Skip to main content
Springer logoLink to Springer
. 2025 Jun 11;53(1):30–47. doi: 10.1007/s00259-025-07390-0

Targeted alpha therapy: a comprehensive analysis of the biological effects from “local-regional-systemic” dimensions

Zhiling Song 1,2,#, Jiajia Zhang 1,2,#, Shanshan Qin 1,2,#, Xiaohui Luan 1,2, Han Zhang 1,2, Mengdie Yang 1,2, Yao Jin 1,2, Gang Yang 1,2, Fei Yu 1,2,
PMCID: PMC12660367  PMID: 40498155

Abstract

Targeted alpha therapy (TAT) has emerged as a promising radiopharmaceutical modality in precision oncology. Compared to beta-emitters, alpha-emitters exhibit superior properties, including higher linear energy transfer, shorter penetration range, enhanced resistance to hypoxic conditions, and convenient radiation protection. Notably, alpha-emitters also demonstrate therapeutic efficacy against a subset of tumors exhibiting resistance to beta-radiotherapy. In 2013, the first α-particle therapeutic agent, ²²³RaCl₂ (Xofigo®), was approved by the FDA for treating bone metastases in advanced castration-resistant prostate cancer, marking a milestone in clinical translation of alpha-emitters. However, the biological mechanisms underlying alpha-particle-mediated therapeutic effects remain incompletely elucidated, which has hindered the optimization of precision treatment strategies. This review systematically analyzes TAT’s tripartite antitumor mechanisms—targeted effects, bystander effects and abscopal effects—thereby constructing a “local-regional-systemic” multidimensional antitumor network. This framework not only clarifies the radiobiological principles of α-emitters but also provides innovative perspectives for advancing TAT applications in tumor precision therapy.

Graphical abstract

graphic file with name 259_2025_7390_Figa_HTML.jpg

Keywords: Targeted Alpha Therapy (TAT), Immunotherapy, Targeted effects, Bystander effects, Abscopal effects

Introduction

Recently, targeted alpha therapy has advanced greatly in tumor therapy. Table 1 provides a systematic summary of the commonly used medical alpha-emitters. Compared with beta-emitting radionuclides, alpha-emitting radionuclides have shown potential advantages. The alpha particles emitted by alpha-nuclides are doubly charged helium-4 nuclei (Inline graphic He2+) characterized by substantial mass and high energy levels (4–9 MeV). These particles exhibit remarkably higher linear energy transfer (LET) than beta particles, with LET values ranging from 50 to 230 keV/µm, in contrast to the 0.2 keV/µm LET typical of beta radiation. This high LET characteristic enables a single alpha particle trajectory to induce irreparable DNA double-strand breaks (DSB). In comparison, beta particle irradiation predominantly causes single-strand breaks (SSB), which are generally reparable, resulting in cytotoxic potency approximately 500-fold lower than that of alpha particles [1]. Moreover, since stronger interactions with matter lead to a shorter path, the substantial mass and charge of alpha particles cause them to interact extensively with the surrounding medium, limiting their range in tissue to about 45–100 μm—roughly the diameter of 5 to 10 cells—thereby effectively reducing damage to neighboring healthy cells compared to the longer range of beta particles [2]. It is worth noting that the short range of alpha-emitters may result in inferior cross-fire effects compared to beta-emitters, leading to suboptimal efficacy in large masses or antigen heterogeneous tumors [3]. Furthermore, the cytotoxicity induced by alpha-emitters is independent of oxygen free radicals generated by radiolysis of water and unaffected by cell cycle phases. The underlying mechanism lies in the dominant direct DSB induced by alpha particles. Significantly, the hypoxia tolerance of alpha-emitters confers significant advantages in treating radiation-resistant refractory tumors due to hypoxic tumor microenvironments [47]. TAT can also effectively overcome resistance issues associated with beta-particle therapies, driving a paradigm shift in radionuclide treatment approaches [8] (Fig. 1). With the aforementioned radiobiological advantages, alpha-emitters are gradually transitioning from preclinical trials to clinical applications. Currently, Several clinical trials based on alpha-emitters have been approved (Table 2), marking an advancement in the clinical translation of this therapeutic strategy.

Table 1.

Common characteristics, delivery strategies, and clinical applications of alpha-emitters

Alpha emitter Half-life Range (um) Emissions Per Decay Max Energy (MeV) Delivery strategies Current clinical applications
223Ra 11.4 d 46 ~ 70 4 α, 2β- 5.716 Natural bone-targeting properties mCRPC
225Ac 9.92 d 50 ~ 90 4 α, 2β- 5.830 Monoclonal antibody, peptides, bispecific antibody, small molecules,

mCRPC, SCLC,

LCNEC GEP-Nets, advanced solid tumours, glioma

211At 7.21 h 55 ~ 80 1 α, 1 EC 5.869 Monoclonal antibody

Acute leukemia,

myelodysplastic syndrome

212Pb 10.64 h 600 1 α, 2β- 6.09 Small molecules, peptides

mCRPC,

solid tumors,

neuroendocrine tumors,

melanoma,

atypical lung carcinoids

212Bi 60.6 min 40 ~ 100 1 α, 1β- 6.051 / /
213Bi 45.6 min 40 ~ 100 1 α, 2β- 5.875 Monoclonal antibody

Leukemia,

myelodysplastic syndrome

149Tb 4.1 h 25 ~ 30 1 α, 1 β+ 3.967 / /
227Th 18.7 d 50 ~ 70 5 α, 2β- 6.038 Antibody mCRPC

Fig. 1.

Fig. 1

Comparison of the characteristics of alpha particles and beta particles. (Created with Biorender.com)

Table 2.

Current clinical trials of targeted alpha therapy are underway

Clinical trial Alpha-emitter Target Drug Approved indications Primary outcome measures
NCT03746431 225Ac IGF-1R 225Ac-FPI-1434 Locally advanced or metastatic solid tumours AE, DLT, ORR
NCT05902247 225Ac PSMA 225Ac-PSMA I&T mCRPC AE, SAE, Safety, Tolerability
NCT04946370 225Ac PSMA 225Ac-J591 mCRPC DLT, Optimal dose, CR
NCT04597411 225Ac PSMA 225Ac-PSMA-617 CRPC RP2D
NCT05496686 225Ac / 225Ac-MTI-201 Metastatic uveal melanoma MTD, DLT, AE, SAE
NCT06732505 225Ac SSTR 225Ac-DOTATATE Inoperable, locally advanced or metastatic, progressive, well-differentiated, SSTR + GEP-Nets Safety, Tolerability, AE, DLT
NCT06147037 225Ac EGFR-cMET 225Ac-FPI-2068 Metastatic and/or recurrent solid tumors Safety, Tolerability, AE, MTD, RP2D
NCT06975332 225Ac NK-1R 225Ac-DOTA-SP Recurrent glioblastoma (WHO G3-G4) Clinical progression
NCT06492122 225Ac PSMA 225Ac-FL-020 mCRPC DLT, RP2D, AE, SAE
NCT06736418 225Ac DLL3 225Ac-ABD147 SCLC and LCNEC of the lung following platinum-based chemotherapy Safety, AE, Tolerability, DLT, RP2D, ORR, DCR, DOR, PFS, OS
NCT06939036 225Ac SSTR2 225Ac-SS0110 ES-SCLC and recurrent locally advanced or metastatic MCC Safety, Tolerability, RP2D, SAE, PFS, DCR, BOR, ORR, DoR
NCT05636618 212Pb SSTR2 212Pb-ADVC001 Advanced SSTR2 positive neuroendocrine tumors DLT, AE, AUC
NCT05655312 212Pb MC1R [212Pb]VMT01 Unresectable and metastatic melanoma DLT, ORR, AE, SAE
NCT06427798 212Pb SSTR [212Pb]VMT-alpha-NET Gastrointestinal neuroendocrine tumors MTD, ORR
NCT05283330 212Pb GRPR 212Pb-DOTAM-GRPR1 Recurrent or metastatic GRPR-expressing Tumors RP2D
NCT03670966 211At CD45 211At-BC8-B10 Relapsed or refractory high-risk acute leukemia or myelodysplastic syndrome Toxicity

mCRPC metastatic castration-resistant prostate cancer

While the inherent radiobiological advantages of alpha-emitters establish a foundation for precision therapy, the multifaceted mechanisms of TAT remain incompletely elucidated. Current research primarily focuses on alpha particle-induced nuclear DNA damage as the dominant on-target effect [9]. However, emerging evidence has increasingly revealed TAT-mediated subcellular structures targeting, although it’s still not comprehensive [10]. Significantly, beyond these direct cytotoxic actions, TAT exerts cytotoxic effects on non-irradiated tumor cells through off-target mechanisms. These effects are predominantly orchestrated by the radiation bystander effect and systemic abscopal immune activation, both critically contributing to therapeutic efficacy. This review systematically synthesizes recent advances in TAT’s biological mechanisms, integrating organelle-specific targeting with systemic off-target regulatory networks to decode its multi-layered tumor eradication principles, thereby providing a theoretical framework for optimizing TAT clinical translation strategies.

Targeted effects

DNA centered effects

For decades, nuclear DNA has been regarded as the primary target of irradiation [11]. DNA damage induced by radionuclides occurs through both direct and indirect mechanisms. Among them, the radiolysis of water produces reactive oxygen species (ROS) including hydroxyl radicals (•OH), superoxide anions (O₂•⁻), and hydrogen peroxide (H₂O₂) [12], a process commonly referred to as the “indirect effect” of radiation. In addition, charged particles can directly ionize biomolecules (DNA, proteins, and lipids). This direct effect is particularly significant in high LET particles (such as α-particles), as alpha particles induce localized high-density ionization in biological tissues. Unlike beta particles that primarily induce repairable SSB in tumor cells through radical-mediated mechanisms, alpha particles directly cause complex, irreparable clustered DSB. This highly efficient mode of DNA damage makes alpha particles uniquely advantageous in radiotherapy [4, 13].

Generally, DNA damage triggers the DNA damage response pathway, which safeguards cell viability and preserves genomic integrity by activating DNA repair mechanisms [14].The repair mechanisms for alpha particle-induced DSBs relies on two distinct molecular pathways: Non-Homologous End Joining (NHEJ) and Homologous Recombination (HR) [15]. NHEJ mediates rapid DNA restoration through direct ligation of broken DNA termini, offering immediate damage resolution at the cost of reduced fidelity, often resulting in accumulation of insertion/deletion mutations. In contrast, HR operates as a high-precision repair system strictly confined to S and G2 phases of the cell cycle, requiring a homologous template [16]. Notably, high-LET alpha particles generate complex DNA lesions that are particularly challenging for cellular repair systems due to their dense ionization patterns.

The high LET properties of alpha particles, arising from their elevated ionization density and localized energy deposition, induce more complex biological damage compared to low-LET radiation (β particles and γ rays) [17]. Such radiation not only generates DNA DSBs but also causes SSBs, base oxidation, and clustered lesions. During damage recognition, the MRE11-RAD50-NBS1 (MRN complex) localizes to DNA damage sites [18], facilitating the recruitment and activation of serine/threonine protein kinases (ATM/ATR). This triggers phosphorylation cascades in key signaling molecules such as Chk2, γ-H2 AX, and p53 [19, 20]. In response to DNA damage, cells can activate repair mechanisms such as HR through the recruitment of repair proteins like 53BP1 and BRCA1 [21, 22]. Simultaneously, cell cycle arrest mediated by proteins such as p21 and Chk1 provides critical time for DNA repair, thereby preventing abnormal replication of damaged DNA [23]. When damage exceeds reparable thresholds, the p53 signaling network shifts to activate a pro-apoptotic program by upregulating effector molecules like Bax, Puma, Noxa, and Fas, driving cells toward irreversible apoptosis [24, 25]. Marques et al. have elucidated that 223RaCl2 induces irreparable DNA double-strand breaks, which activates cell cycle checkpoint through the ATM/CHK2 signaling pathway, consequently triggering G2/M phase arrest and initiating apoptotic cascades [26]. This regulatory mechanism, which coordinates DNA repair and programmed cell death in accordance with the biological relevance of DNA damage [27], establishes a homeostatic equilibrium between genome integrity maintenance and the elimination of genomically compromised cells (Fig. 2). While existing review articles have systematically elucidated the mechanisms underlying ionizing radiation-induced DSBs and their dynamic coordination with cell cycle progression and DNA repair processes, the precise molecular mechanisms underlying TAT regulation remain incompletely elucidated [28].

Fig. 2.

Fig. 2

A concept map of TAT-induced targeted effects. (Created with Biorender.com). The direct ionization by alpha particles induces clustered DNA DSBs, which recruit and activate the ATM/ATR kinase cascade, initiating DNA damage response signaling network. The extent of damage dynamically regulates downstream pathways to determine cellular fate: mild damage activates cell cycle arrest to facilitate DNA repair, while severe damage triggers irreversible cell death via the intrinsic apoptotic pathway. Furthermore, α-particle irradiation specifically activates ASMase, catalyzing sphingomyelin hydrolysis to generate ceramide, which aggregates into lipid rafts, thereby activating the MAPK signaling cascade (e.g., p38/JNK/ERK phosphorylation) to amplify pro-apoptotic signals. The ionization of α-particles induces mitochondrial membrane potential collapse, triggering mitochondrial dysfunction and subsequent calcium overload. Moreover, α-particles provokes endoplasmic reticulum stress and initiates lysosomal-mediated autophagy, forming a coordinated chain of cellular damage responses

It is noteworthy that studies on alpha-particle-based radiopharmaceuticals provide critical insights into elucidating the aforementioned mechanisms. A 213Bi-labeled PSMA inhibitor (²¹³Bi-PSMA I&T) demonstrated rapid tumor-targeting kinetics, with its alpha-particle emissions inducing persistent accumulation of DNA DSBs in tumor tissues [29]. Similarly, the 211At-labeled heterodimeric peptide (iRGD-C6-lys(211At-ATE)-C6-DA7R) has been demonstrated to significantly inhibit the viability of glioma cells. This compound effectively induces cellular apoptosis and triggers cell cycle arrest at the G2/M phase [30]. Recent research has demonstrated that 223RaCl2 induces DNA damage and activates repair pathways in various prostate cancer cells. The finding revealed that it activates DSB repair via the NHEJ mechanism, while simultaneously inducing high levels of apoptosis. The potent cytotoxic effect of 223Ra may be attributed to alpha-particle-induced clustered DNA damage sites [31].

Non-DNA centered effects

Cell membrane

DNA is not the only target of radiation-induced damage. The cell membrane, a dynamic bilayer structure, is composed of lipids (30–80%), proteins (20–60%), and carbohydrates (0–10%). The lipids in the cell membrane contain polyunsaturated fatty acids (PUFA), which are susceptible to oxidation by ROS induced by radiation. This oxidation process generates lipid-derived metabolites, such as malondialdehyde (MDA) and 4-hydroxy-2-nonenal (4-HNE). These breakdown products serve as biomarkers of lipid peroxidation, indicating that irradiation has disrupted the biological functions of the cell membrane [32].

Alpha particle-induced ROS can activate acid sphingomyelinase (ASMase) [33, 34], which hydrolyzes sphingomyelin, a type of membrane phospholipid, to produce ceramide. Ceramide accumulation leads to the formation of large signaling platforms known as lipid rafts. The formation of lipid rafts can activate a variety of signaling pathways [35, 36] and ion channels (such as potassium [37] and calcium [38] channels), ultimately resulting in nuclear DNA damage (Fig. 2).

Several studies have revealed the mediating role of the cell membrane in alpha particle-induced cell death. Seideman et al. indicated that alpha particles released by 225Ac-DOTA-anti-CD3 IgG antibodies can induce dose-dependent apoptosis through activation of the sphingomyelin pathway [39]. Ladjohounlou and his colleagues also demonstrated that the cell membrane serves as a radiosensitive target for radiation-induced damage through their studies using antibodies labeled with alpha-emitting radionuclides (²¹³Bi, ²¹²Pb/²¹²Bi) [40]. Their research further revealed that alpha radioimmunotherapy can induce cell death by activating p38 and Jun N-terminal kinase 1/2 (JNK1/2) signaling pathways through lipid raft-mediated mechanisms. Disrupting the integrity of lipid rafts leads to the absence of p38 and JNK1/2/3 phosphorylation. When pharmacological inhibitors of p38 (SB203580) and JNK1/2/3 (SP600125) were employed in conjunction with alpha particle-irradiated cells, an enhancement in cell viability was observed. This finding suggests that the p38 and JNK mitogen-activated protein kinase (MAPK) signaling pathways are pivotal in the targeted cytotoxicity of alpha-particle irradiation [41]. Additionally, signaling pathways such as NF-κB and PI3 K/AKT have been shown to be activated [36].

Mitochondria

In the cytoplasm, mitochondria, which play a crucial role in cellular metabolism and energy homeostasis, occupy 25% of the cell volume, indicating a high probability of being hit by particles [42]. Mitochondrial DNA, being unprotected by histones, is often observed to have mutations or deletions under radiation exposure. Schilling-Toth et al. elucidated this phenomenon as radiation-induced mitochondrial genomic instability [43, 44]. Notably, intact mitochondrial function is essential for the cells to induce nuclear signaling responses upon cytoplasmic-specific targeting interventions [22].

Although numerous studies have demonstrated that radiation induces mitochondrial DNA damage, the specific mechanisms by which mitochondrial function contributes to radiation-induced cytoplasmic toxicity remain unclear. This is primarily due to the challenge of selectively targeting the cytoplasm without affecting the nucleus. Radiation exposure leads to a sustained increase in mitochondrial oxidative stress, this change is associated with mitochondrial membrane potential, mitochondrial respiration, and the ATPase encoded by the mitochondria [43]. Bo Zhang et al. reported that targeted cytoplasmic irradiation with alpha particles can specifically induce upregulation of DRP1 expression, thereby driving mitochondrial fission, accompanied by impaired mitochondrial respiratory chain function (such as a significant reduction in the activities of cytochrome c oxidase and succinate dehydrogenase within 4 h after irradiation) [45]. It is worth noting that mitochondrial fission gradually recovers to a nearly normal morphology within 24 h, while the respiratory chain function partially recovers. This dynamic change may reflect the adaptive stress response of mitochondria through the fission-fusion balance. Treatment with the DRP1 inhibitor mdivi-1 can block radiation-induced mitochondrial fission but fails to reverse the impairment of respiratory chain function. This suggests that DRP1-dependent fission may indirectly exacerbate functional defects by limiting the clearance of damaged mitochondria. This finding reveals the complex association between mitochondrial dynamics and function, indicating that DRP1-mediated fission is a compensatory stress adaptation mechanism after alpha-particle irradiation. However, excessive mitochondrial fission can lead to reduced respiratory chain function and irreversible mitochondrial damage. Evidence suggests that aberrant mitochondrial fission contributes to impaired respiratory chain activity in neurodegenerative diseases such as Alzheimer’s and Parkinson’s [46, 47]. Furthermore, study had demonstrated that under alpha-particle irradiation, the transcriptional expression of mitochondrial biomarkers such as LONP1 and TFAM, as well as mtDNA-encoded genes MT-CYB and MT-RNR1, was significantly upregulated during the differentiation of human induced pluripotent stem cell-derived cardiomyocytes, which may represent an additional adaptive response mechanism to radiation stress [48]. Yang et al. developed a novel ²²³Ra-loaded Mg/Al-layered double hydroxide (LDH) nanomaterials (223Ra-LDH) through a coprecipitation strategy. This radiopharmaceutical delivery system induces mitochondrial dysfunction via ROS generation and stimulates endogenous Ca²⁺ release. The resultant Ca²⁺ overload ultimately triggers apoptosis, demonstrating precise ²²³Ra delivery within the tumor microenvironment [49].

Other subcellular structures

Current research primarily focuses on these established radiobiological targets, while investigations into the functional mechanisms of subcellular structures such as the endoplasmic reticulum (ER) and lysosomal system in alpha-particle therapy remain limited. Although studies have observed damage responses in these organelles following alpha-irradiation, their underlying molecular regulatory mechanisms have yet to be systematically elucidated. Diniz Filho et al. demonstrated that ²²³RaCl₂ irradiation of MDA-MB-231 cells triggered canonical autophagy pathway activation, characterized by autophagosome-lysosome formation, which represents one of the principal mechanisms underlying radiation-induced cell death [50]. Furthermore, transmission electron microscopy (TEM) revealed ultrastructural alterations, including dilation of both ER and mitochondria following alpha-particle exposure [51]. A recent study demonstrated that ²²³Ra induces ER stress in tumor cells, which markedly upregulates calreticulin (CRT) expression and facilitates its ectopic surface translocation, thereby potentiating susceptibility to cytotoxic T lymphocyte (CTL)-mediated cytolysis [52]. Functioning as both a canonical ER chaperone and pivotal damage-associated molecular patterns (DAMP) [53], CRT plays a pivotal role in radiation-induced immunogenic modulation. This finding elucidated the pivotal role of ER stress in ²²³Ra-mediated immune modulation and established a theoretical framework for developing ER-focused TAT strategies. Besides, research demonstrated that monoclonal antibodies (mAbs) labeled with alpha-emitters (such as 212Bi, 213Bi, and 212Pb) can migrate from the cell surface to lysosomes, accompanied by the gradual accumulation of alpha-emitting radionuclides within lysosomes, thereby ensuring optimal cellular radiation exposure. This discovery provides valuable insights for developing lysosome-targeted strategies based on alpha-emitters [54, 55].

Bystander effects

In addition to the direct cytotoxicity of TAT, alpha-emitters also cause biological effects through non-targeted effects (NTEs), which commonly include bystander effects and abscopal effects induced by immune response activation. Numerous studies have demonstrated that non-irradiated cells can exhibit a series of biological responses when their neighboring cells are exposed to radiation, the phenomenon known as bystander effect. In the 2006 report by the United Nations Scientific Committee on the Effects of Atomic Radiation (UNSCEAR), the term “bystander effect” is defined as “the ability of irradiated cells to convey manifestations of damage to neighboring cells not directly irradiated” [56]. Notably, research has confirmed a close correlation between this bystander effect and the hyper-radiosensitivity of cells to low-dose radiation exposure [57].

In 1992, Nagasawa and Little performed a pioneering study that first revealed bystander effect induced by alpha-radiation. They exposed Chinese hamster ovary cells to low doses of alpha-particle irradiation (0.31 mGy), where only a small fraction of cell nuclei (1% or less) were traversed by alpha particles, yet 30% of the cells demonstrated an increased frequency of sister chromatid exchanges. Under normal conditions, approximately 2.0 Gy of X-ray dose is typically required to achieve a similar biological effect [58]. This study highlighted the potential value of radiation-induced bystander effect (RIBE) in reducing radiation dosage.

Experimental evidence

In recent years, research on alpha-emitting radionuclides in the field of RIBE has been continuously advancing. The study revealed that when lymphocytes were irradiated with 213Bi-labeled monoclonal antibodies, the percentage of cell mortality was significantly higher than the percentage of nuclei actually hit by alpha particles, indicating the presence of a pronounced bystander effect [59]. Leung et al. provided evidence through the establishment of an in vivo model that 223Ra-induced cytotoxic bystander effects are involved in suppressing the growth of disseminated tumor cell xenografts. Their findings demonstrate the therapeutic potential of alpha-emitters in disseminated tumor cells [60]. Subsequently, the team achieved stratification of the direct effects of alpha particles and radiation-induced bystander effects through in vivo measurements of the biological effects of alpha-particles emitted by 223RaCl₂, further highlighting the critical regulatory role of bystander effects in enhancing the overall therapeutic response [61]. Moreover, Rajon et al. developed a model to investigate the role of the bystander effect induced by 223Ra in delaying the growth of breast cancer xenografts in mice [62]. These studies contribute to understanding the mechanisms by which 223Ra slows the progression of metastatic lesions. Other researchers have used 212Pb/212Bi-labeled monoclonal antibodies to reveal the roles of ROS and cell membranes in alpha-particle-emitter-induced bystander effects [40]. Additionally, Yang et al. focused on the biological role of mitochondrial transfer in RIBE, revealing that damaged mitochondria mediate alpha-particle-induced bystander effects through intercellular transmission, though the precise molecular pathways remain to be elucidated [63].

These findings not only reveal the existence of bystander effects induced by alpha-emitter irradiation, but also highlight the indispensable role of intercellular communication in mediating this phenomenon, as cells can be killed without being directly traversed by particles. With the continuous advancement of research in recent years, both direct cell-to-cell communication via gap junctions and the release of signaling molecules have been identified as two primary mechanisms involved in bystander effects (Fig. 3).

Fig. 3.

Fig. 3

RIBE is mediated by multiple mechanisms of intercellular communication. (Created with Biorender.com). The GJIC formed by connexins, which directly transmit oxidative stress signals to bystander cells. Irradiated cells can also signal through the secretion of soluble mediators, or by transporting extracellular vesicles that carry miRNAs and damaged organelles like mitochondria and lysosomes, leading to bystander cell death. These mechanisms collectively result in genomic instability, reduced clonogenic capacity, and non-targeted cell death in bystander cells

Gap junction intercellular communication

Early studies primarily focused on the role of gap junction intercellular communication (GJIC) mediated by direct cell-cell contact in radiation-induced bystander effects [64]. Azzam et al. first identified the critical function of GJIC in bystander signal transmission between cells in 1998 [65], and their follow-up study in 2001 further demonstrated that connexin43 (CX43), as a core regulatory molecule of GJIC, plays an essential mediating role in intercellular communication. Of particular significance, pharmacological inhibition of GJIC using γ-hexachlorocyclohexane (lindane) resulted in a marked reduction of alpha-particle-induced bystander effects, directly confirming the necessity of GJIC in this process [66]. Subsequent research by the Autsavapromporn team expanded the understanding in this field, revealing that GJIC not only transmit radiation-induced damage signals but also amplify oxidative stress-related signaling pathways, thereby serving as central regulators in the cascade reactions of radiation bystander effects [67].

However, RIBE is mediated not only by gap junction channels but also through extracellular mediator-mediated communication, including soluble mediator secretion and vesicle-based molecular transport.

Extracellular mediator-mediated communication

Soluble mediators

Substantial evidence from multiple studies has demonstrated that various soluble signaling factors play crucial regulatory roles in RIBE. Such as reactive species [68, 69] (e.g., ROS, NO), cytokines [7072] (e.g., IL-12, IL-15, IL-18, TNF-α, IFN-γ, TGF-β1), ionic signaling mediators [73] (e.g., Ca²⁺).

Among various signaling molecules, ROS and reactive nitrogen species (RNS) play crucial regulatory roles in the bystander effect [36, 74]. Previous study has demonstrated that ROS/RNS are generated in the mitochondria of target cells through a Ca²⁺-dependent mechanism [75]. In RIBE, ROS serve as the initial triggering factor that initiates the cascade reaction, while nitric oxide radical (NO•) and its active metabolites function as key effector molecules, regulating the activation and transduction of downstream signaling pathways [76]. Research data suggest that under alpha-particle radiation conditions, ROS may participate in the activation process of both the p53-dependent pathway and multiple members of MAPK superfamily in bystander cells through the regulation of redox-sensitive signaling pathways, and this mechanism may be associated with micronucleus formation in bystander cells [77]. ROS act as pivotal signaling molecules that induce cyclooxygenase-2 (COX-2) expression [64]. Research has demonstrated that the COX-2 signaling cascade plays a central regulatory role in RIBE. Zhou et al. substantiated that pharmacological inhibition of COX-2 activity using NS-398, a specific COX-2 inhibitor, significantly attenuated bystander responses. Furthermore, given the pivotal role of COX-2 in activating MAPK pathway, this investigation further elucidates the critical involvement of MAPK signaling cascades in bystander effect mechanisms. Notably, COX-2-associated pathways are indispensable for mediating cellular inflammatory responses, suggesting potential molecular crosstalk between inflammatory mechanisms and bystander effects [78]. NO, serving as a core component of RNS, is synthesized and released through enzymatic catalysis by distinct isoforms of nitric oxide synthase (NOS) [64, 79]. Shao et al. demonstrated that NO is involved in mediating the regulatory effects of radiation-induced bystander effects on cell proliferation and micronucleus formation in human salivary gland cells [80].

Significantly, research findings indicate that dimethyl sulfoxide (DMSO), a ROS scavenger, failed to significantly suppress the bystander effect following irradiation, whereas Filipin, a cell membrane signaling inhibitor, exhibited pronounced inhibitory effects. This discovery suggests that beyond signaling molecule mechanisms, membrane-dependent signal transduction pathways may play a more critical role in regulating bystander effects [71].

In vitro experiments demonstrated that disruption of lipid raft structures using Methyl-β-cyclodextrin, filipin, or ROS scavengers significantly suppressed non-targeted cytotoxicity, confirming the critical role of the cell membrane in radiation-induced NTEs [81]. The co-culture experiments utilizing U937 macrophages and bystander cell HL-7702 hepatocytes demonstrated that alpha-radiation triggers a bidirectional cAMP communication mediated through cell membrane signaling pathways. However, this interaction was not observed when the membrane signaling pathway was blocked using filipin, which inhibits cAMP transmission [82]. Further studies have revealed that alpha-particle irradiation activates membrane-dependent signal transduction pathways, such as the MAPK signaling cascade, which plays a central regulatory role in bystander effects. Key components of this pathway include JNK, ERK1/2, and p38 MAPK [38, 77, 83]. Relevant data revealed that lipid raft-dependent mechanisms contribute approximately 50% to the NTEs induced by alpha-particle irradiation, highlighting the importance of membrane architecture in mediating these biological responses [40, 81].

Vesicles

Additionally, radiation-induced bystander signals can be transmitted between cells via extracellular vesicles (EVs), such as exosomes (50–150 nm), the smallest extracellular vesicles [84]. Compared to other soluble factors, exosomes possess a protective lipid bilayer that enables stable systemic propagation without degradation [85, 86]. They can carry bioactive molecules, including nucleic acids, proteins, and lipids, which are delivered to recipient cells through various mechanisms. This dual functionality allows exosomes to mediate both material transfer and signaling in intercellular communication [87]. Studies have shown that radiation enhances the release of exosomes and promotes the migration of recipient cells [8890]. Molecular profiling further reveals that exosomes are enriched with critical signaling pathway molecules associated with cell migration [88]. Furthermore, as a novel mediator of RIBE, exosomes participate in the systemic immunomodulatory effects of radiation. Exosomes released from irradiated cells can deliver immune-active molecules (e.g., MHC-I molecules, DAMPs, and double-stranded DNA) to unirradiated cells, triggering innate immune responses and reshaping local or systemic immune microenvironments [91].

Study indicated that when high-LET radionuclides are used for intracellular irradiation, RIBE is constrained by a threshold radiation dose. Alpha particles and Auger electrons can induce bystander effects at low radiation doses, but the effects diminish at higher doses [92]. Under conditions where direct radiation effects are limited, low average absorbed doses can trigger high bystander effects [93, 94]. This mechanism expands the killing range of tumor cells in heterogeneous regions (such as areas with uneven radiopharmaceutical biodistribution or clusters of drug-resistant clones), providing an innovative “controlled-by-less” strategy to reduce the recurrence risk of residual lesions.

Abscopal effects

In addition to the bystander effect, NTEs can also produce biological effects through abscopal effect. Although the abscopal effect is not the primary target of irradiation, it may significantly enhance the overall therapeutic outcome. In 1905, Heineke et al. first proposed that radiation might transmit signals from the irradiated area to unirradiated regions through the bloodstream. This suggested that an underlying biological mechanism could allow radiation therapy to exert effects beyond the localized treatment area, potentially influencing distant tissues or organs. The term “abscopal effect (AE)” in its true sense was first coined by British scientist Mole in 1953. He observed that when a local tumor was treated with radiation therapy, distant tumors that had not been irradiated also showed signs of shrinkage or regression. He defined the AE as “the gradual regression of distant tumors that were not irradiated as a result of local radiation” [95].

Accumulating evidence now suggests that the abscopal effect is mediated by the immune system [96]. Studies confirmed that systemic immune responses play a crucial role in radiation-induced anti-tumor effects [97, 98]. Ferreira’s team revealed the unique immunomodulatory advantages of alpha-emitters through comparative studies in immunocompetent prostate cancer models. The alpha-emitting radiopharmaceutical 225Ac-NM600 demonstrated significantly stronger antitumor efficacy compared to the beta-emitting radiopharmaceutical 177Lu-NM600. This enhanced therapeutic effect is attributed to the alpha-emitter’s specific reduction of immunosuppressive regulatory T cell (Treg) infiltration, whereas no such remodeling of the immune microenvironment was observed in the 177Lu-NM600 treatment group [99]. Leung et al. intravenously administered 223RaCl₂ at doses of 0, 50, or 600 kBq/kg to Swiss Webster mice and collected spleens on days 5, 12, and 19 to evaluate changes in immune cell populations. The results demonstrated that 223RaCl₂ affected the functional activity of NK cells and led to a reduction in splenic CD8+ T cells in Swiss Webster mice, depending on both administered activity and time post-administration [100]. Current clinical trials have demonstrated the significant therapeutic efficacy of 223Ra in mCRPC patients, and this success has laid the foundation for other radionuclide therapies in mCRPC treatment. In-depth investigation of the dynamic changes in immune microenvironment before and after treatment, as well as the radionuclide-specific immunomodulatory mechanisms, will provide critical insights for optimizing personalized treatment strategies [101].

Radiation-induced immunogenic cell death (ICD) constitutes a pivotal mechanism underlying abscopal effects [102]. Ionizing radiation triggers the release of DAMPs from tumor cells, including key mediators such as high mobility group box 1 (HMGB1), adenosine triphosphate (ATP), and CRT. These DAMP molecules activate pattern recognition receptors, thereby promoting the maturation and differentiation of antigen-presenting cells (APCs). Subsequently, the matured APCs migrate to tumor-draining lymph nodes, where they present processed tumor-associated antigens to T lymphocytes. This antigen presentation cascade ultimately activates tumor-specific CTL-mediated immune responses [103] (Fig. 4).

Fig. 4.

Fig. 4

α-particle–emitting radiopharmaceuticals induce systemic anti-tumor abscopal effects. (Created with Biorender.com). Alpha particle irradiation triggers tumor cells to release DAMPs, including HMGB1, ATP, and CRT. These DAMPs bind to pattern recognition receptors on the surface of APCs, driving the maturation and differentiation of APCs. Mature APCs migrate to tumor-draining lymph nodes, presenting processed tumor-associated antigens to T lymphocytes and activating CTLs. Activated CTLs eliminate tumor cells by infiltrating primary tumor sites and distant metastases through the peripheral circulation, while inducing long-term immune memory formation

Evidence suggested that cellular exposure to alpha irradiation (227Th) induced the production of DAMPs, triggering immunogenic cell death. This indicates the activation of the immune system under alpha irradiation [104]. Previously, the alpha-emitter 213Bi was also reported to produce similar results. Alpha-particle therapy for multiple myeloma using 213Bi can induce irradiated cells to secrete soluble factors that activate dendritic cells (DCs) and trigger the occurrence of ICD in tumor cells [105]. Gorin et al. also observed that treatment of MC-38 cells with 213Bi induced the release of DAMPs and activated DCs, leading to a systemic and lasting anti-tumor response [106]. Furthermore, irradiation therapy using 211At on small cell lung cancer induced the release of MHC-I molecules and CRT from tumor cells. This process activates endogenous anti-tumor immune responses by modulating the functionality of immune cells within the tumor microenvironment [107]. Similarly, 223Ra has been shown to exhibit similar regulatory mechanisms [52]. Moreover, Yang et al. demonstrated that alpha-emitter radiopharmaceuticals (223Ra-LDH) can activate the STING signaling pathway in vivo, eliciting a cascade effect that effectively induces ICD while concomitantly enhancing DCs maturation and T-lymphocyte activation [49]. The above studies demonstrate that TAT not only enhance T-cell cytotoxic capacity by promoting antigen presentation, but also activate antitumor immune responses through inducing tumor cell ICD. It is noteworthy that tumor immunomodulatory effects triggered by different types of alpha-emitter therapies may vary, therefore further in-depth exploration of their mechanisms of action remains necessary. Intriguingly, TAT can further establish long-term immunological memory by promoting the formation of effector memory T cells, thereby effectively preventing tumor recurrence [108].

Multiple clinical studies evaluating the immunomodulatory effects of alpha-emitters are currently underway. In 2019, a case of multiple cutaneous squamous cell carcinoma was reported, where complete remission of the treated lesion was observed 76 days after implantation of 224Ra-loaded seeds. Two distant lesions showed significant shrinkage. One year after treatment, the treated lesion remained in complete remission, while the untreated distant lesions spontaneously regressed. This may be related to an immune-mediated response [109]. Kim et al. enrolled 15 male patients with metastatic castration-resistant prostate cancer (mCRPC) who received a regimen of 223Ra at 50 kBq/kg. Peripheral blood mononuclear cells were collected before treatment and 3–4 weeks after treatment to analyze the phenotype and functional characteristics of CD8+ T cells. The results showed that after a single 223Ra treatment, the proportion of effector memory CD8+ T cells expressing PD-1 decreased from 20.6 to 14.6% [110]. However, the study had a small sample size, and the significance of the immunological changes and the long-term effects of treatment still require further investigation in future research. Currently, a phase I/II study combining 223Ra with the anti-PD-L1 antibody (atezolizumab) has been completed [111]. Numerous clinical trials exploring the combined use of radionuclide 223Ra and immunomodulatory drugs in the treatment of mCRPC are also actively advancing [112115]. Evidently, combination regimens employing different alpha-emitting radionuclides with immunotherapy have exhibited heterogeneous clinical outcomes, while the synergistic effects between TAT and immune checkpoint inhibitors (ICIs) (Table 3) point to a potentially promising treatment approach [116120]. Existing clinical studies have reported that the combined administration is typically performed on the same day. If concurrent administration is not tolerated, the radiopharmaceuticals treatment is administered first, followed by immunomodulatory drugs. This sequencing strategy may be attributed to the potential of prior radioligand therapy to prime the tumor microenvironment. However, there is still insufficient clinical evidence or established guidelines to determine whether this specific administration sequence yields superior clinical benefits. Further clinical trials are needed to explore the optimal treatment sequencing strategy [111].

Table 3.

Recent studies on the combination of TAT and ICIs

Alpha radionuclide Combination Tumor type Study type Efficiency
Bismuth-213 [121] Anti-PD-1 mAb (CD279) Melanoma Preclinical study Slow down tumor growth rate by 1.5 times (compared to monotherapy)
Astatine-211 [122] Anti-PD-1 mAb Glioblastoma Preclinical study Significantly prolong PFS and 100% complete response
Astatine-211 [108] Anti-PD-L1 mAb

Breast cancer,

Colorectal cancer

Preclinical study

Prolong overall survival,

Form long-term immune-memory effects

Actinium-225 [123] Anti-PD-1 mAb mCRPC Preclinical study Prolong TTP and survival
Actinium-225 [124] Anti-PD-L1 mAb Melanoma Preclinical study Delay tumor growth
Lead-212 [117]

Anti-PD-1 mAb,

Anti-CLTA-4 mAb

(Dual ICIs)

Melanoma Preclinical study 43% complete tumor response
Thorium-227 [119] Anti-PD-L1 mAb

Colorectal cancer,

Ovarian cancer,

Mesothelioma

Preclinical study 58.3% (7/12) complete tumor response (Significantly better than monotherapy)
Radium-223 [111] Atezolizumab (Anti–PD-L1 mAb) mCRPC Clinical trial No clear evidence of clinical benefit
Radium-223 [125]

Pembrolizumab

(Anti-PD-1 mAb)

mCRPC Clinical trial No efficacy improvement

Actinium-225

NCT04946370

Pembrolizumab (Anti-PD-1 mAb) mCRPC Clinical trial Incomplete

Actinium-225

NCT06939036

Atezolizumab, Durvalumab, Avelumab, Pembrolizumab, Retifanlimab (Anti-PD-1/PD-L1 mAb)

ES-SCLC,

MCC

Clinical trial Incomplete

Lead-212

NCT05655312

Nivolumab (Anti-PD-1 mAb) Melanoma Clinical trial Incomplete

PFS, progression-free survival, TTP time to progression, mCRPC metastatic castration-resistant prostate cancer

Challenge and future perspectives

Given the short range of alpha particles, TAT is generally considered to have a high safety profile, with a low incidence of severe acute or subacute adverse events (AEs). However, hematologic toxicity remains the most characteristic AE of TAT, primarily manifested as cytopenia due to bone marrow suppression [126]. Clinical data show that in prostate cancer patients with bone metastases treated with 223RaCl₂, the incidence of neutropenia and thrombocytopenia is significantly increased, indicating hematopoietic suppression [127]. Parlani et al. concluded from animal experiments that the myelotoxicity induced by 223Ra therapy is acute and reversible, without causing permanent damage to the hematopoietic system [128]. However, this conclusion requires validation through extensive clinical trials. This hematologic toxicity remains a key bottleneck limiting the clinical translation of TAT.

Although the incidence of hematologic toxicity with 225Ac-labeled therapies (e.g., 225Ac-PSMA-617) is relatively low [129], its longer half-life may increase the risk of cumulative toxicity. Some studies have observed nephrotoxicity in the clinical application of 225Ac [124], which may be related to the redistribution of daughter nuclides (e.g., 213Bi) due to recoil effects, leading to off-target radiation damage. In one clinical trial of 225Ac-PSMA-617 radioligand therapy for mCRPC, delayed nephrotoxicity was observed [130]. However, data on long-term adverse events remain limited, and future studies should expand clinical trial cohorts and extend follow-up periods to clarify its safety boundaries.

In addition to the aforementioned AEs, the multiple challenges in accurate dosimetry assessment of alpha-emitting radionuclides further hinder their clinical translation.

Alpha-emitters predominantly rely on γ rays emitted during decay for imaging. However, most therapeutic alpha-emitters are administered at low activities, with γ photon yields being particularly low (γ ray abundance < 1%) and the emitted photon energy spectrum is complex. These characteristics result in poor signal-to-noise ratios in conventional SPECT/CT imaging, thereby limiting the reliability of quantitative imaging. Early MIRD monographs made adjustments to dosimetry for alpha-emitters, but calculations for alpha-particle radiopharmaceutical therapy in human tissues were based on assumptions that prove inaccurate. These assumptions fail to account for the heterogeneous microscopic distribution of alpha-particle radiopharmaceutical therapy agents, making dose calculations challenging [131]. Certain alpha-emitters with multiple decay progeny (e.g., 223Ra, 225Ac, 227Th) feature daughter radionuclides with long half-lives. The high-energy nuclear recoil released during alpha decay may cause redistribution of these daughter nuclides within the body [8]. For instance, 213Bi from 225Ac decay tends to relocate to kidneys [132]. Accurate dosimetry of alpha-emitter radiopharmaceuticals requires simultaneous imaging of both the target parent nuclides and their redistributed daughters. Understanding the distribution of long-lived progeny is equally crucial for toxicity assessment, as they may induce off-target organ toxicity [133]. Dosimetric measurement of alpha-emitters remains an urgent challenge that needs to be addressed. Due to the short path length, high LET, and heterogeneous distribution of alpha-emitters within target volumes, small-scale dosimetry and microdosimetry assessments has become an essential approach for accurately evaluating the therapeutic potential of TAT and risks to normal tissues. Current research in targeted radionuclide therapy utilizes both Monte Carlo simulations and analytical methods for small-scale dosimetry and microdosimetry assessments. These computational approaches enable evaluation of radiopharmaceutical biodistribution, energy deposition, and radiobiological modeling. While achieving accurate dose measurements remains challenging, reliable dosimetric quantification is critical for ensuring both therapeutic efficacy and patient safety [134].

Conclusion

TAT offers a highly promising strategy for precision oncology by integrating the multidimensional biological effects of targeted effect, bystander effect, and abscopal effect. The targeted effect achieves efficient eradication of localized lesions through α-particles, while the bystander effect overcomes therapeutic blind spots caused by tumor heterogeneity via intercellular signaling transfer, significantly expanding therapeutic coverage. The abscopal effect further enhances systemic therapeutic potential by activating systemic anti-tumor immune responses, inducing regression of distant metastases. The synergy of these three effects establishes a three-dimensional anti-tumor network at the “local-regional-systemic” levels, pioneering novel pathways for precision cancer therapy. However, the clinical translation of TAT still faces critical bottlenecks: despite the precise tumor cell eradication enabled by alpha-emitters through their high LET and short penetration range, the delivery efficiency and stability of radiopharmaceutical carriers remain critical limitations for the broad application of targeted effects. While the bystander effects extend therapeutic impact beyond directly irradiated cells, it introduces potential risks: the nonlinear dose-response characteristics, distant tissue toxicity, and optimization of personalized dosing require in-depth investigation, necessitating high-precision in vivo models to define safety thresholds. Furthermore, although the abscopal effects induced by alpha-emitters have garnered significant interest, their combination with ICIs has yet to demonstrate substantial clinical benefits, highlighting the need for systematic exploration of synergistic strategies with immunotherapy. In conclusion, overcoming these key challenges is essential to translate the biological advantages of alpha-emitters into clinically accessible precision anticancer tools.

Author contributions

All authors contributed to the study conception and design. Zhiling Song, Jiajia Zhang and Shanshan Qin conceived the ideas and wrote the initial draft. Xiaohui Luan, Han Zhang reviewed and edited the manuscript. Mengdie Yang, Yao Jin and Gang Yang created the figures and tables. Fei Yu coordinated the project and supervised the work. All authors read and approved the final manuscript.

Funding

This work was conducted with the supports by National Natural Science Foundation of China (Grant No. 82272030 and No. 82472028) and China University Industry-University-Research Institute Innovation Fund (No.2023HT066).

Data availability

Data sharing does not apply to this article as no datasets were generated during the current study.

Declarations

Ethics

This is a systematic literature review; therefore, no ethical approval is required.

Competing interests

The authors have no relevant financial or non-financial interests to disclose.

Footnotes

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Zhiling Song, Jiajia Zhang and Shanshan Qin contributed equally to this work.

References

  • 1.Poty S, Francesconi LC, McDevitt MR, Morris MJ, Lewis JS. α-Emitters for radiotherapy: from basic radiochemistry to clinical Studies—Part 1. J Nucl Med. 2018;59:878–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.McDevitt MR, Sgouros G, Sofou S. Targeted and nontargeted α-Particle therapies. Annu Rev Biomed Eng. 2018;20:73–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Aghevlian S, Boyle AJ, Reilly RM. Radioimmunotherapy of cancer with high linear energy transfer (LET) radiation delivered by radionuclides emitting α-particles or Auger electrons. Adv Drug Deliv Rev. 2017;109:102–18. [DOI] [PubMed] [Google Scholar]
  • 4.Sgouros G, Roeske JC, McDevitt MR, Palm S, Allen BJ, Fisher DR, et al. MIRD pamphlet 22 (Abridged): radiobiology and dosimetry of α-Particle emitters for targeted radionuclide therapy. J Nucl Med Off Publ Soc Nucl Med. 2010;51:311–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Wulbrand C, Seidl C, Gaertner FC, Bruchertseifer F, Morgenstern A, Essler M, et al. Alpha-Particle emitting 213Bi-Anti-EGFR immunoconjugates eradicate tumor cells independent of oxygenation. PLoS ONE. 2013;8:e64730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Staudacher AH, Liapis V, Brown MP. Therapeutic targeting of tumor hypoxia and necrosis with antibody α-radioconjugates. Antib Ther. 2018;1:55–63.30406213 [Google Scholar]
  • 7.Karimian A, Ji NT, Song H, Sgouros G. Mathematical modeling of preclinical Alpha-Emitter radiopharmaceutical therapy. Cancer Res. 2020;80:868–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Kratochwil C, Bruchertseifer F, Giesel FL, Weis M, Verburg FA, Mottaghy F, et al. 225Ac-PSMA-617 for PSMA-Targeted α-Radiation therapy of metastatic Castration-Resistant prostate Cancer. J Nucl Med. 2016;57:1941–4. [DOI] [PubMed] [Google Scholar]
  • 9.Miederer M, Benešová-Schäfer M, Mamat C, Kästner D, Pretze M, Michler E, et al. Alpha-Emitting radionuclides: current status and future perspectives. Pharmaceuticals. 2024;17:76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Pouget J-P, Constanzo J. Revisiting the radiobiology of targeted alpha therapy. Front Med. 2021;8:692436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Nikitaki Z, Velalopoulou A, Zanni V, Tremi I, Havaki S, Kokkoris M, et al. Key biological mechanisms involved in high-LET radiation therapies with a focus on DNA damage and repair. Expert Rev Mol Med. 2022;24:e15. [DOI] [PubMed] [Google Scholar]
  • 12.Hirano S, Ichikawa Y, Sato B, Yamamoto H, Takefuji Y, Satoh F. Molecular hydrogen as a potential clinically applicable radioprotective agent. Int J Mol Sci. 2021;22:4566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.O’Neill P, Wardman P. Radiation chemistry comes before radiation biology. Int J Radiat Biol. 2009;85:9–25. [DOI] [PubMed] [Google Scholar]
  • 14.Tu W, Dong C, Fu J, Pan Y, Kobayashi A, Furusawa Y, et al. Both irradiated and bystander effects link with DNA repair capacity and the linear energy transfer. Life Sci. 2019;222:228–34. [DOI] [PubMed] [Google Scholar]
  • 15.Xu G, Chapman JR, Brandsma I, Yuan J, Mistrik M, Bouwman P, et al. REV7 counteracts DNA double-strand break resection and affects PARP Inhibition. Nature. 2015;521:541–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Chen Y, Wu J, Zhai L, Zhang T, Yin H, Gao H, et al. Metabolic regulation of homologous recombination repair by MRE11 lactylation. Cell. 2024;187:294–e31121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Muggiolu G, Torfeh E, Simon M, Devès G, Seznec H, Barberet P. Recruitment kinetics of XRCC1 and RNF8 following MeV proton and α-Particle Micro-Irradiation. Biology. 2023;12:921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Shibata A, Jeggo P, Löbrich M. The pendulum of the Ku-Ku clock. DNA Repair. 2018;71:164–71. [DOI] [PubMed] [Google Scholar]
  • 19.Maréchal A, Zou L. DNA damage sensing by the ATM and ATR kinases. Cold Spring Harb Perspect Biol. 2013;5:a012716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Baidoo KE, Yong K, Brechbiel MW. Molecular pathways: targeted α-Particle radiation therapy. Clin Cancer Res Off J Am Assoc Cancer Res. 2013;19:530–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Panier S, Boulton SJ. Double-strand break repair: 53BP1 comes into focus. Nat Rev Mol Cell Biol. 2014;15:7–18. [DOI] [PubMed] [Google Scholar]
  • 22.Tartier L, Gilchrist S, Burdak-Rothkamm S, Folkard M, Prise KM. Cytoplasmic irradiation induces Mitochondrial-Dependent 53BP1 protein relocalization in irradiated and bystander cells. Cancer Res. 2007;67:5872–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kastan MB, Bartek J. Cell-cycle checkpoints and cancer. Nature. 2004;432:316–23. [DOI] [PubMed] [Google Scholar]
  • 24.Harper JW, Elledge SJ. The DNA damage response: ten years after. Mol Cell. 2007;28:739–45. [DOI] [PubMed] [Google Scholar]
  • 25.Vousden KH, Lane DP. p53 in health and disease. Nat Rev Mol Cell Biol. 2007;8:275–83. [DOI] [PubMed] [Google Scholar]
  • 26.Marques IA, Abrantes AM, Pires AS, Neves AR, Caramelo FJ, Rodrigues T, et al. Kinetics of radium-223 and its effects on survival, proliferation and DNA damage in lymph-node and bone metastatic prostate cancer cell lines. Int J Radiat Biol. 2021;97:714–26. [DOI] [PubMed] [Google Scholar]
  • 27.Pouget JP, Mather SJ. General aspects of the cellular response to low- and high-LET radiation. Eur J Nucl Med. 2001;28:541–61. [DOI] [PubMed] [Google Scholar]
  • 28.Thompson LH. Recognition, signaling, and repair of DNA double-strand breaks produced by ionizing radiation in mammalian cells: the molecular choreography. Mutat Res Mutat Res. 2012;751:158–246. [DOI] [PubMed] [Google Scholar]
  • 29.Nonnekens J, Chatalic KLS, Molkenboer-Kuenen JDM, Beerens CEMT, Bruchertseifer F, Morgenstern A, et al. 213Bi-Labeled prostate-Specific membrane Antigen-Targeting agents induce DNA Double-Strand breaks in prostate Cancer xenografts. Cancer Biother Radiopharm. 2017;32:67–73. [DOI] [PubMed] [Google Scholar]
  • 30.Liu W, Ma H, Liang R, Chen X, Li H, Lan T, et al. Targeted alpha therapy of glioma using 211At-Labeled heterodimeric peptide targeting both VEGFR and integrins. Mol Pharm. 2022;19:3206–16. [DOI] [PubMed] [Google Scholar]
  • 31.Abramenkovs A, Hariri M, Spiegelberg D, Nilsson S, Stenerlöw B. Ra-223 induces clustered DNA damage and inhibits cell survival in several prostate cancer cell lines. Transl Oncol. 2022;26:101543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Pouget J-P, Tardieu M, Poty S. The radiobiology of radiopharmaceutical therapy: the input of imaging and radiomics. In: Bodei L, Lewis JS, Zeglis BM, editors. Radiopharm. Ther. Cham: Springer International Publishing; 2023. pp. 91–122. [Google Scholar]
  • 33.Santana P, Peña LA, Haimovitz-Friedman A, Martin S, Green D, McLoughlin M, et al. Acid sphingomyelinase-deficient human lymphoblasts and mice are defective in radiation-induced apoptosis. Cell. 1996;86:189–99. [DOI] [PubMed] [Google Scholar]
  • 34.Scheel-Toellner D, Wang K, Craddock R, Webb PR, McGettrick HM, Assi LK, et al. Reactive oxygen species limit neutrophil life span by activating death receptor signaling. Blood. 2004;104:2557–64. [DOI] [PubMed] [Google Scholar]
  • 35.Corre I, Guillonneau M, Paris F. Membrane signaling induced by high doses of ionizing radiation in the endothelial compartment. Relevance in radiation toxicity. Int J Mol Sci. 2013;14:22678–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Kolesnick R, Fuks Z. Radiation and ceramide-induced apoptosis. Oncogene. 2003;22:5897–906. [DOI] [PubMed] [Google Scholar]
  • 37.Cs Szabo B, Szabo M, Nagy P, Varga Z, Panyi G, Kovacs T, et al. Novel insights into the modulation of the voltage-gated potassium channel KV1.3 activation gating by membrane ceramides. J Lipid Res. 2024;65:100596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Paillas S, Ladjohounlou R, Lozza C, Pichard A, Boudousq V, Jarlier M, et al. Localized irradiation of cell membrane by Auger electrons is cytotoxic through oxidative Stress-Mediated nontargeted effects. Antioxid Redox Signal. 2016;25:467–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Seideman JH, Stancevic B, Rotolo JA, McDevitt MR, Howell RW, Kolesnick RN, et al. Alpha particles induce apoptosis through the sphingomyelin pathway. Radiat Res. 2011;176:434–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Ladjohounlou R, Lozza C, Pichard A, Constanzo J, Karam J, Le Fur P, et al. Drugs that modify cholesterol metabolism alter the p38/JNK-Mediated targeted and nontargeted response to alpha and Auger radioimmunotherapy. Clin Cancer Res. 2019;25:4775–90. [DOI] [PubMed] [Google Scholar]
  • 41.Pouget J-P, Georgakilas AG, Ravanat J-L. Targeted and Off-Target (Bystander and Abscopal) effects of radiation therapy: redox mechanisms and risk/benefit analysis. Antioxid Redox Signal. 2018;29:1447–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Azzam EI, Jay-Gerin J-P, Pain D. Ionizing radiation-induced metabolic oxidative stress and prolonged cell injury. Cancer Lett. 2012;327:48–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Kam WW-Y, Banati RB. Effects of ionizing radiation on mitochondria. Free Radic Biol Med. 2013;65:607–19. [DOI] [PubMed] [Google Scholar]
  • 44.Schilling-Tóth B, Sándor N, Kis E, Kadhim M, Sáfrány G, Hegyesi H. Analysis of the common deletions in the mitochondrial DNA is a sensitive biomarker detecting direct and non-targeted cellular effects of low dose ionizing radiation. Mutat Res Mol Mech Mutagen. 2011;716:33–9. [DOI] [PubMed] [Google Scholar]
  • 45.Zhang B, Davidson MM, Zhou H, Wang C, Walker WF, Hei TK. Cytoplasmic irradiation results in mitochondrial dysfunction and DRP1-dependent mitochondrial fission. Cancer Res. 2013;73:6700–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Cho D-H, Nakamura T, Fang J, Cieplak P, Godzik A, Gu Z, et al. S-Nitrosylation of Drp1 mediates β-Amyloid–Related mitochondrial fission and neuronal injury. Science. 2009;324:102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Manczak M, Calkins MJ, Reddy PH. Impaired mitochondrial dynamics and abnormal interaction of amyloid beta with mitochondrial protein Drp1 in neurons from patients with alzheimer’s disease: implications for neuronal damage. Hum Mol Genet. 2011;20:2495–509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Baljinnyam E, Venkatesh S, Gordan R, Mareedu S, Zhang J, Xie L, et al. Effect of densely ionizing radiation on cardiomyocyte differentiation from human-induced pluripotent stem cells. Physiol Rep. 2017;5:e13308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Yang M, Li J, Han Z, Luan X, Zhang X, Gao J, et al. Layered double hydroxides for Radium-223 targeted alpha therapy with elicitation of the immune response. Adv Healthc Mater. 2025;14:2403175. [DOI] [PubMed] [Google Scholar]
  • 50.Noguchi M, Hirata N, Tanaka T, Suizu F, Nakajima H, Chiorini JA. Autophagy as a modulator of cell death machinery. Cell Death Dis. 2020;11:517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Diniz Filho JFS, de Barros AOda, Pijeira S, Ricci-Junior MSO, Midlej E, Baroni V. Ultrastructural analysis of Cancer cells treated with the radiopharmaceutical radium dichloride ([223Ra]RaCl2): Understanding the effect on cell structure. Cells. 2023;12:451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Malamas AS, Gameiro SR, Knudson KM, Hodge JW. Sublethal exposure to alpha radiation (223Ra dichloride) enhances various carcinomas’ sensitivity to Lysis by antigen-specific cytotoxic T lymphocytes through calreticulin-mediated Immunogenic modulation. Oncotarget. 2016;7:86937–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Murao A, Aziz M, Wang H, Brenner M, Wang P. Release mechanisms of major damps. Apoptosis. 2021;26:152–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Kelly MP, Lee FT, Tahtis K, Smyth FE, Brechbiel MW, Scott AM. Radioimmunotherapy with α-Particle–Emitting 213Bi-C-Functionalized trans-Cyclohexyl-Diethylenetriaminepentaacetic Acid-Humanized 3S193 is enhanced by combination with Paclitaxel chemotherapy. Clin Cancer Res. 2007;13:s5604–12. [DOI] [PubMed] [Google Scholar]
  • 55.Yao Z, Garmestani K, Wong KJ, Park LS, Dadachova E, Yordanov A, et al. Comparative cellular catabolism and retention of Astatine-, Bismuth-, and Lead-Radiolabeled internalizing monoclonal antibody. J Nucl Med. 2001;42:1538–44. [PubMed] [Google Scholar]
  • 56.UNSCEAR 2006 Report Volume l. United Nations:Scientific Committee on the Effects of Atomic Radiation.
  • 57.Mothersill C, Seymour CB, Joiner MC. Relationship between radiation-induced low-dose hypersensitivity and the bystander effect. Radiat Res. 2002;157:526–32. [DOI] [PubMed] [Google Scholar]
  • 58.Nagasawa H, Little JB. Induction of sister chromatid exchanges by extremely low doses of alpha-particles. Cancer Res. 1992;52:6394–6. [PubMed] [Google Scholar]
  • 59.Chouin N, Bernardeau K, Bardiès M, Faivre-Chauvet A, Bourgeois M, Apostolidis C, et al. Evidence of extranuclear cell sensitivity to alpha-particle radiation using a microdosimetric model. II. Application of the microdosimetric model to experimental results. Radiat Res. 2009;171:664–73. [DOI] [PubMed] [Google Scholar]
  • 60.Leung CN, Canter BS, Rajon D, Bäck TA, Fritton JC, Azzam EI, et al. Dose-Dependent growth delay of breast Cancer xenografts in the bone marrow of mice treated with 223Ra: the role of bystander effects and their potential for therapy. J Nucl Med. 2020;61:89–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Canter BS, Leung CN, Fritton JC, Bäck T, Rajon D, Azzam EI, et al. Radium-223-induced bystander effects cause DNA damage and apoptosis in disseminated tumor cells in bone marrow. Mol Cancer Res MCR. 2021;19:1739–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Rajon DA, Canter BS, Leung CN, Bäck TA, Fritton JC, Azzam EI, et al. Modeling bystander effects that cause growth delay of breast cancer xenografts in bone marrow of mice treated with radium-223. Int J Radiat Biol. 2021;97:1217–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Yang M, Wang L, Qin S, Dai X, Li J, An L, et al. Role of damaged mitochondrial transfer in alpha-particle generator 212Pb radiation-induced bystander effect. Theranostics. 2024;14:6768–82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Klammer H, Mladenov E, Li F, Iliakis G. Bystander effects as manifestation of intercellular communication of DNA damage and of the cellular oxidative status. Cancer Lett. 2015;356:58–71. [DOI] [PubMed] [Google Scholar]
  • 65.Azzam EI, de Toledo SM, Gooding T, Little JB. Intercellular communication is involved in the bystander regulation of gene expression in human cells exposed to very low fluences of alpha particles. Radiat Res. 1998;150:497–504. [PubMed] [Google Scholar]
  • 66.Azzam EI, de Toledo SM, Little JB. Direct evidence for the participation of gap junction-mediated intercellular communication in the transmission of damage signals from α-particle irradiated to nonirradiated cells. Proc Natl Acad Sci U S A. 2001;98:473–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Autsavapromporn N, de Toledo SM, Little JB, Jay-Gerin J-P, Harris AL, Azzam EI. The role of gap junction communication and oxidative stress in the propagation of toxic effects among high-dose α-particle-irradiated human cells. Radiat Res. 2011;175:347–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Shao C, Folkard M, Prise KM. Role of TGF-beta1 and nitric oxide in the bystander response of irradiated glioma cells. Oncogene. 2008;27:434–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Li J, He M, Shen B, Yuan D, Shao C. Alpha particle-induced bystander effect is mediated by ROS via a p53-dependent SCO2 pathway in hepatoma cells. Int J Radiat Biol. 2013;89:1028–34. [DOI] [PubMed] [Google Scholar]
  • 70.Whiteside SK, Snook JP, Williams MA, Weis JJ, Bystander T, Cells. A balancing act of friends and foes. Trends Immunol. 2018;39:1021–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Burdak-Rothkamm S, Short SC, Folkard M, Rothkamm K, Prise KM. ATR-dependent radiation-induced γH2AX foci in bystander primary human astrocytes and glioma cells. Oncogene. 2007;26:993–1002. [DOI] [PubMed] [Google Scholar]
  • 72.Kim T-S, Shin E-C. The activation of bystander CD8 + T cells and their roles in viral infection. Exp Mol Med. 2019;51:1–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Restier-Verlet J, Joubert A, Ferlazzo ML, Granzotto A, Sonzogni L, Al-Choboq J, et al. X-rays-Induced bystander effect consists in the formation of DNA breaks in a Calcium-Dependent manner: influence of the experimental procedure and the individual factor. Biomolecules. 2023;13:542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Garcia-Barros M, Paris F, Cordon-Cardo C, Lyden D, Rafii S, Haimovitz-Friedman A, et al. Tumor response to radiotherapy regulated by endothelial cell apoptosis. Science. 2003;300:1155–9. [DOI] [PubMed] [Google Scholar]
  • 75.Leach JK, Van Tuyle G, Lin P-S, Schmidt-Ullrich R, Mikkelsen RB. Ionizing Radiation-induced, Mitochondria-dependent generation of reactive Oxygen/Nitrogen1. Cancer Res. 2001;61:3894–901. [PubMed] [Google Scholar]
  • 76.Mikkelsen RB, Wardman P. Biological chemistry of reactive oxygen and nitrogen and radiation-induced signal transduction mechanisms. Oncogene. 2003;22:5734–54. [DOI] [PubMed] [Google Scholar]
  • 77.Azzam EI, De Toledo SM, Spitz DR, Little JB. Oxidative metabolism modulates signal transduction and micronucleus formation in bystander cells from alpha-particle-irradiated normal human fibroblast cultures. Cancer Res. 2002;62:5436–42. [PubMed] [Google Scholar]
  • 78.Zhou H, Ivanov VN, Gillespie J, Geard CR, Amundson SA, Brenner DJ, et al. Mechanism of radiation-induced bystander effect: role of the cyclooxygenase-2 signaling pathway. Proc Natl Acad Sci U S A. 2005;102:14641–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Stuehr DJ, Santolini J, Wang Z-Q, Wei C-C, Adak S. Update on mechanism and catalytic regulation in the NO synthases**. J Biol Chem. 2004;279:36167–70. [DOI] [PubMed] [Google Scholar]
  • 80.Shao C, Furusawa Y, Aoki M, Matsumoto H, Ando K. Nitric oxide-mediated bystander effect induced by heavy-ions in human salivary gland tumour cells. Int J Radiat Biol. 2002;78:837–44. [DOI] [PubMed] [Google Scholar]
  • 81.Ferranti CS, Cheng J, Thompson C, Zhang J, Rotolo JA, Buddaseth S, et al. Fusion of lysosomes to plasma membrane initiates radiation-induced apoptosis. J Cell Biol. 2020;219:e201903176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.He M, Dong C, Xie Y, Li J, Yuan D, Bai Y, et al. Reciprocal bystander effect between α-irradiated macrophage and hepatocyte is mediated by cAMP through a membrane signaling pathway. Mutat Res Mol Mech Mutagen. 2014;763–764:1–9. [DOI] [PubMed] [Google Scholar]
  • 83.Shao C, Folkard M, Michael BD, Prise KM. Targeted cytoplasmic irradiation induces bystander responses. Proc Natl Acad Sci U S A. 2004;101:13495–500. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Wan C, Sun Y, Tian Y, Lu L, Dai X, Meng J, et al. Irradiated tumor cell–derived microparticles mediate tumor eradication via cell killing and immune reprogramming. Sci Adv. 2020;6:eaay9789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Smolarz M, Skoczylas Ł, Gawin M, Krzyżowska M, Pietrowska M, Widłak P. Radiation-Induced bystander effect mediated by exosomes involves the replication stress in recipient cells. Int J Mol Sci. 2022;23:4169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Peinado H, Alečković M, Lavotshkin S, Matei I, Costa-Silva B, Moreno-Bueno G, et al. Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through MET. Nat Med. 2012;18:883–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Takahashi A, Okada R, Nagao K, Kawamata Y, Hanyu A, Yoshimoto S, et al. Exosomes maintain cellular homeostasis by excreting harmful DNA from cells. Nat Commun. 2017;8:15287. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Arscott WT, Tandle AT, Zhao S, Shabason JE, Gordon IK, Schlaff CD, et al. Ionizing radiation and glioblastoma exosomes: implications in tumor biology and cell migration. Transl Oncol. 2013;6:638–IN6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Mutschelknaus L, Peters C, Winkler K, Yentrapalli R, Heider T, Atkinson MJ, et al. Exosomes derived from squamous head and neck Cancer promote cell survival after ionizing radiation. PLoS ONE. 2016;11:e0152213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Malla B, Aebersold DM, Dal Pra A. Protocol for serum Exosomal MiRNAs analysis in prostate cancer patients treated with radiotherapy. J Transl Med. 2018;16:223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Mallegol J, Niel GV, Lebreton C, Lepelletier Y, Candalh C, Dugave C, et al. T84-Intestinal epithelial exosomes bear MHC class ii/peptide complexes potentiating antigen presentation by dendritic cells. Gastroenterology. 2007;132:1866–76. [DOI] [PubMed] [Google Scholar]
  • 92.Boyd M, Ross SC, Dorrens J, Fullerton NE, Tan KW, Zalutsky MR, et al. Radiation-Induced biologic bystander effect elicited in vitro by targeted radiopharmaceuticals labeled with α-, β-, and Auger Electron–Emitting radionuclides. J Nucl Med. 2006;47:1007–15. [PubMed] [Google Scholar]
  • 93.Belyakov OV, Mitchell SA, Parikh D, Randers-Pehrson G, Marino SA, Amundson SA, et al. Biological effects in unirradiated human tissue induced by radiation damage up to 1 mm away. Proc Natl Acad Sci U S A. 2005;102:14203–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Hu B, Wu L, Han W, Zhang L, Chen S, Xu A, et al. The time and Spatial effects of bystander response in mammalian cells induced by low dose radiation. Carcinogenesis. 2006;27:245–51. [DOI] [PubMed] [Google Scholar]
  • 95.Mole RH. Whole body Irradiation—Radiobiology or medicine?? Br J Radiol. 1953;26:234–41. [DOI] [PubMed] [Google Scholar]
  • 96.Demaria S, Ng B, Devitt ML, Babb JS, Kawashima N, Liebes L, et al. Ionizing radiation Inhibition of distant untreated tumors (abscopal effect) is immune mediated. Int J Radiat Oncol Biol Phys. 2004;58:862–70. [DOI] [PubMed] [Google Scholar]
  • 97.Wu Q, Allouch A, Martins I, Brenner C, Modjtahedi N, Deutsch E, et al. Modulating both tumor cell death and innate immunity is essential for improving radiation therapy effectiveness. Front Immunol. 2017;8:613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Eapen RS, Williams SG, Macdonald S, Keam SP, Lawrentschuk N, Au L, et al. Neoadjuvant lutetium PSMA, the TIME and immune response in high-risk localized prostate cancer. Nat Rev Urol. 2024;21:676–86. [DOI] [PubMed] [Google Scholar]
  • 99.Ferreira CA, Potluri HK, Massey C, Grudzinski JJ, Carston A, Clemons N, et al. Profound Immunomodulatory effects of 225Ac-NM600 drive enhanced anti-tumor response in prostate cancer. bioRxiv; 2022. 2022.09.26.509374.
  • 100.Leung CN, Howell DM, Howell RW. Radium-223 dichloride causes transient changes in natural killer cell population and cytotoxic function. Int J Radiat Biol. 2021;97:1417–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Lunj S, Smith TAD, Reeves KJ, Currell F, Honeychurch J, Hoskin P, et al. Immune effects of α and β radionuclides in metastatic prostate cancer. Nat Rev Urol. 2024;21:651–61. [DOI] [PubMed] [Google Scholar]
  • 102.Ashrafizadeh M, Farhood B, Eleojo Musa A, Taeb S, Rezaeyan A, Najafi M. Abscopal effect in radioimmunotherapy. Int Immunopharmacol. 2020;85:106663. [DOI] [PubMed] [Google Scholar]
  • 103.van Pul KM, Fransen MF, van de Ven R, de Gruijl TD. Immunotherapy Goes local: the central role of lymph nodes in driving tumor infiltration and efficacy. Front Immunol. 2021;12:643291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Hagemann UB, Ellingsen C, Schuhmacher J, Kristian A, Mobergslien A, Cruciani V, et al. Mesothelin-Targeted Thorium-227 conjugate (MSLN-TTC): preclinical evaluation of a new targeted alpha therapy for Mesothelin-Positive cancers. Clin Cancer Res. 2019;25:4723–34. [DOI] [PubMed] [Google Scholar]
  • 105.Gorin J-B, Gouard S, Ménager J, Morgenstern A, Bruchertseifer F, Faivre-Chauvet A, et al. Alpha particles induce autophagy in multiple myeloma cells. Front Med. 2015;2:74. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Gorin J-B, Ménager J, Gouard S, Maurel C, Guilloux Y, Faivre-Chauvet A, et al. Antitumor immunity induced after α irradiation. Neoplasia. 2014;16:319–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Qin S, Yang Y, Zhang J, Yin Y, Liu W, Zhang H, et al. Effective treatment of SSTR2-Positive small cell lung Cancer using 211At-Containing targeted α-Particle therapy agent which promotes endogenous antitumor immune response. Mol Pharm. 2023;20:5543–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Zhang J, Li F, Yin Y, Liu N, Zhu M, Zhang H, et al. Alpha radionuclide-chelated radioimmunotherapy promoters enable local radiotherapy/chemodynamic therapy to discourage cancer progression. Biomater Res. 2022;26:44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Bellia S, Feliciani G, Duca M, Monti M, Turri V, Sarnelli A, et al. Clinical evidence of abscopal effect in cutaneous squamous cell carcinoma treated with diffusing alpha emitters radiation therapy: a case report. J Contemp Brachytherapy. 2019;11:449–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Kim JW, Shin MS, Kang Y, Kang I, Petrylak DP. Immune analysis of Radium-223 in patients with metastatic prostate Cancer. Clin Genitourin Cancer. 2018;16:e469–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Fong L, Morris MJ, Sartor O, Higano CS, Pagliaro L, Alva A, et al. A phase Ib study of Atezolizumab with Radium-223 dichloride in men with metastatic Castration-Resistant prostate Cancer. Clin Cancer Res Off J Am Assoc Cancer Res. 2021;27:4746–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Body J-J, Casimiro S, Costa L. Targeting bone metastases in prostate cancer: improving clinical outcome. Nat Rev Urol. 2015;12:340–56. [DOI] [PubMed] [Google Scholar]
  • 113.Gartrell BA, Coleman R, Efstathiou E, Fizazi K, Logothetis CJ, Smith MR, et al. Metastatic prostate Cancer and the bone: significance and therapeutic options. Eur Urol. 2015;68:850–8. [DOI] [PubMed] [Google Scholar]
  • 114.Saad F, Carles J, Gillessen S, Heidenreich A, Heinrich D, Gratt J, et al. Radium-223 and concomitant therapies in patients with metastatic castration-resistant prostate cancer: an international, early access, open-label, single-arm phase 3b trial. Lancet Oncol. 2016;17:1306–16. [DOI] [PubMed] [Google Scholar]
  • 115.Morris MJ, Corey E, Guise TA, Gulley JL, Kevin Kelly W, Quinn DI, et al. Radium-223 mechanism of action: implications for use in treatment combinations. Nat Rev Urol. 2019;16:745–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Qin Y, Imobersteg S, Frank S, Blanc A, Chiorazzo T, Berger P, et al. Signaling network response to α-Particle-Targeted therapy with the 225Ac-Labeled Minigastrin analog 225Ac-PP-F11N reveals the radiosensitizing potential of histone deacetylase inhibitors. J Nucl Med Off Publ Soc Nucl Med. 2023;64:873–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Li M, Liu D, Lee D, Cheng Y, Baumhover NJ, Marks BM, et al. Targeted Alpha-Particle radiotherapy and immune checkpoint inhibitors induces cooperative Inhibition on tumor growth of malignant melanoma. Cancers. 2021;13:3676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Vardaki I, Corn P, Gentile E, Song JH, Madan N, Hoang A, et al. Radium 223 treatment increase immune checkpoint expression in extracellular vesicles from the metastatic prostate cancer bone microenvironment. Clin Cancer Res Off J Am Assoc Cancer Res. 2021;27:3253–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Lejeune P, Cruciani V, Berg-Larsen A, Schlicker A, Mobergslien A, Bartnitzky L, et al. Immunostimulatory effects of targeted thorium-227 conjugates as single agent and in combination with anti-PD-L1 therapy. J Immunother Cancer. 2021;9:e002387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Bellavia MC, Patel RB, Anderson CJ. Combined targeted radiopharmaceutical therapy and immune checkpoint blockade: from preclinical advances to the clinic. J Nucl Med. 2022;63:1636–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Jiao R, Allen KJH, Malo ME, Rickles D, Dadachova E. Evaluating the combination of radioimmunotherapy and immunotherapy in a melanoma mouse model. Int J Mol Sci. 2020;21:773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Dabagian H, Taghvaee T, Martorano P, Martinez D, Samanta M, Watkins CM, et al. PARP targeted Alpha-Particle therapy enhances response to PD-1 Immune-Checkpoint Blockade in a syngeneic mouse model of glioblastoma. ACS Pharmacol Transl Sci. 2021;4:344–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Czernin J, Current K, Mona CE, Nyiranshuti L, Hikmat F, Radu CG, et al. Immune-Checkpoint Blockade enhances 225Ac-PSMA617 efficacy in a mouse model of prostate Cancer. J Nucl Med. 2021;62:228–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Ertveldt T, Krasniqi A, Ceuppens H, Puttemans J, Dekempeneer Y, Jonghe KD, et al. Targeted α-Therapy using 225Ac radiolabeled Single-Domain antibodies induces Antigen-Specific immune responses and instills Immunomodulation both systemically and at the tumor microenvironment. J Nucl Med. 2023;64:751–8. [DOI] [PubMed] [Google Scholar]
  • 125.Choudhury AD, Kwak L, Cheung A, Allaire KM, Marquez J, Yang DD, et al. Randomized phase II study evaluating the addition of pembrolizumab to Radium-223 in metastatic Castration-resistant prostate Cancer. Cancer Immunol Res. 2024;12:704–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Activity. Adverse events of Actinium-225-PSMA-617 in advanced metastatic Castration-resistant prostate Cancer after failure of Lutetium-177-PSMA. Eur Urol. 2021;79:343–50. [DOI] [PubMed] [Google Scholar]
  • 127.Hematologic Safety of Radium-223 Dichloride. Baseline prognostic factors associated with myelosuppression in the ALSYMPCA trial. Clin Genitourin Cancer. 2017;15:42–e528. [DOI] [PubMed] [Google Scholar]
  • 128.Parlani M, Boccalatte F, Yeaton A, Wang F, Zhang J, Aifantis I, et al. 223Ra induces transient functional bone marrow toxicity. J Nucl Med. 2022;63:1544–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Lawal IO, Morgenstern A, Vorster M, Knoesen O, Mahapane J, Hlongwa KN, et al. Hematologic toxicity profile and efficacy of [225Ac]Ac-PSMA-617 α-radioligand therapy of patients with extensive skeletal metastases of castration-resistant prostate cancer. Eur J Nucl Med Mol Imaging. 2022;49:3581–92. [DOI] [PubMed] [Google Scholar]
  • 130.Satapathy S, Sharma A, Sood A, Maheshwari P, Gill HJS. Delayed nephrotoxicity after 225Ac-PSMA-617 radioligand therapy. Clin Nucl Med. 2022;47:e466. [DOI] [PubMed] [Google Scholar]
  • 131.Sgouros G, Dosimetry. Radiobiology and synthetic lethality: radiopharmaceutical therapy (RPT) with Alpha-Particle-Emitters. Semin Nucl Med. 2020;50:124–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Schwartz J, Jaggi JS, O’Donoghue JA, Ruan S, McDevitt M, Larson SM, et al. Renal uptake of bismuth-213 and its contribution to kidney radiation dose following administration of actinium-225-labeled antibody. Phys Med Biol. 2011;56:721–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Sgouros G, Frey E, Du Y, Hobbs R, Bolch W. Imaging and dosimetry for alpha-particle emitter radiopharmaceutical therapy: improving radiopharmaceutical therapy by looking into the black box. Eur J Nucl Med Mol Imaging. 2021;49:18–29. [DOI] [PubMed] [Google Scholar]
  • 134.Tronchin S, Forster JC, Hickson K, Bezak E. Dosimetry in targeted alpha therapy. A systematic review: current findings and what is needed. Phys Med Biol. 2022;67:09TR01. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing does not apply to this article as no datasets were generated during the current study.


Articles from European Journal of Nuclear Medicine and Molecular Imaging are provided here courtesy of Springer

RESOURCES