanoparticles (NPs) with RT not only localizes radiation in targeted tissue but also provides significant tumoricidal effect(s) compared to radiation alone. NPs can be functionalized with both biomolecules and therapeutic agents, and their combination significantly reduces the side effects of RT. NP-based RT destroys cancer cells through multiple mechanisms, including ROS generation, which in turn damages DNA and other cellular organelles, inhibiting of the DNA double-strand damage-repair system, obstructing of the cell cycle, regulating of the tumor microenvironment, and killing of cancer stem cells. Furthermore, such combined treatments overcome radioresistance and drug resistance to chemotherapy. Additionally, NP-based RT in combined treatments have shown synergistic therapeutic benefit(s) and enhanced the therapeutic window. Furthermore, a combination of phototherapy, i.e., photodynamic therapy and photothermal therapy with NP-based RT, not only reduces phototoxicity but also offers excellent therapeutic benefits. Moreover, using NPs with RT has shown promise in cancer treatment and shown excellent therapeutic outcomes in clinical trials. Therefore, extensive research in this field will pave the way toward improved RT in cancer treatment.

1. Introduction

Cancer is a great threat to global public health [1]. It is one of the main causes of significant morbidity and global deaths per year [2]. In 2022, approximately 1,918,030 cancer cases and 609,360 cancer related deaths were reported [3], and 26 million new cancer cases are projected to occur by 2030 [4]. Cancer develops by uncontrolled proliferation of cells due to pathophysiological alterations of the inherent cell division process that later disseminate to different tissues [5,6]. Radiotherapy (RT) is actively used to treat primary, as well as terminal, tumors [7]. This course of treatment damages cancer cells upon exposure to ionization radiation [8]. A beam of high-energy ionization radiation destroys intracellular components of tumors, killing cancer cells [9]. Radiation treatment is effectual for more than half of cancer patients [10,11,12] and is applied in two-thirds of cancer treatment regimens in Western countries as a significant treatment modality for locoregional tumors [13]. Approximately 30–50% of all cancer patients receive RT, either alone or as an adjuvant treatment [14].
However, RT has limitations that includes dose heterogeneity, intrinsic as well as acquired resistance to RT, and tumor recurrence [15,16]. In addition, delivering sufficient tumoricidal radiation doses induces toxicity in both cancer and nearby non-cancerous cells since both cells have similar mass energy absorption tendencies [17]. Therefore, RT is restricted by the highest tolerated dose to the surrounding normal tissues [13]. Moreover, hypoxic tumors resist radiation treatment [18]. RT also exhibits side effects (acute, consequential, or late complications) causing radiation toxicity in the skin, mucosa, liver, lungs, kidneys, and heart [19,20,21,22,23].
The advent of nanoparticles (NPs) has assisted in the evolution of traditional RT from a “one-size-fits-all” concept to tailored and dynamic treatment modalities [13]. High surface-to-volume ratios, increased cellular uptake, and adjustability make NPs a suitable choice for tumor-targeted and less toxic treatment approaches [24,25,26,27,28,29]. NPs that are prepared from high Z materials act as radiosensitizers, receiving external beams of ionizing radiation [30,31]. Additionally, NPs can be modified by target molecules, and radiosensitizers are able to penetrate and accumulate in tumors to achieve tumor selectivity, as well as enhanced therapeutic doses [32,33,34,35]. Such NP-based radiosensitizing agents draw contrasts between cancer and non-cancerous cells owing to variability in their mass absorption coefficients [17]. Nanostructured radiosensitizers intensify radiation to increase the local dose and overcome hypoxia and rapid proliferation [36,37]. The combination of phototherapy with NP-based RT further improvises cancer management by increasing significant anticancer activity while reducing the limitations of both photothermal and photodynamic treatment [38].
In this review, we highlight the potentiality of NP-based RT to overcome the limitations of conventional RT, as well as combined treatment of RT with chemotherapy (CT). We have describe the robust and significant tumoricidal activity of the NP-based radiotherapeutic approaches compared to radiation or NP treatment only in cancer, and we explore its effects on cancer stem cells and in clinical trials. In addition, NP-based RT’s potential to overcome radioresistance and drug resistance is discussed. Moreover, we indicate some crucial limitations of NP-based RT and address the future prospects of this treatment modality.

2. Combinational Use of RT and Chemotherapy

Radiotherapy utilizes high-energy beams to destroy cancer cells and thereby shrink tumors [39]. On the other hand, chemotherapy (CT) uses cytotoxic drugs that are capable of killing cancer cells and inhibit cancer cell growth [40]. These conventional treatments have limitations, such as insufficient therapeutic properties and side effects [41]. In this regard, combination of RT with CT offers an improved therapeutic effect compared to using a single approach. CT is widely used in lung cancer [42], esophageal cancer [43], rectal cancer [44], and hepatocellular carcinoma [45]. However, toxic side effects and a lack of selectivity and synergy between RT and CT are key problems in chemotherapeutic treatment. In addition, severe side effects of Pt-based anticancer drugs have restricted their clinical application [41,46]. On the other hand, RT is often off-target and damages surrounding healthy cells, and it is difficult to obtain optimum radiation [47]. The application of nanostructured radiosensitizers in RT has attained great attention recently. Using high Z materials as nano radiosensitizers enhances RT owing to their Compton scattering and photoelectric effects [48]. Moreover, introducing platinum-based anticancer drugs, such as cisplatin (II), oxaliplatin (II), and carboplatin (II), as radiosensitizers has yielded more effective RT strategies [49].
A key advantage to combined treatment is the prospect of achieving a greater organ-preservation rate. Another advantage is its independent cell-destroying effect. RT aims to control the primary tumor, whereas CT eliminates distant metastasis. The combined treatment modality is also advocated based on clinical trial results. Phase II trials have reported convincing results of combined treatments [50]. However, the combined approach has some challenges to overcome. Several parameters, such as the dose, duration, administration sequence, should be optimized properly [51]. Therefore, further research is needed to explore these limitations.

3. How Radiation Reacts with Radiosensitizers of High Z Materials

The bombarding of ionizing radiation with NPs gives rise to several outcomes, including photoelectric effects, the Compton electron effect, and Auger electron effects (Figure 1) [40]. The radiation energy is imparted to the electrons of the NPs’ atoms, causing the ejection of electrons from their orbits [41]. Such electronic ejection occurs with a kinetic energy that is equivalent to the radiation wave energy minus the binding energy of the electrons and assesses the electron range in tissue [42].
Photoelectric effects occur when low-energy photons interacts with materials (<60 keV). The photon energy is exclusively absorbed by inner orbital electrons and is ejected from its orbit. This phenomenon causes the electrons of outer orbits to shift to inner orbits and empty space. Thus, liberated fluorescence photons with specific wavelengths depend on the difference between the energy of two orbits, called secondary radiation (Figure 1). Later, Auger electrons are emitted when outer orbit electrons fill the empty space of the inner orbit due to photoelectric effects. This process relinquishes energy to outer orbits electrons, leading to ejection of electrons from higher orbits (Figure 1) [17]. The Auger electrons have high linear energy transfer properties and hence could be extremely injurious to cells [43]. The probability of photoelectric effects occurring is assessed by the formula (Z/E)3, where E = the incoming photon’s energy, and Z = the absorber molecule’s atomic number. Thus, the possibility of photoelectric effects is enhanced with increased absorber atoms, but it decreases with increased energy of incident photons. Photoelectric effects contribute more to radiation interaction with high Z metal NPs than to absorption in soft tissue. Therefore, photoelectrons, secondary photons, and Auger electrons released from high-Z metal NPs enhance localized doses, along with focal ionization of nearby cells via photoelectric effects. Since photoelectric effects decrease with increased energy of photons, most of the nanoparticles combining with radiation treatment use keV photons to optimize the radiosensitization and enhance the local dose by 10–150 times [44,45,46].
The Compton interaction dominates within 25 keV–25 MeV of photon energy. Since most RT is performed at an energy level of 6–20 MeV, this effect is the most common interaction between incident photons and cancer tissue. In the case of the Compton effect, incident photons strike weakly bound outer orbit electrons and donate part of their energy to the electrons, stimulating electrons to leave the outer orbit. Concurrently, the photons become scattered after giving part of their energy and further interacting with other atoms (Figure 1). Afterward, the emitted electrons continue to ionize adjacent tissues. The possibility of a Compton interaction depends inversely on the incoming photon’s energy but is not dependent on the material’s atomic number. Therefore, high-Z metal NPs do not have a substantial role in the Compton effect [17,40,46,47].

4. Biological Response of NPs-Based RT

Complementing nanotherapeutics with ionizing radiation exhibits an enhanced biological response in cancer management through several approaches. The major mechanisms include inhibiting DNA-repair processes, producing reactive oxygen species (ROS), which damage DNA or other biomolecules by oxidation, inhibiting tumor metastasis by controlling the tumor microenvironment (TME), and arresting the cell cycle (Figure 2).

4.1. DNA Damage

Ionizing with X-rays, γ-rays, or proton radiation itself causes spontaneous double strand breaks (DSBs) in DNA by breaking atomic and molecular bonds [48,49,50], and un-repairing of DSB leads to genetic instability, cell division termination, and reduced proliferation and consequently to death [51,52]. However, such DSBs are repaired by the cellular DNA damage response (DDR) [53]. Three main proteins from the phosphatidylinositol 3-kinase-related kinase family, namely ataxia-telangiectasia mutated (ATM), ATM and Rad3 related (ATR), and DNA-dependent protein kinase (DNA-PK), are involved in identifying and repairing DNA DSBs [54]. Three different types of sensor protein complexes are responsible for the recruitment, as well as activation, of these three proteins of the PI3K family at damaged DNA sites, i.e., MRE11/RAD50/NBS1 (MRN) for ATM, RPA/ATRIP for ATR, and KU70–KU80/86 for DNA-PK [55].
The main goal of DDR machinery is to delay the progression of the cell cycle and fix the damage [56]. When cancer cells are exposed to IR, they undergo transient cell cycle arrest to repair DSBs either by non-homologous end joining (NHEJ) throughout the entire cell cycle or by homologous recombination (HR) during the S and G2 phase [57]. ATM is primarily responsible for activating the HR pathway. Other proteins, including breast and ovarian susceptibility protein (Brca2), Rad51, and X-ray repair cross complementing protein 2 (XRCC2), are also involved in the HR pathway [58]. On the other hand, DNA-PK mainly regulates the NHEJ pathway in association with DNA ligase IV and X-ray repair cross complementing protein 4 [59,60]. Another pathway involves both ATM and ATR. Here, MRN sensor protein complex senses damage sites and activates ATM [61]. Autophosphorylation of ATM kinase sends signals to transducers such as checkpoint kinase 2 (Chk2) and the transcription factor p53. p53 controls the expression of p21, which interacts with cyclin-dependent kinase (CDK) complex and arrests the G1 phase of the cell cycle [62]. Modification of chromatin also occurs together with the process, and then the DNA repair process is initiated. However, mutation of the p53 makes the G1 checkpoint defective in most of the cancer cells. Hence, the G2 checkpoint plays a crucial role for surviving cancer cells. RPA/ATRIP sensor protein complex recognizes and activates ATR. Here, ATR phosphorylates checkpoint kinase 1 (Chk1), which degrades cell division cycle 25A (CDC25A) through further phosphorylation and slows the progression of DNA replication during the S phase. The ATM-Chk2 and ATR-Chk1 signaling pathway acts together with DNA-PK phosphorylate p53, which controls genes required for DNA repair, arresting the cell cycle, and apoptosis [55,61,63].
For this purpose, DNA double-strand repair inhibition (DSBRI) appears to a promising strategy for RT [64]. However, it is a challenge to achieve tumor selective DSBRI-based radiotherapeutic treatment since such approaches often sensitize normal cells [65]. Interestingly, introduction of NP-based radiotherapeutic approaches mediates tumor-specific DSBRI owing to their increased permeability and retention effects [66,67]. Zhang et al. developed nano-constructure by combining androgen receptor (AR) with shRNA and folate-targeted H1 nanopolymer (NP AR-shRNA). NP AR-shRNA selectively destroyed prostate cancer cells by mimicking DNA DSBs and activated kinase activity, in turn impeding DNA damage repair signaling pathways (Figure 2) [68]. The presence of γ-H2AX is used as an indicator to detect DNA DSB in higher eukaryotes [69,70]. NP AR-shRNA in combination with IR (4 Gy) increases the expression of γ-H2AX in PC3 and 22Rv1 cells by nearly three-fold compared to IR alone (4 Gy). In vivo experiments showed that irradiation downregulated the expression of AR protein while increasing γ-H2AX expression in 22Rv1 tumors. Additionally, mice bearing PC3 and 22Rv1 cells were exposed to NP AR-shRNA under X-ray (total 9 Gy) and displayed significant tumor reduction compared to only X-ray treatment [68]. Similarly, Yao et al. formulated nanoparticles by conjugating DSB bait (Dbait) with H1 polymer (Dbait@H1 NPs), selectively killing prostate cancer cells upon exposure to radiation by inhibiting DSB repair [71]. Dbait@H1 NPs attached selectively to folate receptor and mimicked DNA DSBs upon release into the nucleus [72,73]. Dbait of Dbait@H1 NPs activates DNA-PK and phosphorylate γ-H2AX. Then, factors associated with DNA damage repair are assembled at the free end of Dbait, preventing them from affecting the DSB sites of the real chromosome and resulting in prolonged defects in DSB repair [74,75,76,77]. Irradiating (4 Gy) PC-3 cells with Dbait@H1 NPs induced γ-H2AX foci numbers three times higher than radiation alone (4 Gy) (Table 1). Moreover, the same radiation dose also increases phosphorylation of DNA-PK and H2AX of both PC-3 and 22Rv1 cells. Moreover, an in vivo study showed that mice carrying both PC-3 and 22Rv1 cells were exposed to 9 Gy of radiation and 60 μg/kg of Dbait@H1 NPs and exhibited 1.67- and 2.5-fold reduced tumor volumes, respectively, compared to only radiation (9 Gy) (Table 1) [71]. Similarly, HeLa cells exposed to AuNPs and irradiation at 4 Gy of 220 kVp and 6 MVp caused enhanced γ-H2AX, suggesting induction of possible DNA DSB [78]. Combined treatment with PEGylated-AuNPs and 4 Gy RT (150 kVp) increased DNA damage 1.7-fold in U251 cells compared to radiation alone [79]. Additionally, Zheng et al. reported that AuNPs at a radiation dose of 6 Gy induced DSB in HepG2 cells [80].
Chen et al. modified AuNPs with bovine serum albumin (BSA@AuNPs), which induced a 2.02-fold increase in γ-H2AX density compared to X-ray radiation only in U87 cells upon exposure to a 3-Gy dose of 160 kVp X-ray. Moreover, treating mice with BSA@AuNPs under X-ray radiation (5 Gy) reduced tumor volume significantly compared to X-ray radiation alone [81]. Similarly, a nanoformulation of KU55933 (NPs@KU55933) impeded the repair process of DSBs and exhibited enhanced tumor volume reduction in vivo. Exposing two types of lung cancer cells (i.e., H460 and A549 cells) to NPs@KU55933 at a radiation dose of 15 Gy showed a remarkable tumor size reduction compared to the X-ray irradiation group (Table 1) [82].

4.2. Reactive Oxygen Species (ROS)

Ionization in combination with NPs brings about indirect necrosis or apoptosis via oxidation of biomolecules, including proteins, lipids, and DNA, along with mitochondrial dysfunction [7,83,84]. High-intensity ionizing radiation generates ROS, including superoxide anion radicals (O2), hydrogen peroxide (H2O2), and hydroxyl radicals (OH), through water radiolysis, and they interact with cellular biomolecules, leading to apoptosis/cellular death [7,85,86,87,88,89], as well as suppressing tumor progression [90]. Therefore, sufficient ROS production is crucial to mediate DNA damage, as well as suppress DNA repair (Figure 2).
Zhao et al. fabricated Gd-bearing polyoxometalates linked with chitosan nanospheres and integrated with hypoxia inducible factor 1α (HIF-1α) siRNA (GdW10@CS_HsiRNA) that showed enhanced radiosensitization in hypoxic tumors. GdW10@CS functions as an external radiosensitizer for depositing ionizing radiation doses and as a nanocarrier of HIF-1α siRNA to stop DNA DSB restoration. Additionally, GdW10@CS annihilates intracellular reduced glutathione (GSH) levels upon exposure to X-ray radiation, leading to overproduction of ROS through W6+-triggered GSH oxidation, thereby facilitating radiotherapeutic efficiency. Irradiating BEL-7402 cells with 6 Gy of X-ray and GdW10@CS (100 μM) produced Compton and Auger electrons, which interact with surrounding H2O or O2 molecules, thus producing 10-fold more ROS, as well as exhausting GSH levels three times more compared to X-ray treatment alone. Furthermore, 20 μL of GdW10@CS along with 10 Gy of X-ray radiation were administered into BALB/c mice bearing BEL-7402 tumors and reduced tumor volumes by nearly five- and eight-fold compared to GdW10@CS_HsiRNA without RT and RT without GdW10@CS_HsiRNA, respectively [91]. Zhan et al. designed a nano-enabled coordination platform with bismuth and cisplatin prodrug (NP@PVP) that improves the efficiency of chemoradiotherapy by X-ray radiation. The bismuth in NP@PVP increases generation of ROS to intensify DNA damage after X-ray irradiation. Treating EMT-6 cells with NP@PVP and X-ray irradiation (5 Gy) generated 3.21-fold more ROS compared to platinum-based drugs along with X-ray irradiation at the same dose. In vivo results showed that mice carrying EMT-6 tumors treated with NP@PVP (2 mg kg−1) under irradiation with X-rays (5 Gy) showed 54.7% inhibited tumor growth compared to X-ray treatment only [92]. Choi et al. synthesized radiation-responsive PEGylated gold nanoparticles containing dihydrohodamine 123 (DHR-123) (RPAuNPs). RPAuNPs absorbs the X-ray energy and transfers it to nearby molecules through electrons, including photoelectrons, Compton election, and Auger electrons, which generate ROS by water radiolysis. Exposure of RPAuNPs (7.75 μg/mL) at to Gy of X-ray radiation increases the fluorescence by 7-fold owing to the local generation of ROS. The same treatment of MDA-MB-231 cells yields similar results, with increased ROS levels in nearby cells. Mice bearing MDA-MB-231 xenografts treated with RPAuNPs at a concentration of 6 μg/100 μL and irradiation of 6 Gy at 225 kVp exhibited 3- to 6-fold higher ROS production compared to the treatment with RPAuNPs without X-ray irradiation [93]. In addition, mitochondria are an important source of cellular ROS production [94]. Tang et al. constructed Gd-doped titanium dioxide nanosensitizer (G@TiO2 NPs), which targets mitochondria for effective RT. G@TiO2 NPs generates ROS effectively since it possesses a large photoelectric cross-section for X-rays [95].

4.3. Tumor Microenvironment (TME)

The tumor microenvironment (TME) consists of various types of cells (endothelial cells, immune cells, fibroblasts, etc.) and extracellular components (extracellular matrix, growth factors, cytokines, etc.) that surrounds tumors and are nourished by blood and the lymphatic vascular network [96,97]. The TME functions critically in the regulation of tumor progression, immune escape, and metastasis [98]. It has significant influence on therapeutic effectivity [99]. Therefore, TME-associated NP-based RT offers potentiality in destroying cancer cells efficiently [100].
ROS are among the key players to alternate the tumor microenvironment (TME) during RT [101]. Among many other features of the TME, hypoxia mostly compromises tumor sensitivity to anticancer drugs, along with reactivity toward free radicals, thus creating hurdles in RT [7]. The anoxic and hypoxic TME of solid tumors compromises the production of ROS, causing poor responses to tumor cells [102,103]. Furthermore, hypoxia=activated hypoxia inducing factor-1 promotes resistance to RT and increases expression of genes involved in angiogenesis and metastasis of tumors [104,105,106]. Therefore, management of the TME is critical in killing cancer cells.
O2 is indispensable during RT since it reacts with DNA breaks to avoid repair of DNA by tumor cells, thus relieving hypoxia and enhancing RT-mediated cell killing [107]. Considering this point, Liu et al. prepared a nanostructure system, PFC@PLGA-RBCM, by enveloping perfluorocarbon (PFC) within poly(d,l-lactide-co-glycolide) (PLGA), which then was coated with a red-blood-cell membrane (RBCM). PFC@PLGA-RBCM NPs contains a PFC core, which is able to dissolve O2 to a great extent, and the RBCM coating contributes to enhanced blood circulation of NPs. PFC@PLGA-RBCM NPs deliver oxygen efficiently to the TME, which helps to relieve tumor hypoxia and enhance the efficacy of RT. Injecting 4T1-tumor-bearing mice with PFC@PLGA-RBCM NPs (200 µL) and exposing them to X-ray radiation (8 Gy) showed enhanced tumor volume reduction by nearly 8- and 2.5-fold compared to PFC@PLGA-RBCM NPs without irradiation and X-ray irradiation only, respectively [108]. Chen et al. fabricated NPs via encapsulating catalase (Cat) by poly(lactic-co-glycolic) acid and hydrophobic imiquimod (Cat@PLGA_R837). Upon irradiation (8 Gy) in CT-26 cell lines, Cat loaded inside Cat@PLGA_R837 decomposes H2O2 to produce O2, thus relieve the hypoxic TME significantly (Table 1) [89]. Moreover, X-ray irradiation induced gold NPs with silica cores (SAuNPs) to exhibit enhanced antitumor effects under a hypoxic environment. At 8 Gy of radiation, SAuNPs caused 20% more cellular death of CT26 cells than in a control group under hypoxic conditions, while only 5% death occurred under normoxic condition. In addition, irradiation with X-rays increased ROS production by 40% in normoxic conditions compared to 20% in hypoxic conditions [109].

4.4. Targeting the Cell Cycle

Targeting the cell cycle is an effective approach in cancer treatment. Different nano formulations in combination with radiation mediate disruption of the cell cycle, leading to apoptosis [85]. Furthermore, cell cycle phases have distinctive effects on radiosensitivity. For instance, the late S-phase is the most radioresistant, while G2 is the most radiosensitive phase [110]. Cells activate cell cycle checkpoints in the G1, S, and G2 phases in response to radiation to repair genomic defects, maintenance integrity, or prevent cell division through activation of cell death mechanisms [111]. The literatures has reported that NPs with radiation arrest mostly the G2/M phase while reducing cells in the G0/G1 phase of the cell cycle (Figure 2) [112,113,114].
Roa et al. constructed gold NPs capped with glucose (Glu-AuNPs), showing improved cell-targeting capacity and excellent radio-sensitization. Glu-AuNPs stimulated activation of cyclin dependent kinase (CDK), leading to an accelerated G0/G1 phase and halting the G2/M phase of the cell cycle by activating CDK1 and CDK2. Irradiating DU-145 cells with 2 Gy of ortho-voltage together with Glu-AuNPs (15 nM) arrested the cell cycle at the G2/M phase and exhibited enhanced growth inhibition by 1.5- to 2-fold compared to X-ray alone. Glu-GNPs inhibited cyclin A’s expression by 42.3%. Cyclin A together with CDK2 form cyclin A–CDK2 complex and initiate the transition of G2/M. Hence, inhibiting cyclin A caused G2/M transition delay [115]. Chen et al. synthesized ultra-small selenium NPs (SeNPs) of 27.5 nm by chemical methods [116]. SeNPs have shown excellent biological activity and low toxicity [117,118]. Upon irradiation, SeNPs arrested the G2/M phase while accelerating the G1/S phase. X-ray irradiation of SeNPs (0.15 μg/mL) at 6 Gy on MCF-7 cells upsurged the G2/M phase proportion by 7.4-fold compared to radiation alone [116]. Xu et al. conjugated gold with glycine (G), arginine (R), and aspartate (D) peptides (Au-G-R-D). While exposed to 6-mV X-rays with a 4-Gy dose, Au-G-R-D (50 μg/mL) arrested the G2/M phase significantly (6.4%) in A375 melanoma cells compared to radiation alone [119].
Table 1. In vivo effects of NP-based RT.
Table 1. In vivo effects of NP-based RT.
Nanoparticle FormulationTest Animal ModelAnimal AgeCancer TypeCell LineTumor Growth Delay Compared to Control GroupConcentrationRadiation DoseObservationsRef.
RTNanoparticleNanoparticle + RTStatistical Significance Method
KU55933Nu/Nu mice6–8 weeksNSCLH460SignificantSignificantSignificantResponse features analysis method500
mg/kg
15 GyImpedes the repair process of DNA double-stranded breaks [82]
A549SignificantSignificantSignificant
Dbait@H1 NPs Male nude mice4–6 weeksPCaPC-3Significant
p = 0.032
NRSignificant
p = 0.004
Two-sample t-tests and analysis
of variance
60 μg/kg9 GySignaling pathways for repairing DNA damage are inhibited by Dbait@H1 NPs, which are competitive inhibitors of DSB. [71]
22Rv1Significant
p = 0.001
NRSignificant
p < 0.001
Cat@PLGA_R837Balb/c miceNRCRCCT-26No appreciable inhibition of tumor growthNRSignificantly suppresses tumor growthOne-way ANOVA using the Tukey’s post-testCat = 0.5 mg/kg; R837 = 0.6 mg/kg 8 GyDecomposes H2O2 to produce O2 in the tumor microenvironment.[120]
IPI549@HMPBalb/c mice6–8 weeksCRCLuc+ CT26 cellsSignificantModerately significantSignificant Student’s two-tailed unpaired t-testMnO2 =
7.5 mg/kg; IPI549 = 1.5 mg/kg
6 GyMyeloid cells are selectively targeted by IPI549@HMP and decompose endogenous H2O2 to O2 upon X-ray irradiation [121]
AuNPsBalb/C miceNRMCEMT-6Non-significant Significantly
delayed tumor growth but did not reduce tumor volume significantly
Significantly decreased tumor growth and tumor volume Wilcoxon’s non-parametric two-sample rank-sum1.35 g/kg30 GyAuNPs, being high-Z nanoparticles, preferentially absorb X-rays and subsequently reduce tumor. [122]
NPs-lncAFAP1-AS1 siRNABALB/c normal mice and nude mice 4–5 weeks BCMDA-MBA-231RSignificantly decreased tumor volume
p < 0.001
Significantly decreased tumor volume
p < 0.01
Significantly decreased tumor volume
p < 0.001
Student’s unpaired two-sided t-test and one-way ANOVA1 nmol siRNA10 GyTumor growth is reduced due to blockage of the Wnt/β Catenin signaling pathway and scavenging intracellular GSH.[123]
Au@Tat-R-EK NPsBALB/c mice6–8 weeks LCLM3Moderately significant Not SignificantSignificantOne-way analysis of
variance (ANOVA) in Origin software
25 mg/kg6 GyAu@Tat-R-EK NPs
respond to overexpressed cathepsin B in the tumor
microenvironment that induces site-specific
enhancement of tumor cell uptake and afterward damages DNA effectively upon X-ray irradiation.
[124]
AuNP@CRZNude mice 10–11 weeksACC of the salivary glandACCImpeded tumor
growth (p < 0.001) but did not significantly reduce tumor volume to less than its size at T0 (p > 0.05)
Significantly decreased tumor
volume (p < 0.001) and caused a reduction in tumor
volume (p < 0.001).
Decreased tumor volume compared to previously (p < 0.001), and it caused significant shrinkage of the tumor to near disappearance
(p = 0.007).
Unpaired two-sided t-test2 mg/kg18 GyTumor-cell repair mechanism is reduced [125]
Lu–Au-
NLS-RGD-anti-VEGF aptamer
Athymic male mice 6–7 weeks MGU87MGTumor size progression was significantly slower (p < 0.05)Tumor size progression was significantly slower (p < 0.05)Tumor size progression was significantly slower (p < 0.05)ANOVA75.1980 GyTumor development is inhibited by halting the formation of new blood vessels [126]
Doc-NPsBALB/c mice 4–5 weeksGCBGC823Not significant in tumor doubling timeNot significant in tumor doubling timeSignificant effect on tumor doubling timet-test5 mg/kg15 GyDoc-NPs cause cell cycle arrest in G2-M phase whereas irradiation leads to ROS generation which induce DNA damage[127]
AuNPsC3H/HeJ mice8–10 weeksSCCSCCVIITime for doubling of tumor volume and survival time increased significantly (p < 0.04)Time for doubling of tumor volume and survival time increased by 23 days and 37%, respectivelyTime for doubling of tumor volume and survival time increased most significantly (p < 0.04)2-sided Z-statistics test1.9 g/kg42 GyAuNPs boost the radiation treatment in a radioresistant squamous cell carcinoma [128]
AuNPsSyngeneic black B6C3f1 miceNRMBTTu-2449No abatement of tumor growth18% long-term survival56% long-term survivalLog-rank (Mantel–Cox) survival
analysis test with 95% CIs using GraphPad Prism®
software
4 g/kg35 GyAuNPs combined with radiation cause about 53% (>1 year) tumor-free survival [129]
NP@PVP + B + Pt BALB/c mice5–6 weeksBCEMT-6Tumor growth was inhibited by 45.4%Tumor growth was inhibited by 51.9%Tumor growth was inhibited by 91.2%Student’s t-test4 mg/kg5 GyThe bismuth in NP@PVP + B + Pt functions as a radiosensitizer and
increases the production of ROS, which damage DNA under X-ray irradiation
[92]
Abbreviations: NP = Nanoparticle; Dbait@H1 NPs = NP-DNA double-strand breaks bait (consisting of polycation polyethylenimine, cross-linked with β-cyclodextrin and conjugated with folic acid); Cat@PLGA_R837 = NP@Poly(lactic-co-glycolic) acid (PLGA) + water-soluble catalase (Cat) + hydrophobic imiquimod (R837), IPI549@HMP = PI3-kinase γ (PI3kγ) inhibitor (IPI549) + PEGylated HMnO2 (HMP); AuNPs = Gold-NPs; NPs-lncAFAP1-AS1 siRNA = Long noncoding actin filament-associated protein 1 antisense RNA1 with Small interfering RNA; Au@Tat-R-EK NPs = Gold + nuclear targeting peptide sequence (GRKKRRQRRRPQ) + peptide sequence (GFLG) + zwitterionic peptide sequence consisting of alternative glutamic Acid (E) and lysine (K); AuNPs@CRZ = AuNPs + Crizotinib; Lu–Au-NLS-RGD-anti-VEGF aptamer = Lutetium177 + Gold + Nuclear Localization Sequence-Arg-Gly-Asp + anti-VEGF aptamer; Doc-NPs = Docetaxel-NPs. NP@PVP + B + Pt = Polyvinylpyrrolidone with bismuth and cisplatin Gy = Gray (unit of ionizing radiation dose in the International System of Units); NR = Not reported; NSCL = Non-small cell lung cancer; PCa = Prostate cancer; CRC = Colorectal cancer; MC = Mammary carcinoma; BC = Breast cancer; LC = Liver cancer; ACC = Adenoid cystic carcinoma; MG = Malignant glioma; GC = Gastric cancer; SCC = Squamous cell carcinoma; MBT = Malignant brain rumor.

5. Uptake and Excretion of NPs during RT

The success of the NP-based RT largely depends on internalization of the nanostructured radiosensitizers into the tumor, along with clearance from the body to avoid any toxicity. Yi et al. explored the possible effect of X-ray irradiation on the uptake and efflux of NPs by cancer cells. It was reported that exposure to X-rays (6 Gy) enhanced the uptake of different nano-structured compounds, including melanin-coated CuS NPs with PEG, AuNPs, silica NPs, and HAS NPs, by 32%, 25%, 33%, and 20%, respectively, by 4T1 cells compared to without X-ray radiation. Furthermore, looking into the probable mechanism, irradiation causes a reduction in the G0/G1 phase while increasing the G2/M phase [130]. Cells show varying levels of endocytosis at different phases of the cell cycle: G2/M > S > G0/G1 [131,132]. Thus, increased G2/M phase after X-ray radiation increased the uptake of NPs by cancers cells [130]. Similarly, Davies et al. prepared nanosized liposomal doxorubicin and reported that RT increased the uptake of doxorubicin by two- to four-fold in the tumor [133]. However, traditional radiosensitizers such as cisplatin exhibited significant tumor uptake and excellent RT sensitization. However, the poor renal clearance caused kidney toxicity [134]. Larger NPs sized 20–100 nm accumulated in the spleen and liver where they lasted for few months, thus mediating long-term toxicity [135,136]. NPs smaller than 5 nm are eliminated shortly in vivo via renal clearance [137]. Zhang et al. fabricated polymer micelles with ultra-small AuNPs (PMG NPs) sized 1.9 nm. PMG NPs improved RT and were removed from the body without causing any toxicity [138].

6. Radiation-Induced Bystander Effects

The biological effects of radiation were believed to be restricted only within targeted tumor areas but have been found to affect adjacent non-targeted tissue surrounding the targeted area [139]. This event is called the radiation-induced bystander effect (RIBE), in which nearby non-irradiated cells act like radiation-exposed cells (Figure 3a) [140,141,142,143]. RIBE arises through signal transmission from irradiated to nearby non-irradiated healthy cells, direct cellular contact, or secreting of soluble factors in the neighboring area [142,144]. These bystander signals may alter genetic expression, exchange sister chromatid, cause genomic stability, damage DNA, reduce cell proliferation, and alter the translation process of non-exposed cells [142,145]. Major bystander signaling molecules are ROS, micro-ribonucleic acid (miRNA), and extracellular oxidized DNA [146,147], which reach surrounding healthy cells via binding to receptors or passive diffusion [147]. For instance, cells died by RT-released cell free chromatin particle (cfchp), which integrates into the bystander cells’ genome and causes DNA damage along with inflammation [148]. However, the abovementioned RIBE can be abrogated by cfchp inactivating agents, such as anti-histone antibody complexed NPs (CNPs). Kirolikar et al. reported that γ-rays (50 Gy) activated RIBE biomarkers, including H2AX, NFκB, IL-6, and caspase-3, in brain cells. Concurrent treatment with CNPs abrogated such RIBE biomarker activation and thus prevented cfchp-induced cell death (Figure 3b) [149]. Likewise, Zainudin et al. irradiated MCF-7 and hFOB 1.19 cells with a photon beam of 6 MV along with bismuth oxide NPs (Bi2O3 NPs). Exposure to ionization radiation plus Bi2O3 NPs did not induce ROS generation or DNA fragmentation due to RIBE. Therefore, the survival rate increased by 3–8% to both bystander cells [150]. Furthermore, Rostami et al. investigated the consequences of RIBE on MCF7 and QUDB cells after treating both cell lines with glucose-covered gold NPs (G-AuNPs) and 2 Gy of 100-kVp X-ray radiation. The results showed 13.2% reduced cell viability and an 11.5% decreased survival fraction in QUDB bystander cells compared to irradiated bystander cells without G-AuNPs. However, MCF7 cells did not show any RIBE upon radiation exposure [144].

7. The Effects of NPs-Based RT in Cancer Stem Cells

Cancer stem cells (CSCs) are common in most cancers; they cause metastases and function as a cancer cell reservoir, causing tumor relapse after CT, radiotherapy, or surgery [151]. CSCs’ ability to proliferate unlimitedly and their resistance to drugs pose great threats in cancer management [152]. CSCs are characteristically resistant to CT due to their quiescence, capacity to repair DNA, and ABC-transporter expression [153]. The self-renewal property allows them to extend tumor cell numbers after chemo- or radiotherapy [154,155]. Conventional treatments are not able to destroy CSCs; therefore, novel treatment strategies are highly demanded.
Fiorillo et al. developed graphene oxide (GO)-based nanostructures that were able to prevent tumor-sphere formation in six different cancer cell lines, including MCF7 for breast cancer, SKOV3 for ovarian cancer, PC3 for prostate cancer, MIA-PaCa-2 for pancreatic cancer, A549 for lung cancer, and U87-MG for brain cancer. They applied tumor sphere assay to measure the formation of tumor spheres to evaluate the effect on GO. The results suggested that GO targets the phenotypic prosperity of CSCs and reduces bona fide CSC numbers by inducing their differentiation and inhibiting their proliferation. More specifically, GO-based treatment inhibited several major signal pathways, including WNT- and Notch-driven signaling, STAT1/3 signaling, and the NRF2-dependent anti-oxidant response, together with inducing the differentiation of CSC, thus decreasing the general stemness [156]. Likewise, Yao et al. designed gastric CSCs targeting carbon nanotubes based on chitosan and loaded with salinomycin with hyaluronic acid (SWCNTs), which selectively eradicate gastric CSCs [157]. Later, Al Faraj et al. modified SWCNTs with CD44 antibodies and showed enhanced targeting of breast CSCs and promise in clinical studies [158].

8. NPs-Based RT Improving Phototherapy

NP-mediated phototherapy, including photothermal therapy (PTT) and photodynamic therapy (PDT), has exhibited promising efficiency in treating superficial and internal tumors [159]. However, the tremendous advantages of phototherapy come with some limitations. PTT has high efficiency in cancer treatment due to its controllable, accurate, and non-invasive properties [160,161,162]. PTT transforms photon energy to hyperthermia through photothermal agents (PTAs) [163]. However, a characteristic drawback of laser attenuation, PTA’s nonuniform distribution, and unwanted phototoxicity to healthy tissue limit PTT [164,165]. On the other hand, PDT employs photosensitizers to generate ROS in response to near infrared irradiation (NIR) [166,167]. Nevertheless, PDT also has limitation since a hypoxic TME hampers its application [168]. However, NP-based RT could be considered a promising adjuvant treatment to complement phototherapy [169]. Low energy and long wavelengths of near infrared light in phototherapy show limited penetration depth, while the high energy and short wavelength of X-rays or γ-rays are free from depth restrictions [170]. Furthermore, their combination reduces the radiation dose while increasing therapeutic efficiency [171].
Treating KB cells with 20 μg/mL of iron oxide coated with gold NPs (Au@Fe2O3 NPs) for 4 h, followed by PTT with irradiation at 808 nm, 6 W/cm2 for 10 min, and 6-MV X-ray irradiation (2 Gy), decreased cell viability at ~40% and ~20% compared to treating them with only RT and only PTT, respectively [172]. Further investigation revealed that such combination treatment in KB cells enhanced the expression of Bax/Bcl2 genes by five- and 6.67-fold, as well as HSP-70 protein by 1.84- and 2-fold, compared to radiation and laser irradiation, respectively [173]. Bax/Bcl2 genes are involved in regulating apoptosis and their overexpression causing cellular apoptosis [174]. In addition, HSP-70 gene overexpression denotes heating, oxidative stress, and inflammation, leading to cellular death [175]. Movahedi et al. reported similar results in KB cells with gold nanorods with folic acid (AuNRs-FA). AuNRs-FA (15 µg/mL), laser irradiation at 808 nm, 2 W/cm2, along with 6-MV X-ray irradiation, reduced cell viability to ~70% compared to single RT or PTT treatment [176]. Gonza’lez-Ruı’z et al. developed lutetium-177 labeled gold NPs with the nuclear localization sequence-Arg-Gly-Asp and an anti-VEGF aptamer nano system that significantly reduced U87MG tumor progression compared to only RT and only PTT under laser irradiation (532 nm, 1.19 W/cm2, 3.5 min) and 89 Gy of X-ray radiation [126]. Another in vivo study showed that PEG-[64Cu]CuS NPs inhibited the growth of anaplastic thyroid carcinoma and provided eight days more survival time than radiation or PTT treatment alone [177]. Liu et al. designed liquid-metal NPs consisting of metronidazole (MN) and GRD peptides (containing glycine-arginine-aspartic acid) linked to polyethylene glycol and polyacrylic acid (GRD-PEG-PAAMN@LM). Administering RGD-PEG-PAAMN@LM within HepG2 tumor-containing mice under NIR irradiation at a wavelength of 808 nm, 2.0 W cm−2 for 5 min and exposing them to 6 Gy for 2 min after 48 h almost abolished tumors after 14 days. Hence, the synergistic effect proved to be better than a single treatment. Regarding deep mechanisms, the authors explained that RGD-PEG-PAAMN@LM targeted ανβ3 integrin over expressive blood vessel walls of tumors and accumulated through endocytosis. RGD-PEG-PAAMN@LM produced excessive thermal energy along with ROS by means of PTT and PDT, which induced tumor apoptosis. At the same time, the MN part entered nucleus to enhance the X-ray radiation-mediated DNA damage [178]. Some other examples of combinational approaches are also described in the literature [179,180].
However, the highest synergistic effect of RT and phototherapy can be obtained when the two modalities have an optimal interval. Safari et al. synthesized NPs with gold and iron oxide cores and coated by alginate (Fe3O4@Au/Alg NPs). Exposure of KB cells to Fe3O4@Au/Alg NPs under 1 W/cm2 of laser irradiation for 5 min arrested the G2/M phase of the cell cycle and 24-h post-treatment with 6 Gy X-ray radiation showed maximum radiosensitivity with 68% apoptosis [181].

9. NPs-Based RT to Overcome Radioresistance and Drug Resistance

Radioresistance (RR) is a major constraint of RT, which leads to the recurrence of cancer [182]. Several improvements in radiotherapeutic approaches have been made out to increase the safety and efficacy while minimizing the RR of tumors. Among different approaches, combinational treatment of radiation with NPs stood as a prominent avenue to overcome RR [16]. Several mechanisms, including overexpression of DNA repair enzyme and anti-apoptotic proteins, contribute to the development of RR [183]. Moreover, progression of a hypoxic TME after irradiation is the one of the primary factors in RR development [184]. However, RR can be reduced by downregulating selective genes such as vascular endothelial growth factor (VEGF) expression that reduce hypoxia and result in a better radiotherapeutic response [185]. Li et al. constructed AuNPs encapsulating recombinant human endostatin (Au-RHES). RHES is a vascular angiogenesis–disrupting agent that normalized transient vascularization in the H22 xenograft model and diminished hypoxia [186]. Additionally, AuNPs tend to inhibit heparin-binding growth factors (HB-GFs) and basic fibroblast growth factor (bFGF), which are involved in tumor metastasis [187]. AuNPs encapsulating siRNA facilitated myelocytomatosis oncogene (c-myc) knockdown in cancerous HeLa cells [188]. In another study, Lee et al. discovered that high expression of low density lipoprotein receptor-related protein-1 (LRP-1) contributed to radio-resistant CRC [189]. LRP-1 plays a vital role in maintaining the TME and regulating cancer invasion due to its association with intracellular signaling and endocytosis of different types of cancer [190,191]. LRP-1 could be considered a marker protein for CRC’s RR [189]. These authors engineered a nano formulation consisting of human serum albumin possessing the LRP-1-binding peptide and B5 for tumor targeting, along with 5-FU and Cy7 (Cy7–B5–HSA–5-FU). Irradiation of Cy7–B5–HSA–5-FU resulted in the reversal of the radio resistance of CRC greatly and inhibited tumor growth significantly with treatment of 2 Gy for five days [189]. Furthermore, Li et al. reported that 1.7-nm platinum NPs increased γ ray radiation effects by >40%, thus overcoming RR in Deinococcus radiodurans [192].
Moreover, drug resistance is responsible for chemotherapeutic failure in 90% of patients with metastatic tumors and is still a significant obstacle to achieving success in CT [193]. Among a number of causes, hypoxia is one of the main reasons for drug resistance in tumor cells [194,195]. Therefore, releasing tumor cells from hypoxic conditions is essential to overcoming drug resistance. Nitric oxide (NO) has hypoxia-relieving properties. It can reverse cancer cells’ drug resistance and enhance their sensitivity to radiotherapy [196,197,198]. However, NO has a short half-life, and its sensitivity toward biological substances limits its clinical application [199]. On the basis of these findings, Zhang et al. fabricated a nanotheranostic agent by functionalizing bismuth with S-nitrosothiol (Bi-SNO NPs). Upon exposure to radiation, X-ray broke down the S-N bond and stimulated release of a great amount of NO. Bi-SNO NPs (300 µg mL−1) 36 nm in size under 5 Gy of X-ray irradiation on HepG2 cells released NO, subsequently damaged DNA robustly, and overcame drug resistance [200].

10. Clinical Trials

Clinical trials help to ensure the safety and effectiveness of newly formulated treatment modalities, which is why clinical trials hold immense importance in developing treatment approaches and assessing recommended drug doses (RDs) [201]. A phase I study was performed to determine the RDs and safety profiles of hafnium oxide (HfO2) nanoparticles, called NBTXR3, and external beam RT (EBRT) in adult patients with locally advanced soft tissue sarcoma (STS) (trial registration number: NCT01433068). NBTXR3 at a dose of 53.3 g/L was initially injected with 50-Gy EBRT for five weeks, and the dose was escalated. The result demonstrated that NBTXR3 at 10% of tumor volume was the recommended rose, and intratumoral injection resulted in a 40% median tumor shrinkage rate and a 26% median percentage of residual viable tumor cells [202]. Later, a multicenter, randomized, phase II/III trial in 180 STS patients was conducted to determine the safety and efficacy of radiation-assisted NBTXR3 and compared it with radiotherapy alone (trial registration number: NCT02379845). Patients were divided into two groups randomly: one group received EBRT alone (5 × 2 Gy by week) over 5 weeks, while the other group received NBTXR3-mediated EBRT at the same radiation dose and time. The NBTXR3-mediated EBRT group provided a two-fold greater pathologic response rate compared to EBRT alone. The results of the trial also showed that 39% patients in the NBTXR3-mediated EBRT group experienced emergent side effects, whereas 30% patients in the RT group experienced serious adverse events, such as injection site pain, postoperative wound complications, radiation skin injuries, and hypotension [203]. Other clinical trials data demonstrated similar results. However, numbers of registered clinical trials have not showed desired results and came up with some complications. However, overall, the NP-based RT was more effective than treatment with radiation only. Trial numbers NCT01946867 and NCT02901483 (head and neck cancer), NCT02721056 (liver cancer), NCT02805894 (prostate cancer), and NCT02465593 (rectal cancer) have been registered at clinicaltrials.gov but terminated afterward. However, overall, NP-based RT showed better efficacy than treatment with radiation only [204].

11. Limitations and Future Perspective of NPs-Based RT

Despite of the tremendous advantages of NP-based RT, there are still challenges that need additional development [205]. One of the primary challenges to improving the therapeutic efficiency of NP-based RT is the delivery of optimal concentrations of required NPs to cancer cells with minimal or no side effects [206,207]. Administration of NP-based radiosensitizers through the systemic route is deemed to have negative effects on other organs, in addition to tumors. This outcome reduces the patient’s general health, specifically when localized radiosensitization is required [207,208]. Therefore, the therapeutic approach to combining NP-based radiosensitizers with RT should be carefully determined due to radiosensitizers’ toxicity to normal tissues [208]. Another shortcoming of systemic delivery is that NPs that have extended circulating times may not reach the tumor during RT. Such a delay in treatment may lead to dose enhancement compared to localized delivery [209]. Delivering NP-based radiosensitizers via implantation has been used largely in RT. Nevertheless, this strategy is limited to only specialized cancers, such as those of the lungs, breast, prostate, etc., where spacers along with fiducial markers are commonly used [209,210]. Similarly, the inhalation route is only effective in lung cancer [209]. Therefore, further extensive research is essential to develop the present strategies for systemic delivery of targeted NPs.
However, the application of NP-based RT has great prospects, notably for implanting radiosensitizers for in situ dose painting in RT [211,212]. Brachytherapy Application with In-situ Dose-painting through Gold Nanoparticle Eluters (BANDAGE) is considered as an excellent treatment window to enhance therapeutic efficiency during brachytherapy [211]. BANDAGE is expected to elevate the survival rate of prostate cancer patients who need salvage therapy but have reached the dose limit of radiotherapy. Moreover, BANDAGE is also anticipated to be used in prostate-seed brachytherapy treatment to stimulate local radiotherapeutic effects without toxicity [205]. NP-based RT plays a vital role in the advancement of cancer treatment, including concomitant chemoradiotherapy. The ultimate aim of NP-based RT is to obtain greater efficacy with minimum side effects and reduced tumor reoccurrence, helping to improve cancer prognosis and prolong patients’ lives [213]. The ideal cancer treatment modality would require the perfect radiosensitizer, which has not yet been reported [214]. Hence, thorough research is required in this field. However, improvising a delivery route may help to achieve substantial dose enhancement with negligible amounts of radiation and drug toxicity.

12. Conclusions

RT is one of the main and most important treatment strategies against cancer with up to 50% of all cancer patients receiving radiation treatment. RT kills tumors selectively by delivering intense ionizing radiation. Such targeted treatment has been used to cure solid, as well as metastatic, tumors. However, RT has disadvantages, such as causing injury to nearby normal cells. Additionally, tumors distant from the radiation site receive less intense ionizing radiation. Furthermore, RR is one of main causes of the failure of RT and subsequent relapses of tumors. Numbers of cases of radiotoxicity have been reported in the literature. The introduction of NPs in RT provides a state-of-the-art strategy in the field of radiation treatment by not only bypassing the limitations and side effects of RT but also showing significant tumoricidal activity compared to radiation only or NP-based treatment against all types of cancer. NPs with high Z elements show robust radiation absorption cross-sections and thus are used as radiosensitizers in external RT. Apart from high Z materials, nanostructured radiosensitizing agents of different types are also applied in NP-based radiotherapeutic treatment. The incident photon interacts with the NPs’ atoms and stimulate the ejection of photoelectrons or auger electrons, which cause the destruction of cancer cells. The synergistic effects of NPs and ionizing radiation interfere with DNA-repair processes, also producing ROS, which damage tumor DNA. Moreover, their combination arrest the cell cycle of cancer cells and regulate the TME to inhibit tumor progression. Additionally, conjugation of NPs with radiation treatment restricts bystander effects and keeps surrounding non-cancerous cells undamaged. Such associated treatment reduces hypoxia, maintains the TME, and obstructs the tumor metastasis process, contributing to increased RR of tumors. Moreover, introducing nano RT with phototherapy diminishes phototoxicity, at the same time offering complementary treatment against cancer. Overall, NPs can be employed as radiosensitizers in RT, offering new opportunities to progress and improve radiotherapeutic treatment. Their synergistic effects develop the clinical efficiency of RT against various types of cancers. Therefore, more clinical research needs to be conducted to heighten the NP-based radiotherapeutic approach to cancer treatment.

Author Contributions

Conceptualization, M.S.S. and M.H.; writing—original draft preparation, M.H., K.M.M. and M.S.S.; writing—review and editing, M.H. and M.S.S.; project administration, M.H. and M.S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gaidai, O.; Yan, P.; Xing, Y. Future world cancer death rate prediction. Sci. Rep. 2023, 13, 303. [Google Scholar] [CrossRef] [PubMed]
  2. Chhikara, B.S.; Parang, K. Global Cancer Statistics 2022: The trends projection analysis. Chem. Biol. Lett. 2023, 10, 451. [Google Scholar]
  3. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef] [PubMed]
  4. Thun, M.J.; DeLancey, J.O.; Center, M.M.; Jemal, A.; Ward, E.M. The global burden of cancer: Priorities for prevention. Carcinogenesis 2010, 31, 100–110. [Google Scholar] [CrossRef] [Green Version]
  5. Shakil, M.S.; Mahmud, K.M.; Sayem, M.; Niloy, M.S.; Halder, S.K.; Hossen, M.S.; Uddin, M.F.; Hasan, M.A. Using Chitosan or Chitosan Derivatives in Cancer Therapy. Polysaccharides 2021, 2, 795–816. [Google Scholar] [CrossRef]
  6. Matthews, H.K.; Bertoli, C.; de Bruin, R.A.M. Cell cycle control in cancer. Nat. Reviews Mol. Cell Biol. 2022, 23, 74–88. [Google Scholar] [CrossRef]
  7. Chen, Y.; Yang, J.; Fu, S.; Wu, J. Gold Nanoparticles as Radiosensitizers in Cancer Radiotherapy. Int. J. Nanomed. 2020, 15, 9407–9430. [Google Scholar] [CrossRef]
  8. Her, S.; Jaffray, D.A.; Allen, C. Gold nanoparticles for applications in cancer radiotherapy: Mechanisms and recent advancements. Adv. Drug Deliv. Rev. 2017, 109, 84–101. [Google Scholar] [CrossRef]
  9. Leroi, N.; Lallemand, F.; Coucke, P.; Noel, A.; Martinive, P. Impacts of Ionizing Radiation on the Different Compartments of the Tumor Microenvironment. Front. Pharmacol. 2016, 7, 78. [Google Scholar] [CrossRef] [Green Version]
  10. Song, G.; Cheng, L.; Chao, Y.; Yang, K.; Liu, Z. Emerging Nanotechnology and Advanced Materials for Cancer Radiation Therapy. Adv. Mater. 2017, 29, 1700996. [Google Scholar] [CrossRef]
  11. Liu, Y.; Zhang, P.; Li, F.; Jin, X.; Li, J.M.; Chen, W.; Li, Q.J.T. Metal-based NanoEnhancers for Future Radiotherapy: Radiosensitizing and Synergistic Effects on Tumor Cells. Theranostics 2018, 8, 1824–1849. [Google Scholar] [CrossRef]
  12. Baumann, M.; Krause, M.; Overgaard, J.; Debus, J.; Bentzen, S.M.; Daartz, J.; Richter, C.; Zips, D.; Bortfeld, T. Radiation oncology in the era of precision medicine. Nat. Rev. Cancer 2016, 16, 234–249. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, H.H.W.; Kuo, M.T. Improving radiotherapy in cancer treatment: Promises and challenges. Oncotarget 2017, 8, 62742–62758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Barton, M.B.; Jacob, S.; Shafiq, J.; Wong, K.; Thompson, S.R.; Hanna, T.P.; Delaney, G.P. Estimating the demand for radiotherapy from the evidence: A review of changes from 2003 to 2012. Radiother. Oncol. J. Eur. Soc. Ther. Radiol. Oncol. 2014, 112, 140–144. [Google Scholar] [CrossRef]
  15. Morrison, R.; Schleicher, S.M.; Sun, Y.; Niermann, K.J.; Kim, S.; Spratt, D.E.; Chung, C.H.; Lu, B. Targeting the Mechanisms of Resistance to Chemotherapy and Radiotherapy with the Cancer Stem Cell Hypothesis. J. Oncol. 2011, 2011, 941876. [Google Scholar] [CrossRef]
  16. Nagi, N.M.S.; Khair, Y.A.M.; Abdalla, A.M.E. Capacity of gold nanoparticles in cancer radiotherapy. Jpn. J. Radiol. 2017, 35, 555–561. [Google Scholar] [CrossRef]
  17. Butterworth, K.T.; McMahon, S.J.; Currell, F.J.; Prise, K.M. Physical basis and biological mechanisms of gold nanoparticle radiosensitization. Nanoscale 2012, 4, 4830–4838. [Google Scholar] [CrossRef] [PubMed]
  18. Sørensen, B.S.; Horsman, M.R. Tumor Hypoxia: Impact on Radiation Therapy and Molecular Pathways. Front. Oncol. 2020, 10, 562. [Google Scholar] [CrossRef] [Green Version]
  19. Dörr, W.; Hamilton, C.S.; Boyd, T.; Reed, B.; Denham, J.W. Radiation-induced changes in cellularity and proliferation in human oral mucosa. Int. J. Radiat. Oncol. 2002, 52, 911–917. [Google Scholar] [CrossRef]
  20. Adams, M.J.; Hardenbergh, P.H.; Constine, L.S.; Lipshultz, S.E. Radiation-associated cardiovascular disease. Rev. Oncol./Hematol. 2003, 45, 55–75. [Google Scholar] [CrossRef]
  21. Nakajima, T.; Ninomiya, Y.; Nenoi, M. Radiation-Induced Reactions in the Liver—Modulation of Radiation Effects by Lifestyle-Related Factors. Int. J. Mol. Sci. 2018, 19, 3855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Klaus, R.; Niyazi, M.; Lange-Sperandio, B. Radiation-induced kidney toxicity: Molecular and cellular pathogenesis. Radiat. Oncol. 2021, 16, 43. [Google Scholar] [CrossRef] [PubMed]
  23. Boerma, M.; Sridharan, V.; Mao, X.-W.; Nelson, G.A.; Cheema, A.K.; Koturbash, I.; Singh, S.P.; Tackett, A.J.; Hauer-Jensen, M. Effects of ionizing radiation on the heart. Mutat. Res./Rev. Mutat. Res. 2016, 770, 319–327. [Google Scholar] [CrossRef] [Green Version]
  24. Xie, D.; Wang, M.P.; Qi, W.H. A simplified model to calculate the surface-to-volume atomic ratio dependent cohesive energy of nanocrystals. J. Phys. Condens. Matter 2004, 16, L401. [Google Scholar] [CrossRef]
  25. Mahmud, K.M.; Hossain, M.M.; Polash, S.A.; Takikawa, M.; Shakil, M.S.; Uddin, M.F.; Alam, M.; Ali Khan Shawan, M.M.; Saha, T.; Takeoka, S.; et al. Investigation of Antimicrobial Activity and Biocompatibility of Biogenic Silver Nanoparticles Synthesized using Syzigyum cymosum Extract. ACS Omega 2022, 7, 27216–27229. [Google Scholar] [CrossRef]
  26. Pallares, R.M.; Choo, P.; Cole, L.E.; Mirkin, C.A.; Lee, A.; Odom, T.W. Manipulating Immune Activation of Macrophages by Tuning the Oligonucleotide Composition of Gold Nanoparticles. Bioconjugate Chem. 2019, 30, 2032–2037. [Google Scholar] [CrossRef] [PubMed]
  27. Engels, E.; Westlake, M.; Li, N.; Vogel, S.; Gobert, Q.; Thorpe, N.; Rosenfeld, A.; Lerch, M.; Corde, S.; Tehei, M. Thulium Oxide Nanoparticles: A new candidate for image-guided radiotherapy. Biomed. Phys. Eng. Express 2018, 4, 044001. [Google Scholar] [CrossRef] [Green Version]
  28. Lu, A.H.; Salabas, E.L.; Schüth, F. Magnetic nanoparticles: Synthesis, protection, functionalization, and application. Angew. Chem. 2007, 46, 1222–1244. [Google Scholar] [CrossRef]
  29. Sedlmeier, A.; Gorris, H.H. Surface modification and characterization of photon-upconverting nanoparticles for bioanalytical applications. Chem. Soc. Rev. 2015, 44, 1526–1560. [Google Scholar] [CrossRef] [Green Version]
  30. Hainfeld, J.F.; Dilmanian, F.A.; Slatkin, D.N.; Smilowitz, H.M. Radiotherapy enhancement with gold nanoparticles. J. Pharm. Pharmacol. 2010, 60, 977–985. [Google Scholar] [CrossRef] [Green Version]
  31. Roeske, J.C.; Nunez, L.; Hoggarth, M.; Labay, E.; Weichselbaum, R.R. Characterization of the theorectical radiation dose enhancement from nanoparticles. Technol. Cancer Res. Treat. 2007, 6, 395–401. [Google Scholar] [CrossRef]
  32. Wu, P.H.; Onodera, Y.; Ichikawa, Y.; Rankin, E.B.; Giaccia, A.J.; Watanabe, Y.; Qian, W.; Hashimoto, T.; Shirato, H.; Nam, J.M. Targeting integrins with RGD-conjugated gold nanoparticles in radiotherapy decreases the invasive activity of breast cancer cells. Int. J. Nanomed. 2017, 12, 5069–5085. [Google Scholar] [CrossRef] [Green Version]
  33. Du, F.; Lou, J.; Jiang, R.; Fang, Z.; Zhao, X.; Niu, Y.; Zou, S.; Zhang, M.; Gong, A.; Wu, C. Hyaluronic acid-functionalized bismuth oxide nanoparticles for computed tomography imaging-guided radiotherapy of tumor. Int. J. Nanomed. 2017, 12, 5973–5992. [Google Scholar] [CrossRef] [Green Version]
  34. Sykes, E.A.; Dai, Q.; Sarsons, C.D.; Chen, J.; Rocheleau, J.V.; Hwang, D.M.; Zheng, G.; Cramb, D.T.; Rinker, K.D.; Chan, W.C.W. Tailoring nanoparticle designs to target cancer based on tumor pathophysiology. Proc. Natl. Acad. Sci. USA 2016, 113, E1142–E1151. [Google Scholar] [CrossRef] [Green Version]
  35. Bennie, L.A.; McCarthy, H.O.; Coulter, J.A. Enhanced nanoparticle delivery exploiting tumour-responsive formulations. Cancer Nanotechnol. 2018, 9, 10. [Google Scholar] [CrossRef] [PubMed]
  36. Bonnet, M.; Hong, C.R.; Wong, W.W.; Liew, L.P.; Shome, A.; Wang, J.; Gu, Y.; Stevenson, R.J.; Qi, W.; Anderson, R.F.; et al. Next-Generation Hypoxic Cell Radiosensitizers: Nitroimidazole Alkylsulfonamides. J. Med. Chem. 2018, 61, 1241–1254. [Google Scholar] [CrossRef] [PubMed]
  37. Gal, O.; Betzer, O.; Rousso-Noori, L.; Sadan, T.; Motiei, M.; Nikitin, M.; Friedmann-Morvinski, D.; Popovtzer, R.; Popovtzer, A. Antibody Delivery into the Brain by Radiosensitizer Nanoparticles for Targeted Glioblastoma Therapy. J. Nanotheranostics 2022, 3, 177–188. [Google Scholar] [CrossRef] [PubMed]
  38. Yoon, S.W.; Tsvankin, V.; Shrock, Z.; Meng, B.; Zhang, X.; Dewhirst, M.; Fecci, P.; Adamson, J.; Oldham, M. Enhancing Radiation Therapy Through Cherenkov Light-Activated Phototherapy. Int. J. Radiat. Oncol. Biol. Phys. 2018, 100, 794–801. [Google Scholar] [CrossRef]
  39. Cao, W.; Gu, Y.; Meineck, M.; Xu, H. The combination of chemotherapy and radiotherapy towards more efficient drug delivery. Chem.–Asian J. 2014, 9, 48–57. [Google Scholar] [CrossRef]
  40. Laprise-Pelletier, M.; Simão, T.; Fortin, M.-A. Gold Nanoparticles in Radiotherapy and Recent Progress in Nanobrachytherapy. Adv. Healthc. Mater. 2018, 7, 1701460. [Google Scholar] [CrossRef]
  41. McMahon, S.J.; Paganetti, H.; Prise, K.M. Optimising element choice for nanoparticle radiosensitisers. Nanoscale 2016, 8, 581–589. [Google Scholar] [CrossRef] [Green Version]
  42. Kwatra, D.; Venugopal, A.; Anant, S. Nanoparticles in radiation therapy: A summary of various approaches to enhance radiosensitization in cancer. Transl. Cancer Res. 2013, 2, 330–342. [Google Scholar]
  43. Ku, A.; Facca, V.J.; Cai, Z.; Reilly, R.M. Auger electrons for cancer therapy-a review. EJNMMI Radiopharm. Chem. 2019, 4, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Kuncic, Z.; Lacombe, S. Nanoparticle radio-enhancement: Principles, progress and application to cancer treatment. Phys. Med. Biol. 2018, 63, 02TR01. [Google Scholar] [CrossRef] [PubMed]
  45. Rancoule, C.; Magné, N.; Vallard, A.; Guy, J.-B.; Rodriguez-Lafrasse, C.; Deutsch, E.; Chargari, C. Nanoparticles in radiation oncology: From bench-side to bedside. Cancer Lett. 2016, 375, 256–262. [Google Scholar] [CrossRef]
  46. Choi, J.; Kim, G.; Cho, S.B.; Im, H.-J. Radiosensitizing high-Z metal nanoparticles for enhanced radiotherapy of glioblastoma multiforme. J. Nanobiotechnology 2020, 18, 122. [Google Scholar] [CrossRef]
  47. Retif, P.; Pinel, S.; Toussaint, M.; Frochot, C.; Chouikrat, R.; Bastogne, T.; Barberi-Heyob, M. Nanoparticles for Radiation Therapy Enhancement: The Key Parameters. Theranostics 2015, 5, 1030–1044. [Google Scholar] [CrossRef] [Green Version]
  48. Cordelli, E.; Eleuteri, P.; Grollino, M.G.; Benassi, B.; Blandino, G.; Bartoleschi, C.; Pardini, M.C.; Di Caprio, E.V.; Spanò, M.; Pacchierotti, F.; et al. Direct and delayed X-ray-induced DNA damage in male mouse germ cells. Environ. Mol. Mutagen. 2012, 53, 429–439. [Google Scholar] [CrossRef] [PubMed]
  49. Kotb, O.M.; Brozhik, D.S.; Verbenko, V.N.; Gulevich, E.P.; Ezhov, V.F.; Karlin, D.L.; Pak, F.A.; Paston, S.V.; Polyanichko, A.M.; Khalikov, A.I.; et al. Investigation of DNA Damage Induced by Proton and Gamma Radiation. Biophysics 2021, 66, 202–208. [Google Scholar] [CrossRef]
  50. 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] [Green Version]
  51. Ui, A.; Chiba, N.; Yasui, A. Relationship among DNA double-strand break (DSB), DSB repair, and transcription prevents genome instability and cancer. Cancer Sci. 2020, 111, 1443–1451. [Google Scholar] [CrossRef] [PubMed]
  52. Xu, Y.; Xu, D. Repair pathway choice for double-strand breaks. Essays Biochem. 2020, 64, 765–777. [Google Scholar] [CrossRef] [PubMed]
  53. Ciccia, A.; Elledge, S.J. The DNA damage response: Making it safe to play with knives. Mol. Cell 2010, 40, 179–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Hoeijmakers, J.H. Genome maintenance mechanisms for preventing cancer. Nature 2001, 411, 366–374. [Google Scholar] [CrossRef]
  55. Menolfi, D.; Zha, S. ATM, ATR and DNA-PKcs kinases—The lessons from the mouse models: Inhibition ≠ deletion. Cell Biosci. 2020, 10, 8. [Google Scholar] [CrossRef]
  56. Huang, R.-X.; Zhou, P.-K. DNA damage response signaling pathways and targets for radiotherapy sensitization in cancer. Signal Transduct. Target. Ther. 2020, 5, 60. [Google Scholar] [CrossRef]
  57. Iliakis, G.; Wang, H.; Perrault, A.R.; Boecker, W.; Rosidi, B.; Windhofer, F.; Wu, W.; Guan, J.; Terzoudi, G.; Pantelias, G. Mechanisms of DNA double strand break repair and chromosome aberration formation. Cytogenet. Genome Res. 2004, 104, 14–20. [Google Scholar] [CrossRef]
  58. Elbakry, A.; Löbrich, M. Homologous Recombination Subpathways: A Tangle to Resolve. Front. Genet. 2021, 12, 723847. [Google Scholar] [CrossRef]
  59. Khanna, K.K.; Jackson, S.P. DNA double-strand breaks: Signaling, repair and the cancer connection. Nat. Genet. 2001, 27, 247–254. [Google Scholar] [CrossRef]
  60. Jackson, S.P.; Bartek, J. The DNA-damage response in human biology and disease. Nature 2009, 461, 1071–1078. [Google Scholar] [CrossRef] [Green Version]
  61. Li, L.-Y.; Guan, Y.-D.; Chen, X.-S.; Yang, J.-M.; Cheng, Y. DNA Repair Pathways in Cancer Therapy and Resistance. Front. Pharmacol. 2021, 11, 629266. [Google Scholar] [CrossRef]
  62. Alhmoud, J.F.; Woolley, J.F.; Al Moustafa, A.-E.; Malki, M.I. DNA Damage/Repair Management in Cancers. Cancers 2020, 12, 1050. [Google Scholar] [CrossRef] [PubMed]
  63. Andrs, M.; Korabecny, J.; Jun, D.; Hodny, Z.; Bartek, J.; Kuca, K. Phosphatidylinositol 3-Kinase (PI3K) and Phosphatidylinositol 3-Kinase-Related Kinase (PIKK) Inhibitors: Importance of the Morpholine Ring. J. Med. Chem. 2015, 58, 41–71. [Google Scholar] [CrossRef]
  64. Curtin, N.J. Inhibiting the DNA damage response as a therapeutic manoeuvre in cancer. Br. J. Pharmacol. 2013, 169, 1745–1765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Bhattacharya, S.; Asaithamby, A. Repurposing DNA repair factors to eradicate tumor cells upon radiotherapy. Transl. Cancer Res. 2017, 6, S822–S839. [Google Scholar] [CrossRef] [PubMed]
  66. Jain, R.K.; Stylianopoulos, T. Delivering nanomedicine to solid tumors. Nat. Reviews Clin. Oncol. 2010, 7, 653–664. [Google Scholar] [CrossRef] [Green Version]
  67. Mahmud, K.M.; Niloy, M.S.; Shakil, M.S.; Islam, M.A. Ruthenium Complexes: An Alternative to Platinum Drugs in Colorectal Cancer Treatment. Pharmaceutics 2021, 13, 1295. [Google Scholar] [CrossRef]
  68. Zhang, X.; Liu, N.; Shao, Z.; Qiu, H.; Yao, H.; Ji, J.; Wang, J.; Lu, W.; Chen, R.C.; Zhang, L. Folate-targeted nanoparticle delivery of androgen receptor shRNA enhances the sensitivity of hormone-independent prostate cancer to radiotherapy. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 1309–1321. [Google Scholar] [CrossRef]
  69. Löbrich, M.; Shibata, A.; Beucher, A.; Fisher, A.; Ensminger, M.; Goodarzi, A.A.; Barton, O.; Jeggo, P.A. gammaH2AX foci analysis for monitoring DNA double-strand break repair: Strengths, limitations and optimization. Cell Cycle 2010, 9, 662–669. [Google Scholar] [CrossRef] [Green Version]
  70. Olive, P.L. Detection of DNA damage in individual cells by analysis of histone H2AX phosphorylation. Methods Cell Biol. 2004, 75, 355–373. [Google Scholar]
  71. Yao, H.; Qiu, H.; Shao, Z.; Wang, G.; Wang, J.; Yao, Y.; Xin, Y.; Zhou, M.; Wang, A.Z.; Zhang, L. Nanoparticle formulation of small DNA molecules, Dbait, improves the sensitivity of hormone-independent prostate cancer to radiotherapy. Nanomed. Nanotechnol. Biol. Med. 2016, 12, 2261–2271. [Google Scholar] [CrossRef]
  72. Xia, W.; Low, P.S. Folate-targeted therapies for cancer. J. Med. Chem. 2010, 53, 6811–6824. [Google Scholar] [CrossRef]
  73. Quanz, M.; Berthault, N.; Roulin, C.; Roy, M.; Herbette, A.; Agrario, C.; Alberti, C.; Josserand, V.; Coll, J.L.; Sastre-Garau, X.; et al. Small-molecule drugs mimicking DNA damage: A new strategy for sensitizing tumors to radiotherapy. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2009, 15, 1308–1316. [Google Scholar] [CrossRef] [Green Version]
  74. Quanz, M.; Chassoux, D.; Berthault, N.; Agrario, C.; Sun, J.S.; Dutreix, M. Hyperactivation of DNA-PK by double-strand break mimicking molecules disorganizes DNA damage response. PLoS ONE 2009, 4, e6298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Biau, J.; Devun, F.; Jdey, W.; Kotula, E.; Quanz, M.; Chautard, E.; Sayarath, M.; Sun, J.S.; Verrelle, P.; Dutreix, M. A preclinical study combining the DNA repair inhibitor Dbait with radiotherapy for the treatment of melanoma. Neoplasia 2014, 16, 835–844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Devun, F.; Biau, J.; Huerre, M.; Croset, A.; Sun, J.S.; Denys, A.; Dutreix, M. Colorectal cancer metastasis: The DNA repair inhibitor Dbait increases sensitivity to hyperthermia and improves efficacy of radiofrequency ablation. Radiology 2014, 270, 736–746. [Google Scholar] [CrossRef]
  77. Herath, N.I.; Devun, F.; Lienafa, M.C.; Herbette, A.; Denys, A.; Sun, J.S.; Dutreix, M. The DNA Repair Inhibitor DT01 as a Novel Therapeutic Strategy for Chemosensitization of Colorectal Liver Metastasis. Mol. Cancer Ther. 2016, 15, 15–22. [Google Scholar] [CrossRef] [Green Version]
  78. Chithrani, D.B.; Jelveh, S.; Jalali, F.; van Prooijen, M.; Allen, C.; Bristow, R.G.; Hill, R.P.; Jaffray, D.A. Gold Nanoparticles as Radiation Sensitizers in Cancer Therapy. J. Radiat. Res. 2010, 173, 719–728. [Google Scholar] [CrossRef] [PubMed]
  79. Joh, D.Y.; Sun, L.; Stangl, M.; Al Zaki, A.; Murty, S.; Santoiemma, P.P.; Davis, J.J.; Baumann, B.C.; Alonso-Basanta, M.; Bhang, D.; et al. Selective Targeting of Brain Tumors with Gold Nanoparticle-Induced Radiosensitization. PLoS ONE 2013, 8, e62425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Zheng, Q.; Yang, H.; Wei, J.; Tong, J.-l.; Shu, Y.-q. The role and mechanisms of nanoparticles to enhance radiosensitivity in hepatocellular cell. Biomed. Pharmacother. 2013, 67, 569–575. [Google Scholar] [CrossRef]
  81. Chen, N.; Yang, W.; Bao, Y.; Xu, H.; Qin, S.; Tu, Y. BSA capped Au nanoparticle as an efficient sensitizer for glioblastoma tumor radiation therapy. RSC Adv. 2015, 5, 40514–40520. [Google Scholar] [CrossRef]
  82. Tian, X.; Lara, H.; Wagner, K.T.; Saripalli, S.; Hyder, S.N.; Foote, M.; Sethi, M.; Wang, E.; Caster, J.M.; Zhang, L.; et al. Improving DNA double-strand repair inhibitor KU55933 therapeutic index in cancer radiotherapy using nanoparticle drug delivery. Nanoscale 2015, 7, 20211–20219. [Google Scholar] [CrossRef] [Green Version]
  83. Porosnicu, I.; Butnaru, C.M.; Tiseanu, I.; Stancu, E.; Munteanu, C.V.A.; Bita, B.I.; Duliu, O.G.; Sima, F. Y2O3 Nanoparticles and X-ray Radiation-Induced Effects in Melanoma Cells. Molecules 2021, 26, 3403. [Google Scholar] [CrossRef]
  84. Hasegawa, T.; Takahashi, J.; Nagasawa, S.; Doi, M.; Moriyama, A.; Iwahashi, H. DNA Strand Break Properties of Protoporphyrin IX by X-Ray Irradiation against Melanoma. Int. J. Mol. Sci. 2020, 21, 2302. [Google Scholar] [CrossRef] [Green Version]
  85. Rosa, S.; Connolly, C.; Schettino, G.; Butterworth, K.T.; Prise, K.M. Biological mechanisms of gold nanoparticle radiosensitization. Cancer Nanotechnol. 2017, 8, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Goel, S.; Ni, D.; Cai, W. Harnessing the Power of Nanotechnology for Enhanced Radiation Therapy. ACS Nano 2017, 11, 5233–5237. [Google Scholar] [CrossRef] [PubMed]
  87. Li, Y.; Yang, J.; Sun, X. Reactive Oxygen Species-Based Nanomaterials for Cancer Therapy. Front. Chem. 2021, 9, 650587. [Google Scholar] [CrossRef]
  88. Xie, A.; Li, H.; Hao, Y.; Zhang, Y. Tuning the Toxicity of Reactive Oxygen Species into Advanced Tumor Therapy. Nanoscale Res. Lett. 2021, 16, 142. [Google Scholar] [CrossRef]
  89. Aggarwal, V.; Tuli, H.S.; Varol, A.; Thakral, F.; Yerer, M.B.; Sak, K.; Varol, M.; Jain, A.; Khan, M.A.; Sethi, G.J.B. Role of reactive oxygen species in cancer progression: Molecular mechanisms and recent advancements. Biomolecules 2019, 9, 735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Singh, R.; Manna, P.P. Reactive oxygen species in cancer progression and its role in therapeutics. Explor. Med. 2022, 3, 43–57. [Google Scholar] [CrossRef]
  91. Yong, Y.; Zhang, C.; Gu, Z.; Du, J.; Guo, Z.; Dong, X.; Xie, J.; Zhang, G.; Liu, X.; Zhao, Y. Polyoxometalate-Based Radiosensitization Platform for Treating Hypoxic Tumors by Attenuating Radioresistance and Enhancing Radiation Response. ACS Nano 2017, 11, 7164–7176. [Google Scholar] [CrossRef] [PubMed]
  92. Ma, Y.-C.; Tang, X.-F.; Xu, Y.-C.; Jiang, W.; Xin, Y.-J.; Zhao, W.; He, X.; Lu, L.-G.; Zhan, M.-X. Nano-enabled coordination platform of bismuth nitrate and cisplatin prodrug potentiates cancer chemoradiotherapy via DNA damage enhancement. Biomater. Sci. 2021, 9, 3401–3409. [Google Scholar] [CrossRef] [PubMed]
  93. Choi, B.J.; Jung, K.O.; Graves, E.E.; Pratx, G. A gold nanoparticle system for the enhancement of radiotherapy and simultaneous monitoring of reactive-oxygen-species formation. Nanotechnology 2018, 29, 504001. [Google Scholar] [CrossRef] [PubMed]
  94. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [Green Version]
  95. Chen, Y.; Li, N.; Wang, J.; Zhang, X.; Pan, W.; Yu, L.; Tang, B. Enhancement of mitochondrial ROS accumulation and radiotherapeutic efficacy using a Gd-doped titania nanosensitizer. Theranostics 2019, 9, 167–178. [Google Scholar] [CrossRef] [PubMed]
  96. Chen, F.; Zhuang, X.; Lin, L.; Yu, P.; Wang, Y.; Shi, Y.; Hu, G.; Sun, Y. New horizons in tumor microenvironment biology: Challenges and opportunities. BMC Med. 2015, 13, 45. [Google Scholar] [CrossRef] [Green Version]
  97. Weinberg, F.; Ramnath, N.; Nagrath, D.J.C. Reactive oxygen species in the tumor microenvironment: An overview. Cancers 2019, 11, 1191. [Google Scholar] [CrossRef] [Green Version]
  98. Whiteside, T.L. The tumor microenvironment and its role in promoting tumor growth. Oncogene 2008, 27, 5904–5912. [Google Scholar] [CrossRef] [Green Version]
  99. Wu, T.; Dai, Y. Tumor microenvironment and therapeutic response. Cancer Lett. 2017, 387, 61–68. [Google Scholar] [CrossRef]
  100. Zhu, S.; Gu, Z.; Zhao, Y. Harnessing Tumor Microenvironment for Nanoparticle-Mediated Radiotherapy. Adv. Ther. 2018, 1, 1800050. [Google Scholar] [CrossRef]
  101. Gu, H.; Huang, T.; Shen, Y.; Liu, Y.; Zhou, F.; Jin, Y.; Sattar, H.; Wei, Y. Reactive Oxygen Species-Mediated Tumor Microenvironment Transformation: The Mechanism of Radioresistant Gastric Cancer. Oxidative Med. Cell. Longev. 2018, 2018, 5801209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Wang, Y.; Shang, W.; Niu, M.; Tian, J.; Xu, K. Hypoxia-active nanoparticles used in tumor theranostic. Int. J. Nanomed. 2019, 14, 3705–3722. [Google Scholar] [CrossRef] [Green Version]
  103. Horsman, M.R.; Overgaard, J. The impact of hypoxia and its modification of the outcome of radiotherapy. J. Radiat. Res. 2016, 57, i90–i98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Semenza, G.L. Hypoxia-inducible factors: Mediators of cancer progression and targets for cancer therapy. Trends Pharmacol. Sci. 2012, 33, 207–214. [Google Scholar] [CrossRef] [Green Version]
  105. Schito, L.; Semenza, G.L. Hypoxia-Inducible Factors: Master Regulators of Cancer Progression. Trends Cancer 2016, 2, 758–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Patiar, S.; Harris, A.L. Role of hypoxia-inducible factor-1alpha as a cancer therapy target. Endocr.-Relat. Cancer 2006, 13 (Suppl. 1), S61–S75. [Google Scholar] [CrossRef]
  107. Vaupel, P. Tumor microenvironmental physiology and its implications for radiation oncology. Semin. Radiat. Oncol. 2004, 14, 198–206. [Google Scholar] [CrossRef]
  108. Gao, M.; Liang, C.; Song, X.; Chen, Q.; Jin, Q.; Wang, C.; Liu, Z. Erythrocyte-Membrane-Enveloped Perfluorocarbon as Nanoscale Artificial Red Blood Cells to Relieve Tumor Hypoxia and Enhance Cancer Radiotherapy. Adv. Mater. 2017, 29, 1701429. [Google Scholar] [CrossRef]
  109. Kim, M.S.; Lee, E.-J.; Kim, J.-W.; Chung, U.S.; Koh, W.-G.; Keum, K.C.; Koom, W.S. Gold nanoparticles enhance anti-tumor effect of radiotherapy to hypoxic tumor. Radiat. Oncol. J. 2016, 34, 230–238. [Google Scholar] [CrossRef] [Green Version]
  110. Pawlik, T.M.; Keyomarsi, K. Role of cell cycle in mediating sensitivity to radiotherapy. Int. J. Radiat. Oncol. Biol. Phys. 2004, 59, 928–942. [Google Scholar] [CrossRef]
  111. Kastan, M.B.; Bartek, J. Cell-cycle checkpoints and cancer. Nature 2004, 432, 316–323. [Google Scholar] [CrossRef] [PubMed]
  112. Geng, F.; Song, K.; Xing, J.Z.; Yuan, C.; Yan, S.; Yang, Q.; Chen, J.; Kong, B. Thio-glucose bound gold nanoparticles enhance radio-cytotoxic targeting of ovarian cancer. Nanotechnology 2011, 22, 285101. [Google Scholar] [CrossRef] [PubMed]
  113. Mackey, M.A.; Saira, F.; Mahmoud, M.A.; El-Sayed, M.A. Inducing Cancer Cell Death by Targeting Its Nucleus: Solid Gold Nanospheres versus Hollow Gold Nanocages. Bioconjugate Chem. 2013, 24, 897–906. [Google Scholar] [CrossRef] [Green Version]
  114. Turner, J.; Koumenis, C.; Kute, T.E.; Planalp, R.P.; Brechbiel, M.W.; Beardsley, D.; Cody, B.; Brown, K.D.; Torti, F.M.; Torti, S.V. Tachpyridine, a metal chelator, induces G2 cell-cycle arrest, activates checkpoint kinases, and sensitizes cells to ionizing radiation. Blood 2005, 106, 3191–3199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Roa, W.; Zhang, X.; Guo, L.; Shaw, A.; Hu, X.; Xiong, Y.; Gulavita, S.; Patel, S.; Sun, X.; Chen, J.; et al. Gold nanoparticle sensitize radiotherapy of prostate cancer cells by regulation of the cell cycle. Nanotechnology 2009, 20, 375101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Chen, F.; Zhang, X.H.; Hu, X.D.; Liu, P.D.; Zhang, H.Q. The effects of combined selenium nanoparticles and radiation therapy on breast cancer cells in vitro. Artif. Cells Nanomed. Biotechnol. 2018, 46, 937–948. [Google Scholar] [CrossRef] [Green Version]
  117. Zhang, J.; Wang, X.; Xu, T. Elemental selenium at nano size (Nano-Se) as a potential chemopreventive agent with reduced risk of selenium toxicity: Comparison with se-methylselenocysteine in mice. Toxicol. Sci. Off. J. Soc. Toxicol. 2008, 101, 22–31. [Google Scholar] [CrossRef] [Green Version]
  118. Bhattacharjee, A.; Basu, A.; Ghosh, P.; Biswas, J.; Bhattacharya, S. Protective effect of Selenium nanoparticle against cyclophosphamide induced hepatotoxicity and genotoxicity in Swiss albino mice. J. Biomater. Appl. 2014, 29, 303–317. [Google Scholar] [CrossRef]
  119. Xu, W.; Luo, T.; Li, P.; Zhou, C.; Cui, D.; Pang, B.; Ren, Q.; Fu, S. RGD-conjugated gold nanorods induce radiosensitization in melanoma cancer cells by downregulating α(v)β₃ expression. Int. J. Nanomed. 2012, 7, 915–924. [Google Scholar]
  120. Chen, Q.; Chen, J.; Yang, Z.; Xu, J.; Xu, L.; Liang, C.; Han, X.; Liu, Z. Nanoparticle-enhanced radiotherapy to trigger robust cancer immunotherapy. Adv. Mater. 2019, 31, 1802228. [Google Scholar] [CrossRef]
  121. Guan, X.; Sun, L.; Shen, Y.; Jin, F.; Bo, X.; Zhu, C.; Han, X.; Li, X.; Chen, Y.; Xu, H.; et al. Nanoparticle-enhanced radiotherapy synergizes with PD-L1 blockade to limit post-surgical cancer recurrence and metastasis. Nat. Commun. 2022, 13, 2834. [Google Scholar] [CrossRef]
  122. Hainfeld, J.F.; Slatkin, D.N.; Smilowitz, H.M. The use of gold nanoparticles to enhance radiotherapy in mice. Phys. Med. Biol. 2004, 49, N309. [Google Scholar] [CrossRef]
  123. Bi, Z.; Li, Q.; Dinglin, X.; Xu, Y.; You, K.; Hong, H.; Hu, Q.; Zhang, W.; Li, C.; Tan, Y.; et al. Nanoparticles (NPs)-Meditated LncRNA AFAP1-AS1 Silencing to Block Wnt/β-Catenin Signaling Pathway for Synergistic Reversal of Radioresistance and Effective Cancer Radiotherapy. Adv. Sci. 2020, 7, 2000915. [Google Scholar] [CrossRef] [PubMed]
  124. Ding, Y.; Sun, Z.; Tong, Z.; Zhang, S.; Min, J.; Xu, Q.; Zhou, L.; Mao, Z.; Xia, H.; Wang, W. Tumor microenvironment-responsive multifunctional peptide coated ultrasmall gold nanoparticles and their application in cancer radiotherapy. Theranostics 2020, 10, 5195–5208. [Google Scholar] [CrossRef]
  125. Hazkani, I.; Motiei, M.; Betzer, O.; Sadan, T.; Bragilovski, D.; Lubimov, L.; Mizrachi, A.; Hadar, T.; Levi, M.; Ben-Aharon, I.; et al. Can molecular profiling enhance radiotherapy? Impact of personalized targeted gold nanoparticles on radiosensitivity and imaging of adenoid cystic carcinoma. Theranostics 2017, 7, 3962–3971. [Google Scholar] [CrossRef] [Green Version]
  126. González-Ruíz, A.; Ferro-Flores, G.; Jiménez-Mancilla, N.; Escudero-Castellanos, A.; Ocampo-García, B.; Luna-Gutiérrez, M.; Santos-Cuevas, C.; Morales-Avila, E.; Isaac-Olivé, K. In vitro and in vivo synergistic effect of radiotherapy and plasmonic photothermal therapy on the viability of cancer cells using 177Lu–Au-NLS-RGD-Aptamer nanoparticles under laser irradiation. J. Radioanal. Nucl. Chem. 2018, 318, 1913–1921. [Google Scholar] [CrossRef]
  127. Cui, F.-B.; Li, R.-T.; Liu, Q.; Wu, P.-Y.; Hu, W.-J.; Yue, G.-F.; Ding, H.; Yu, L.-X.; Qian, X.-P.; Liu, B.-R. Enhancement of radiotherapy efficacy by docetaxel-loaded gelatinase-stimuli PEG-Pep-PCL nanoparticles in gastric cancer. Cancer Lett. 2014, 346, 53–62. [Google Scholar] [CrossRef]
  128. Hainfeld, J.F.; Dilmanian, F.A.; Zhong, Z.; Slatkin, D.N.; Kalef-Ezra, J.A.; Smilowitz, H.M. Gold nanoparticles enhance the radiation therapy of a murine squamous cell carcinoma. Phys. Med. Biol. 2010, 55, 3045. [Google Scholar] [CrossRef] [PubMed]
  129. Hainfeld, J.F.; Smilowitz, H.M.; O’Connor, M.J.; Dilmanian, F.A.; Slatkin, D.N. Gold nanoparticle imaging and radiotherapy of brain tumors in mice. Nanomedicine 2013, 8, 1601–1609. [Google Scholar] [CrossRef] [Green Version]
  130. Yi, X.; Chen, L.; Chen, J.; Maiti, D.; Chai, Z.; Liu, Z.; Yang, K. Biomimetic Copper Sulfide for Chemo-Radiotherapy: Enhanced Uptake and Reduced Efflux of Nanoparticles for Tumor Cells under Ionizing Radiation. Adv. Funct. Mater. 2018, 28, 1705161. [Google Scholar] [CrossRef]
  131. Kim, J.A.; Åberg, C.; Salvati, A.; Dawson, K.A. Role of cell cycle on the cellular uptake and dilution of nanoparticles in a cell population. Nat. Nanotechnol. 2012, 7, 62–68. [Google Scholar] [CrossRef] [PubMed]
  132. Liu, Y.; Chen, W.; Zhang, P.; Jin, X.; Liu, X.; Li, P.; Li, F.; Zhang, H.; Zou, G.; Li, Q. Dynamically-enhanced retention of gold nanoclusters in HeLa cells following X-rays exposure: A cell cycle phase-dependent targeting approach. Radiother. Oncol. 2016, 119, 544–551. [Google Scholar] [CrossRef] [PubMed]
  133. Davies Cde, L.; Lundstrøm, L.M.; Frengen, J.; Eikenes, L.; Bruland, S.Ø.; Kaalhus, O.; Hjelstuen, M.H.; Brekken, C. Radiation improves the distribution and uptake of liposomal doxorubicin (caelyx) in human osteosarcoma xenografts. Cancer Res. 2004, 64, 547–553. [Google Scholar] [CrossRef] [Green Version]
  134. Sheikh-Hamad, D.; Timmins, K.; Jalali, Z. Cisplatin-induced renal toxicity: Possible reversal by N-acetylcysteine treatment. J. Am. Soc. Nephrol. JASN 1997, 8, 1640–1644. [Google Scholar] [CrossRef] [PubMed]
  135. Liu, W.; Chen, B.; Zheng, H.; Xing, Y.; Chen, G.; Zhou, P.; Qian, L.; Min, Y. Advances of Nanomedicine in Radiotherapy. Pharmaceutics 2021, 13, 1757. [Google Scholar] [CrossRef]
  136. Tan, G.-R.; Feng, S.-S.; Leong, D.T. The reduction of anti-cancer drug antagonism by the spatial protection of drugs with PLA–TPGS nanoparticles. Biomaterials 2014, 35, 3044–3051. [Google Scholar] [CrossRef]
  137. Soo Choi, H.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J.P.; Itty Ipe, B.; Bawendi, M.G.; Frangioni, J.V. Renal clearance of quantum dots. Nat. Biotechnol. 2007, 25, 1165–1170. [Google Scholar] [CrossRef] [Green Version]
  138. Zhang, X.D.; Luo, Z.; Chen, J.; Shen, X.; Song, S.; Sun, Y.; Fan, S.; Fan, F.; Leong, D.T.; Xie, J. Ultrasmall Au(10-12)(SG)(10-12) nanomolecules for high tumor specificity and cancer radiotherapy. Adv. Mater. 2014, 26, 4565–4568. [Google Scholar] [CrossRef] [Green Version]
  139. Marín, A.; Martín, M.; Liñán, O.; Alvarenga, F.; López, M.; Fernández, L.; Büchser, D.; Cerezo, L. Bystander effects and radiotherapy. Rep. Pract. Oncol. Radiother. 2015, 20, 12–21. [Google Scholar] [CrossRef] [Green Version]
  140. Mothersill, C.; Seymour, C. Uncomfortable issues in radiation protection posed by low-dose radiobiology. Radiat. Environ. Biophys. 2013, 52, 293–298. [Google Scholar] [CrossRef]
  141. Mancuso, M.; Pasquali, E.; Giardullo, P.; Leonardi, S.; Tanori, M.; Di Majo, V.; Pazzaglia, S.; Saran, A. The radiation bystander effect and its potential implications for human health. Curr. Mol. Med. 2012, 12, 613–624. [Google Scholar] [CrossRef] [PubMed]
  142. Najafi, M.; Fardid, R.; Hadadi, G.; Fardid, M. The mechanisms of radiation-induced bystander effect. J. Biomed. Phys. Eng. 2014, 4, 163–172. [Google Scholar] [PubMed]
  143. Heuskin, A.-C.; Wéra, A.-C.; Riquier, H.; Michiels, C.; Lucas, S. Low-Dose Hypersensitivity and Bystander Effect are Not Mutually Exclusive in A549 Lung Carcinoma Cells after Irradiation with Charged Particles. J. Radiat. Res. 2013, 180, 491–498. [Google Scholar] [CrossRef] [PubMed]
  144. Rostami, A.; Toossi, M.T.B.; Sazgarnia, A.; Soleymanifard, S. The effect of glucose-coated gold nanoparticles on radiation bystander effect induced in MCF-7 and QUDB cell lines. Radiat. Environ. Biophys. 2016, 55, 461–466. [Google Scholar] [CrossRef] [PubMed]
  145. Zhang, L.; Luo, Y.; Lu, Z.; He, J.; Wang, L.; Zhang, L.; Zhang, Y.; Liu, Y. Astragalus Polysaccharide Inhibits Ionizing Radiation-Induced Bystander Effects by Regulating MAPK/NF-kB Signaling Pathway in Bone Mesenchymal Stem Cells (BMSCs). Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2018, 24, 4649–4658. [Google Scholar] [CrossRef]
  146. Azzam, I.E.; de Toledo, M.S.; Little, B.J. Stress Signaling from Irradiated to Non-Irradiated Cells. Curr. Cancer Drug Targets 2004, 4, 53–64. [Google Scholar] [CrossRef]
  147. Barber, G.N. Cytoplasmic DNA innate immune pathways. Immunol. Rev. 2011, 243, 99–108. [Google Scholar] [CrossRef]
  148. Mittra, I.; Samant, U.; Sharma, S.; Raghuram, G.V.; Saha, T.; Tidke, P.; Pancholi, N.; Gupta, D.; Prasannan, P.; Gaikwad, A.; et al. Cell-free chromatin from dying cancer cells integrate into genomes of bystander healthy cells to induce DNA damage and inflammation. Cell Death Discov. 2017, 3, 17015. [Google Scholar] [CrossRef] [Green Version]
  149. Kirolikar, S.; Prasannan, P.; Raghuram, G.V.; Pancholi, N.; Saha, T.; Tidke, P.; Chaudhari, P.; Shaikh, A.; Rane, B.; Pandey, R.; et al. Prevention of radiation-induced bystander effects by agents that inactivate cell-free chromatin released from irradiated dying cells. Cell Death Discov. 2018, 9, 1142. [Google Scholar] [CrossRef] [Green Version]
  150. Mohd Zainudin, N.H.; Talik Sisin, N.N.; Rashid, R.A.; Jamil, A.; Khairil Anuar, M.A.; Razak, K.A.; Abdullah, R.; Rahman, W.N. Cellular analysis on the radiation induced bystander effects due to bismuth oxide nanoparticles with 6 MV photon beam radiotherapy. J. Radiat. Res. Appl. Sci. 2022, 15, 318–325. [Google Scholar] [CrossRef]
  151. Sinha, N.; Mukhopadhyay, S.; Das, D.N.; Panda, P.K.; Bhutia, S.K. Relevance of cancer initiating/stem cells in carcinogenesis and therapy resistance in oral cancer. Oral Oncol. 2013, 49, 854–862. [Google Scholar] [CrossRef]
  152. Aponte, P.M.; Caicedo, A. Stemness in Cancer: Stem Cells, Cancer Stem Cells, and Their Microenvironment. Stem Cells Int. 2017, 2017, 5619472. [Google Scholar] [CrossRef]
  153. Qin, W.; Huang, G.; Chen, Z.; Zhang, Y. Nanomaterials in Targeting Cancer Stem Cells for Cancer Therapy. Front. Pharmacol. 2017, 8, 1. [Google Scholar] [CrossRef]
  154. Dean, M.; Fojo, T.; Bates, S. Tumour stem cells and drug resistance. Nat. Rev. Cancer 2005, 5, 275–284. [Google Scholar] [CrossRef]
  155. Eyler, C.E.; Rich, J.N. Survival of the Fittest: Cancer Stem Cells in Therapeutic Resistance and Angiogenesis. J. Clin. Oncol. Off. J. Am. Soc. Clin. Oncol. 2008, 26, 2839–2845. [Google Scholar] [CrossRef] [Green Version]
  156. Fiorillo, M.; Verre, A.F.; Iliut, M.; Peiris-Pagés, M.; Ozsvari, B.; Gandara, R.; Cappello, A.R.; Sotgia, F.; Vijayaraghavan, A.; Lisanti, M.P. Graphene oxide selectively targets cancer stem cells, across multiple tumor types: Implications for non-toxic cancer treatment, via “differentiation-based nano-therapy”. Oncotarget 2015, 6, 3553–3562. [Google Scholar] [CrossRef] [Green Version]
  157. Yao, H.-J.; Zhang, Y.-G.; Sun, L.; Liu, Y. The effect of hyaluronic acid functionalized carbon nanotubes loaded with salinomycin on gastric cancer stem cells. Biomaterials 2014, 35, 9208–9223. [Google Scholar] [CrossRef]
  158. Faraj, A.A.; Shaik, A.S.; Sayed, B.A.; Halwani, R.; Jammaz, I.A. Specific targeting and noninvasive imaging of breast cancer stem cells using single-walled carbon nanotubes as novel multimodality nanoprobes. Nanomedicine 2016, 11, 31–46. [Google Scholar] [CrossRef]
  159. Shi, H.; Sadler, P.J. How promising is phototherapy for cancer? Br. J. Cancer 2020, 123, 871–873. [Google Scholar] [CrossRef]
  160. Chen, J.; Ning, C.; Zhou, Z.; Yu, P.; Zhu, Y.; Tan, G.; Mao, C. Nanomaterials as photothermal therapeutic agents. Prog. Mater. Sci. 2019, 99, 1–26. [Google Scholar] [CrossRef]
  161. Ban, Q.; Bai, T.; Duan, X.; Kong, J. Noninvasive photothermal cancer therapy nanoplatforms via integrating nanomaterials and functional polymers. Biomater. Sci. 2017, 5, 190–210. [Google Scholar] [CrossRef]
  162. Gai, S.; Yang, G.; Yang, P.; He, F.; Lin, J.; Jin, D.; Xing, B. Recent advances in functional nanomaterials for light–triggered cancer therapy. Nano Today 2018, 19, 146–187. [Google Scholar] [CrossRef]
  163. Suo, X.; Zhang, J.; Zhang, Y.; Liang, X.-J.; Zhang, J.; Liu, D. A nano-based thermotherapy for cancer stem cell-targeted therapy. J. Mater. Chem. B 2020, 8, 3985–4001. [Google Scholar] [CrossRef]
  164. Zhang, L.; Jing, D.; Wang, L.; Sun, Y.; Li, J.J.; Hill, B.; Yang, F.; Li, Y.; Lam, K.S. Unique Photochemo-Immuno-Nanoplatform against Orthotopic Xenograft Oral Cancer and Metastatic Syngeneic Breast Cancer. Nano Lett. 2018, 18, 7092–7103. [Google Scholar] [CrossRef]
  165. Li, J.; Pu, K. Semiconducting Polymer Nanomaterials as Near-Infrared Photoactivatable Protherapeutics for Cancer. Acc. Chem. Res. 2020, 53, 752–762. [Google Scholar] [CrossRef]
  166. Dąbrowski, J.M. Chapter Nine-Reactive Oxygen Species in Photodynamic Therapy: Mechanisms of Their Generation and Potentiation. In Advances in Inorganic Chemistry; van Eldik, R., Hubbard, C.D., Eds.; Academic Press: Cambridge, MA, USA, 2017; Volume 70, pp. 343–394. [Google Scholar]
  167. Zhou, Z.; Song, J.; Nie, L.; Chen, X. Reactive oxygen species generating systems meeting challenges of photodynamic cancer therapy. Chem. Soc. Rev. 2016, 45, 6597–6626. [Google Scholar] [CrossRef] [Green Version]
  168. Li, X.; Kwon, N.; Guo, T.; Liu, Z.; Yoon, J. Innovative Strategies for Hypoxic-Tumor Photodynamic Therapy. Angew. Chem. Int. Ed. 2018, 57, 11522–11531. [Google Scholar] [CrossRef]
  169. Wang, G.D.; Nguyen, H.T.; Chen, H.; Cox, P.B.; Wang, L.; Nagata, K.; Hao, Z.; Wang, A.; Li, Z.; Xie, J. X-Ray Induced Photodynamic Therapy: A Combination of Radiotherapy and Photodynamic Therapy. Theranostics 2016, 6, 2295–2305. [Google Scholar] [CrossRef]
  170. Wang, J.; Wu, X.; Shen, P.; Wang, J.; Shen, Y.; Shen, Y.; Webster, T.J.; Deng, J. Applications of Inorganic Nanomaterials in Photothermal Therapy Based on Combinational Cancer Treatment. Int. J. Nanomed. 2020, 15, 1903–1914. [Google Scholar] [CrossRef] [Green Version]
  171. Hadi, F.; Tavakkol, S.; Laurent, S.; Pirhajati, V.; Mahdavi, S.R.; Neshastehriz, A.; Shakeri-Zadeh, A. Combinatorial effects of radiofrequency hyperthermia and radiotherapy in the presence of magneto-plasmonic nanoparticles on MCF-7 breast cancer cells. J. Cell. Physiol. 2019, 234, 20028–20035. [Google Scholar] [CrossRef]
  172. Hosseini, V.; Mirrahimi, M.; Shakeri-Zadeh, A.; Koosha, F.; Ghalandari, B.; Maleki, S.; Komeili, A.; Kamrava, S.K. Multimodal cancer cell therapy using Au@Fe2O3 core–shell nanoparticles in combination with photo-thermo-radiotherapy. Photodiagnosis Photodyn. Ther. 2018, 24, 129–135. [Google Scholar] [CrossRef]
  173. Movahedi, M.M.; Alamzadeh, Z.; Hosseini-Nami, S.; Shakeri-Zadeh, A.; Taheripak, G.; Ahmadi, A.; Zare-Sadeghi, A.; Ghaznavi, H.; Mehdizadeh, A. Investigating the mechanisms behind extensive death in human cancer cells following nanoparticle assisted photo-thermo-radiotherapy. Photodiagnosis Photodyn. Ther. 2020, 29, 101600. [Google Scholar] [CrossRef]
  174. Ghasemi, A.; Khanzadeh, T.; Zadi Heydarabad, M.; Khorrami, A.; Jahanban Esfahlan, A.; Ghavipanjeh, S.; Gholipour Belverdi, M.; Darvishani Fikouhi, S.; Darbin, A.; Najafpour, M.; et al. Evaluation of BAX and BCL-2 Gene Expression and Apoptosis Induction in Acute Lymphoblastic Leukemia Cell Line CCRFCEM after High-Dose Prednisolone Treatment. Asian Pac. J. Cancer Prev. APJCP 2018, 19, 2319–2323. [Google Scholar] [PubMed]
  175. Schildkopf, P.; Frey, B.; Ott, O.J.; Rubner, Y.; Multhoff, G.; Sauer, R.; Fietkau, R.; Gaipl, U.S. Radiation combined with hyperthermia induces HSP70-dependent maturation of dendritic cells and release of pro-inflammatory cytokines by dendritic cells and macrophages. Radiother. Oncol. J. Eur. Soc. Ther. Radiol. Oncol. 2011, 101, 109–115. [Google Scholar] [CrossRef] [PubMed]
  176. Movahedi, M.M.; Mehdizadeh, A.; Koosha, F.; Eslahi, N.; Mahabadi, V.P.; Ghaznavi, H.; Shakeri-Zadeh, A. Investigating the photo-thermo-radiosensitization effects of folate-conjugated gold nanorods on KB nasopharyngeal carcinoma cells. Photodiagnosis Photodyn. Ther. 2018, 24, 324–331. [Google Scholar] [CrossRef]
  177. Zhou, M.; Chen, Y.; Adachi, M.; Wen, X.; Erwin, B.; Mawlawi, O.; Lai, S.Y.; Li, C. Single agent nanoparticle for radiotherapy and radio-photothermal therapy in anaplastic thyroid cancer. Biomaterials 2015, 57, 41–49. [Google Scholar] [CrossRef] [Green Version]
  178. Liu, T.; Song, Y.; Huang, Z.; Pu, X.; Wang, Y.; Yin, G.; Gou, L.; Weng, J.; Meng, X. Photothermal photodynamic therapy and enhanced radiotherapy of targeting copolymer-coated liquid metal nanoparticles on liver cancer. Colloids Surf. B Biointerfaces 2021, 207, 112023. [Google Scholar] [CrossRef]
  179. Yong, Y.; Cheng, X.; Bao, T.; Zu, M.; Yan, L.; Yin, W.; Ge, C.; Wang, D.; Gu, Z.; Zhao, Y. Tungsten Sulfide Quantum Dots as Multifunctional Nanotheranostics for In Vivo Dual-Modal Image-Guided Photothermal/Radiotherapy Synergistic Therapy. ACS Nano 2015, 9, 12451–12463. [Google Scholar] [CrossRef] [PubMed]
  180. Yi, X.; Yang, K.; Liang, C.; Zhong, X.; Ning, P.; Song, G.; Wang, D.; Ge, C.; Chen, C.; Chai, Z.; et al. Imaging-Guided Combined Photothermal and Radiotherapy to Treat Subcutaneous and Metastatic Tumors Using Iodine-131-Doped Copper Sulfide Nanoparticles. Adv. Funct. Mater. 2015, 25, 4689–4699. [Google Scholar] [CrossRef]
  181. Safari, A.; Sarikhani, A.; Shahbazi-Gahrouei, D.; Alamzadeh, Z.; Beik, J.; Dezfuli, A.S.; Mahabadi, V.P.; Tohfeh, M.; Shakeri-Zadeh, A. Optimal scheduling of the nanoparticle-mediated cancer photo-thermo-radiotherapy. Photodiagnosis Photodyn. Ther. 2020, 32, 102061. [Google Scholar] [CrossRef]
  182. Galeaz, C.; Totis, C.; Bisio, A. Radiation Resistance: A Matter of Transcription Factors. Front. Oncol. 2021, 11, 662840. [Google Scholar] [CrossRef]
  183. Al-Dimassi, S.; Abou-Antoun, T.; El-Sibai, M. Cancer cell resistance mechanisms: A mini review. Clin. Transl. Oncol. 2014, 16, 511–516. [Google Scholar] [CrossRef] [PubMed]
  184. Barker, H.E.; Paget, J.T.; Khan, A.A.; Harrington, K.J. The tumour microenvironment after radiotherapy: Mechanisms of resistance and recurrence. Nat. Reviews Cancer 2015, 15, 409–425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Carmeliet, P.; Jain, R.K. Principles and mechanisms of vessel normalization for cancer and other angiogenic diseases. Nat. Reviews Drug Discov. 2011, 10, 417–427. [Google Scholar] [CrossRef]
  186. Li, W.; Zhao, X.; Du, B.; Li, X.; Liu, S.; Yang, X.-Y.; Ding, H.; Yang, W.; Pan, F.; Wu, X.; et al. Gold Nanoparticle–Mediated Targeted Delivery of Recombinant Human Endostatin Normalizes Tumour Vasculature and Improves Cancer Therapy. Sci. Rep. 2016, 6, 30619. [Google Scholar] [CrossRef] [Green Version]
  187. Jayne, D.G.; Perry, S.L.; Morrison, E.; Farmery, S.M.; Guillou, P.J. Activated mesothelial cells produce heparin-binding growth factors: Implications for tumour metastases. Br. J. Cancer 2000, 82, 1233–1238. [Google Scholar] [CrossRef] [Green Version]
  188. Child, H.W.; Hernandez, Y.; Conde, J.; Mullin, M.; Baptista, P.; de la Fuente, J.M.; Berry, C.C. Gold nanoparticle-siRNA mediated oncogene knockdown at RNA and protein level, with associated gene effects. Nanomedicine 2015, 10, 2513–2525. [Google Scholar] [CrossRef]
  189. Lee, K.J.; Ko, E.J.; Park, Y.-Y.; Park, S.S.; Ju, E.J.; Park, J.; Shin, S.H.; Suh, Y.-A.; Hong, S.-M.; Park, I.J.; et al. A novel nanoparticle-based theranostic agent targeting LRP-1 enhances the efficacy of neoadjuvant radiotherapy in colorectal cancer. Biomaterials 2020, 255, 120151. [Google Scholar] [CrossRef]
  190. Roy, A.; Coum, A.; Marinescu, V.D.; Põlajeva, J.; Smits, A.; Nelander, S.; Uhrbom, L.; Westermark, B.; Forsberg-Nilsson, K.; Pontén, F.; et al. Glioma-derived plasminogen activator inhibitor-1 (PAI-1) regulates the recruitment of LRP1 positive mast cells. Oncotarget 2015, 6, 23647–23661. [Google Scholar] [CrossRef] [Green Version]
  191. Kang, H.S.; Kim, J.; Lee, H.J.; Kwon, B.M.; Lee, D.K.; Hong, S.H. LRP1-dependent pepsin clearance induced by 2’-hydroxycinnamaldehyde attenuates breast cancer cell invasion. Int. J. Biochem. Cell Biol. 2014, 53, 15–23. [Google Scholar] [CrossRef] [PubMed]
  192. Li, S.; Porcel, E.; Remita, H.; Marco, S.; Réfrégiers, M.; Dutertre, M.; Confalonieri, F.; Lacombe, S. Platinum nanoparticles: An exquisite tool to overcome radioresistance. Cancer Nanotechnol. 2017, 8, 4. [Google Scholar] [CrossRef]
  193. Luqmani, Y.A. Mechanisms of drug resistance in cancer chemotherapy. Med. Princ. Pract. Int. J. Kuwait Univ. Health Sci. Cent. 2005, 14 (Suppl. 1), 35–48. [Google Scholar] [CrossRef]
  194. Rohwer, N.; Cramer, T. Hypoxia-mediated drug resistance: Novel insights on the functional interaction of HIFs and cell death pathways. Drug Resist. Updates 2011, 14, 191–201. [Google Scholar] [CrossRef]
  195. McAleese, C.E.; Choudhury, C.; Butcher, N.J.; Minchin, R.F. Hypoxia-mediated drug resistance in breast cancers. Cancer Lett. 2021, 502, 189–199. [Google Scholar] [CrossRef]
  196. Wei, G.; Yang, G.; Wei, B.; Wang, Y.; Zhou, S. Near-infrared light switching nitric oxide nanoemitter for triple-combination therapy of multidrug resistant cancer. Acta Biomater. 2019, 100, 365–377. [Google Scholar] [CrossRef]
  197. Zhao, Y.; Ouyang, X.; Peng, Y.; Peng, S. Stimuli Responsive Nitric Oxide-Based Nanomedicine for Synergistic Therapy. Pharmaceutics 2021, 13, 1917. [Google Scholar] [CrossRef]
  198. Huang, X.; Xu, F.; Hou, H.; Hou, J.; Wang, Y.; Zhou, S. Stimuli-responsive nitric oxide generator for light-triggered synergistic cancer photothermal/gas therapy. Nano Res. 2019, 12, 1361–1370. [Google Scholar] [CrossRef]
  199. Bekeschus, S.; von Woedtke, T.; Emmert, S.; Schmidt, A. Medical gas plasma-stimulated wound healing: Evidence and mechanisms. Redox Biol. 2021, 46, 102116. [Google Scholar] [CrossRef]
  200. Zhang, F.; Liu, S.; Zhang, N.; Kuang, Y.; Li, W.; Gai, S.; He, F.; Gulzar, A.; Yang, P. X-ray-triggered NO-released Bi–SNO nanoparticles: All-in-one nano-radiosensitizer with photothermal/gas therapy for enhanced radiotherapy. Nanoscale 2020, 12, 19293–19307. [Google Scholar] [CrossRef]
  201. Akhondzadeh, S. The Importance of Clinical Trials in Drug Development. Avicenna J. Med. Biotechnol. 2016, 8, 151. [Google Scholar]
  202. Bonvalot, S.; Le Pechoux, C.; De Baere, T.; Kantor, G.; Buy, X.; Stoeckle, E.; Terrier, P.; Sargos, P.; Coindre, J.M.; Lassau, N.; et al. First-in-Human Study Testing a New Radioenhancer Using Nanoparticles (NBTXR3) Activated by Radiation Therapy in Patients with Locally Advanced Soft Tissue Sarcomas. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2017, 23, 908–917. [Google Scholar] [CrossRef] [Green Version]
  203. Bonvalot, S.; Rutkowski, P.L.; Thariat, J.; Carrère, S.; Ducassou, A.; Sunyach, M.P.; Agoston, P.; Hong, A.; Mervoyer, A.; Rastrelli, M.; et al. NBTXR3, a first-in-class radioenhancer hafnium oxide nanoparticle, plus radiotherapy versus radiotherapy alone in patients with locally advanced soft-tissue sarcoma (Act.In.Sarc): A multicentre, phase 2-3, randomised, controlled trial. Lancet Oncol. 2019, 20, 1148–1159. [Google Scholar] [CrossRef]
  204. Jin, J.; Zhao, Q. Engineering nanoparticles to reprogram radiotherapy and immunotherapy: Recent advances and future challenges. J. Nanobiotechnol. 2020, 18, 75. [Google Scholar] [CrossRef]
  205. Ngwa, W.; Boateng, F.; Kumar, R.; Irvine, D.J.; Formenti, S.; Ngoma, T.; Herskind, C.; Veldwijk, M.R.; Hildenbrand, G.L.; Hausmann, M.; et al. Smart Radiation Therapy Biomaterials. Int. J. Radiat. Oncol. Biol. Phys. 2017, 97, 624–637. [Google Scholar] [CrossRef] [Green Version]
  206. Ngwa, W.; Kumar, R.; Moreau, M.; Dabney, R.; Herman, A. Nanoparticle Drones to Target Lung Cancer with Radiosensitizers and Cannabinoids. Front. Oncol. 2017, 7, 208. [Google Scholar] [CrossRef] [Green Version]
  207. Ngwa, W.; Kumar, R.; Sridhar, S.; Korideck, H.; Zygmanski, P.; Cormack, R.A.; Berbeco, R.; Makrigiorgos, G.M. Targeted radiotherapy with gold nanoparticles: Current status and future perspectives. Nanomedicine 2014, 9, 1063–1082. [Google Scholar] [CrossRef] [Green Version]
  208. Pottier, A.; Borghi, E.; Levy, L. The future of nanosized radiation enhancers. Br. J. Radiol. 2015, 88, 20150171. [Google Scholar] [CrossRef]
  209. Boateng, F.; Ngwa, W. Delivery of Nanoparticle-Based Radiosensitizers for Radiotherapy Applications. Int. J. Mol. Sci. 2020, 21, 273. [Google Scholar] [CrossRef] [Green Version]
  210. Sinha, N.; Cifter, G.; Sajo, E.; Kumar, R.; Sridhar, S.; Nguyen, P.L.; Cormack, R.A.; Makrigiorgos, G.M.; Ngwa, W. Brachytherapy Application with In Situ Dose Painting Administered by Gold Nanoparticle Eluters. Int. J. Radiat. Oncol. Biol. Phys. 2015, 91, 385–392. [Google Scholar] [CrossRef] [Green Version]
  211. Boateng, F.; Ngwa, W. Novel bioerodable eluting-spacers for radiotherapy applications with in situ dose painting. Br. J. Radiol. 2019, 92, 20180745. [Google Scholar] [CrossRef]
  212. Moreau, M.; Yasmin-Karim, S.; Kunjachan, S.; Sinha, N.; Gremse, F.; Kumar, R.; Chow, K.F.; Ngwa, W. Priming the Abscopal Effect Using Multifunctional Smart Radiotherapy Biomaterials Loaded with Immunoadjuvants. Front. Oncol. 2018, 8, 56. [Google Scholar] [CrossRef] [Green Version]
  213. Miller, S.M.; Wang, A.Z. Nanomedicine in chemoradiation. Ther. Deliv. 2013, 4, 239–250. [Google Scholar] [CrossRef] [Green Version]
  214. Kvols, L.K. Radiation sensitizers: A selective review of molecules targeting DNA and non-DNA targets. J. Nucl. Med. Off. Publ. Soc. Nucl. Med. 2005, 46 (Suppl. 1), 187s–190s. [Google Scholar]
Figure 1. NP interaction with ionizing radiation. (A) The collision of incident photons with inner orbital electrons causes a photoelectric effect. The inner electron absorbs photon energy and ejects it as a photoelectron. Due to the ejection of photoelectrons, electrons from outer shell fill the gap resulting in the emitting of X-rays or the Auger electron. (B) On the other hand, the collision of incident photons with outer orbital electrons causes the Compton effect. The Compton electron absorbing photon energy is ejected from the atom, which may excite and ionize subsequent atoms. The photon loses a fraction of its energy and either continues on its course or is alternatively involved in further process, such as photoelectric effects or the Compton effect.
Figure 1. NP interaction with ionizing radiation. (A) The collision of incident photons with inner orbital electrons causes a photoelectric effect. The inner electron absorbs photon energy and ejects it as a photoelectron. Due to the ejection of photoelectrons, electrons from outer shell fill the gap resulting in the emitting of X-rays or the Auger electron. (B) On the other hand, the collision of incident photons with outer orbital electrons causes the Compton effect. The Compton electron absorbing photon energy is ejected from the atom, which may excite and ionize subsequent atoms. The photon loses a fraction of its energy and either continues on its course or is alternatively involved in further process, such as photoelectric effects or the Compton effect.
Cancers 15 01892 g001
Figure 2. Biological responses of NPs aiding RT. Ionizing radiation causes direct damage to DNA, which is repaired by the cells’ own repair mechanism. However, NPs with RT interfere with the DNA-repair process, leading to cancer cell death. Moreover, ionization induces ROS generation, which damage DNA-killing cancer cells. In addition, radiation together with NPs arrests the G2/M phase of the cell cycle and results in cancer cells death eventually.
Figure 2. Biological responses of NPs aiding RT. Ionizing radiation causes direct damage to DNA, which is repaired by the cells’ own repair mechanism. However, NPs with RT interfere with the DNA-repair process, leading to cancer cell death. Moreover, ionization induces ROS generation, which damage DNA-killing cancer cells. In addition, radiation together with NPs arrests the G2/M phase of the cell cycle and results in cancer cells death eventually.
Cancers 15 01892 g002
Figure 3. Radiation-induced bystander effects. (a) Irradiating cancer cells causes both targeted and non-targeted cell destructions. (b) Radiation along with NPs only kills tumor tissues, leaving the nearby healthy cells intact.
Figure 3. Radiation-induced bystander effects. (a) Irradiating cancer cells causes both targeted and non-targeted cell destructions. (b) Radiation along with NPs only kills tumor tissues, leaving the nearby healthy cells intact.
Cancers 15 01892 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

Haque, M.; Shakil, M.S.; Mahmud, K.M. The Promise of Nanoparticles-Based Radiotherapy in Cancer Treatment. Cancers 2023, 15, 1892. https://doi.org/10.3390/cancers15061892

AMA Style

Haque M, Shakil MS, Mahmud KM. The Promise of Nanoparticles-Based Radiotherapy in Cancer Treatment. Cancers. 2023; 15(6):1892. https://doi.org/10.3390/cancers15061892

Chicago/Turabian Style

Haque, Munima, Md Salman Shakil, and Kazi Mustafa Mahmud. 2023. "The Promise of Nanoparticles-Based Radiotherapy in Cancer Treatment" Cancers 15, no. 6: 1892. https://doi.org/10.3390/cancers15061892

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