Next Article in Journal
Cervical Cancer Treatment and Fertility: What We Know and What We Do
Previous Article in Journal
A Preliminary Study on Deep Learning-Based Plan Quality Prediction in Gamma Knife Radiosurgery for Brain Metastases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Actinium-225/Bismuth-213 as Potential Leaders for Targeted Alpha Therapy: Current Supply, Application Barriers, and Future Prospects

by
Mohamed F. Nawar
1,2,
Adli A. Selim
3,
Basma M. Essa
2,
Alaa F. El-Daoushy
1,2,
Mohamed M. Swidan
3,
Claudia G. Chambers
4,5,6,
Mohammed H. Al Qahtani
7,
Charles J. Smith
6,8,9,* and
Tamer M. Sakr
2,9,*
1
Department of Chemistry, Biochemistry and Pharmaceutical Sciences (DCBP), Faculty of Science, University of Bern, CH-3012 Bern, Switzerland
2
Radioactive Isotopes and Generator Department, Hot Labs Center, Egyptian Atomic Energy Authority, Cairo P.O. Box 13759, Egypt
3
Labeled Compounds Department, Hot Labs Center, Egyptian Atomic Energy Authority, Cairo P.O. Box 13759, Egypt
4
Department of Chemistry, University of Missouri, Columbia, MO 65201, USA
5
Molecular Imaging and Theranostics Center, University of Missouri, Columbia, MO 65201, USA
6
Research Division, Harry S. Truman Memorial Veterans’ Hospital, Columbia, MO 65201, USA
7
Cyclotron and Radiopharmaceuticals Department, King Faisal Specialist Hospital and Research Center, Riyadh 11211, Saudi Arabia
8
University of Missouri Research Reactor Center, University of Missouri, Columbia, MO 65201, USA
9
Department of Radiology, School of Medicine, University of Missouri, Columbia, MO 65201, USA
*
Authors to whom correspondence should be addressed.
Cancers 2025, 17(18), 3055; https://doi.org/10.3390/cancers17183055
Submission received: 18 August 2025 / Revised: 14 September 2025 / Accepted: 16 September 2025 / Published: 18 September 2025
(This article belongs to the Section Cancer Therapy)

Simple Summary

This review assesses the potential of Actinium-225 and Bismuth-213 as leading radionuclides for targeted alpha therapy, highlighting their ability to deliver potent cytotoxic radiation with high tumor specificity. It addresses current challenges related to their production and radiolabeling, and provides a comprehensive overview of recent preclinical and clinical evaluations demonstrating their promise as effective radiotheranostic agents.

Abstract

Alpha therapy (TAT) relies on combining alpha-emitting radionuclides with specific cell-targeting vectors to deliver a high payload of cytotoxic radiation capable of destroying tumor tissues. TAT efficacy comes from the tissue selectivity of the targeting vector, the high linear energy transfer (LET) of the radionuclide, and the short range of alpha particles in tissues. Recent research studies have been directed to evaluate TAT on a preclinical and clinical scale, including evaluating damage to tumor tissues with minimal toxic radiation effects on surrounding healthy tissues. This review highlights the use of Actinium-225/Bismuth-213 radionuclides as promising candidates for TAT. Herein, we begin with a discussion on the production and supply of [225Ac]Ac/[213Bi]Bi followed by the formulation of [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals using different radiolabeling techniques. Finally, we have summarized the preclinical and clinical evaluation of these potential radiotheranostic agents.

1. Introduction

Targeted Alpha Therapy (TAT) offers an innovative approach to cancer treatment. The novel technique of pairing alpha-emitting radionuclides with specific targeting agents, including antibodies or peptides, offers an effective means of treating metastatic tumors. This unique strategy provides a direct cytotoxic radiation dose to targeted cancer cells [1,2,3,4,5,6,7,8].
The use of alpha-emitting radionuclides for cancer treatment has been explored over different conventional methods. This may be attributed to two specific reasons. First, the short penetration length of alpha particles inside the human body (˂0.1 mm) dramatically helps to affect only the targeted cancer cells, protecting the neighboring healthy tissues. Secondly, the high linear energy transfer (LET) of high-energy alpha radiation results in highly effective cell destruction [9,10]. Consequently, alpha-emitters can be more effective in killing metastatic cells that do not respond to other treatment strategies, such as beta- and gamma-emitting irradiations and/or chemotherapeutic medications. As a result, TAT creates a better therapeutic opportunity for cases resistant to conventional therapies [11,12,13,14,15,16]. In this regard, the development of radiopharmaceuticals incorporating alpha-emitting radionuclides for TAT has become a competitive interest for both academic research and commercial industries [17,18,19,20].
Recently, two alpha emitters have been introduced as promising radionuclides for cancer treatment. These radionuclides are [225Ac]Ac and its daughter [213Bi]Bi, which are a result of the 233U decay chain. Despite their potential, several challenges face their optimum usage on a large scale. For example, the worldwide supply of [225Ac]Ac is limited to approximately 63 GBq (1.7 Ci) annually. This limited amount of radionuclide can only sustain the treatment of 100–200 patients per year on the global stage. Moreover, tedious post-irradiation separation and purification steps are required to isolate [225Ac]Ac from any undesired co-produced isotopes during the production process. Furthermore, the notable lack of a clear understanding of the chemistry of [225Ac]Ac has dramatically decreased the chances of producing highly stable, radiolabeled cell targeting agents in an acceptable yield with sufficient pharmacodynamic properties. The daughter of [225Ac]Ac, [213Bi]Bi, can be produced on-site from the [225Ac]Ac/[213Bi]Bi generator. However, this daughter radionuclide shares some of the same production challenges as its parent [21].
The primary goal of this article is to provide an in-depth assessment of the production techniques of [225Ac]Ac and [213Bi]Bi that have been made commercially available for TAT. In addition, we discuss new supply routes that could have the potential to elevate the roles of [225Ac]Ac and [213Bi]Bi in various medical applications. Moreover, this work highlights the chemistry, the current radiolabeling strategies, and recent biological studies of [225Ac]Ac and [213Bi]Bi radiopharmaceuticals.

2. Chemistry and Radiochemistry of [225Ac]Ac and [213Bi]Bi

Actinium is the first element in the actinide series. Its electron configuration is [Rn] 6d1 7s2, and it usually exists as the Ac3+ cation with the configuration of [Rn] 6d0 7s0. There are no electrons in the outer shell, making actinium invisible to most forms of spectroscopy, consequently making its characterization difficult [22]. Actinium exists in an aqueous solution in the most stable state (+3), and is hydrolyzed at pH ranges from 8.6 to 10.4 [23]. Ac3+ polarizes the coordinated water molecule, affecting the proton release to form [Ac(OH)3-x]x+. Hence, in radiolabeling investigations, higher radiochemical yields (RCY) are expected when using alkaline buffers due to a decrease in [Ac(OH)3-x]x+ formation [24]. The chemical behavior of Ac3+ is very similar to lanthanum (La3+), and La3+ is often used as a suitable inactive surrogate for Ac3+. Because of a large ionic radius of 1.12 Å and high coordination number, [225Ac]Ac is often complexed with large polydentate chelating ligands. However, because of this larger radius, Ac3+ forms a weaker electrostatic bond with the donor atoms of the complexing ligand, which may result in instability [25]
The impetus driving [225Ac]Ac to be used as a radiotherapeutic isotope results from the inherent physical characteristics that it possesses. [225Ac]Ac is an alpha-emitting radionuclide known for its relatively long half-life (t1/2 = 9.92d) [26]. It decays via the emission of six subsequent daughter radionuclides to reach the most stable long-lived [209Bi]Bi (t1/2 = 2.01 × 1019 y). These daughters result from four alpha (α) decays with energies ranging between 5.8 MeV and 8.4 MeV and two beta (β) decays with energies of 0.093 and 0.660 MeV, which present 225Ac as a radiotherapeutic radioisotope as shown in Figure 1 [10].
The co-emission of gamma (γ) rays is also generated within the [225Ac]Ac decay chain, which is the result of the decay of [221Fr]Fr (218 keV, 11.6% emission probability) and [213Bi]Bi (440 keV, 26.1% emission probability). These gamma rays are useful for in vivo imaging applications.
The electron configuration of [213Bi]Bi is [Xe] 4f14 5d10 6s2 6p3. It forms complexes with coordinating ligands via the three electrons in 6p. Hence, the oxidation state of [213Bi]Bi is predominantly +3. In some cases, however, the +5 oxidation state may be seen, as in the case of the two electrons of 6s also participating in bonding. The coordination number of [213Bi]Bi ranges from three to ten. Bi+3 is a hard Lewis acid, so it has a high affinity for nitrogen and oxygen donors. Yet, complexes with halogens and sulfur are also well known.
[213Bi]Bi is a short-lived radionuclide (t1/2 = 45.6 min) and emits both alpha and beta particles. It decays mainly by the emission of beta particles to produce the ultra-short-lived and pure alpha emitter, [213Po]Po (t1/2 = 3.71 µs and Eα = 8.4 MeV), with a branching ratio of 97.8%. The residual 2.2% of [213Bi]Bi disintegrates with the emission of alpha particles to produce [209Tl]Tl with Eα = 5.549 MeV (0.16%) and Eα = 5.869 MeV (2.01%). Both [213Po]Po and [209Tl]Tl decay to yield [209Pb]Pb (t1/2 = 3.24 h), which in turn decays by beta particles to produce long-lived [209Bi]Bi (t1/2 = 1.9 × 1019 y). [213Po]Po decays by emitting an 8.4 MeV alpha particle with a penetrating power of 85 µm in human tissue. That makes it responsible for greater than 98% of the emitted alpha particle energy by the decay of [213Bi]Bi and is believed to be responsible for its cytotoxic effects. Approximately 92.7% of the overall particle energy that is generated from [213Bi]Bi decay results from alpha decay. Simultaneously, the remaining 7.3% of the decay energy results from beta decay [10]. [213Bi]Bi also decays by emitting gamma rays with a photon energy of 440 keV, which facilitates the detection of [213Bi]Bi biodistribution using a SPECT gamma scintigraph and also for performing pharmacokinetic and dosimetric studies [27].

3. Production Techniques of [225Ac]Ac

3.1. Current [225Ac]Ac Sources from Thorium Generators

Currently, [229Th]Th generators are considered the primary source of [225Ac]Ac production, and they can supply pure [225Ac]Ac free from any other actinium isotopes. These generators are being milked to produce hundreds of mCi of [225Ac]Ac on a monthly basis. The idea is built on the generation of [229Th]Th from the decay of 233U hoards, followed by the decay of [229Th]Th to [225Ra]Ra. Finally, [225Ra]Ra decays to produce [225Ac]Ac. Most [233U]U was produced in the period between 1954 and 1970 through the neutron activation of [232Th]Th, which was produced to test its utilization in nuclear reactors and the manufacturing of nuclear weapons. Since 1995, [229Th]Th has been extracted from [233U]U hoards that were stocked at four main places:
  • The US Department of Energy, Oak Ridge National Laboratory (ORNL) in Oak Ridge, TN, USA: 704 mg of [229Th]Th with a radioactivity level of 5.55 GBq (150 mCi) [28].
  • The Directorate for Nuclear Safety and Security of the JRC of the European Commission in Karlsruhe, Germany (formerly known as Institute for Trans-Uranium Elements (ITU) in Karlsruhe, Germany: this source consists of 215 mg of [229Th]Th and produces 1.7 GBq (46 mCi) [11,29].
  • Lipinski Institute for Physics and Power Engineering (IPPE) in Obninsk, Russia: this site contains 704 mg of [229Th]Th, which has a radioactivity level of 5.55 GBq (150 mCi).
  • At the Belgian Nuclear Research Centre (SCK CEN) in Mol, Belgium, very pure sources of [229Th]Th were identified, processed, and used for pre-clinical studies [30].
These four locations are considered to be the primary sources of the worldwide [225Ac]Ac supply and produce approximately 63 GBq (1.7 Ci) of [225Ac]Ac annually [31,32,33]. The market share of [225Ac]Ac between ORNL (USA), ITU (Germany), and IPPE (Russia) is 26.6 GBq (720 mCi), 13.1 GBq (350 mCi) [34], and 26.6 GBq (720 mCi), respectively. In addition, IPPE in Russia does not supply [225Ac]Ac regularly [35]. Accordingly, this small amount of [225Ac]Ac is sufficient for only 100–200 patient treatments per year. Moreover, on one hand, this low [225Ac]Ac production does not satisfy the researcher’s needs or those of increasing clinical trials, which are estimated to require approximately 185 GBq (5 Ci) every year [12,20]. On the other hand, for the development of new [213Bi]Bi-based radiopharmaceutical applications, the demand for [225Ac]Ac will be even greater.

3.2. Developed [225Ac]Ac Production Methods

Various methods for large-scale production of [225Ac]Ac have been explored, including spallation of natural uranium targets with energetic protons, as well as irradiation of [226Ra]Ra targets. The most advanced method is the spallation of [232Th]Th, successfully demonstrated at the Institute for Nuclear Research in Russia and Los Alamos National Laboratory in the United States. Routine production of [225Ac]Ac via this method has been established under the US Department of Energy’s Tri-Lab initiative, which encompasses Oak Ridge National Laboratory, Brookhaven National Laboratory, and the Los Alamos National Laboratory. Irradiations are conducted at Brookhaven (200 MeV at 165 mA) and Los Alamos (100 MeV at 275 mA), with processing and distribution of the final product occurring at ORNL. A key limitation of this process is the coproduction of long-lived [227Ac]Ac, with a half-life of 21.8 years, at activity levels of 0.1% to 0.2% at the end of bombardment (EOB) [36,37,38]. Thus, the vast growth of [225Ac]Ac utilization has promoted the exploration of new production techniques. These methods were proposed and tested experimentally, including the production of [225Ac]Ac either from [226Ra]Ra, or from [232Th]Th targets in nuclear reactors and electron or proton accelerators, according to but not limited to the following nuclear reactions:
  • 226Ra(γ,n)225Ra → 225Ac (electron accelerator; electrons of ~25–30 MeV) [35,37,38,39].
  • 226Ra(p,2n) 225Ac (proton accelerator; low-energy protons ~16 MeV) [37,40,41].
  • 226Ra(n,2n)225Ra → 225Ac (nuclear reactor; high-energy neutrons >6.4 MeV) [37,40,41].
  • 232Th(p,x)225Ac (Thorium spallation; high-energy protons > 70 MeV) [37,42].
  • 238U(p,x)225Ac (Uranium spallation; high-energy protons >100 MeV) [37,38].

3.2.1. [225Ac]Ac Production from [226Ra]Ra Targets

[225Ac]Ac Production via Electron Accelerators
Generally, a linear electron accelerator (linac) employs incident electron beams on tungsten targets to generate bremsstrahlung x-rays, which can be used for an external radiation beam therapy. Then, [226Ra]Ra needles are subjected to the generated x-ray beams to produce [225Ra]Ra (t1/2 = 14.9 d), which decays to give [225Ac]Ac according to the following designed and experimentally tested photonuclear reaction [29,37,43]:
226 Ra( γ , n )225Ra 225 Ac
Hence, 20 mg of [226Ra]Ra is placed at a distance of 12.5 cm from the bremsstrahlung target. It is then irradiated for one hour using 18 MeV bremsstrahlung x-rays and with an average current of 26 µA [44], resulting in the production of 2.44 MBq (66 µCi) of [225Ac]Ac. Based on experimental results, [225Ac]Ac production can reach 48 GBq (1.3 Ci) per month using 1 g of [226Ra]Ra target. In this regard, different electron accelerators—for instance, the Advanced Rare Isotope Laboratory (ARIEL) facility at TRI-University Meson Facility (TRIUMF, Vancouver, BC, Canada) in Canada and Accélérateur Linéaire et Tandem à Orsay (ALTO) in France—can be used for the production of [225Ac]Ac using this technique.
ARIEL is used as an Isotope Separation On-Line (ISOL) facility for radioisotope production via a photofission reaction at 35 MeV and an electron beam of 10 mA. It is mainly for research requirements rather than for medical utilization [45]. The experimental measurements show the possibility of producing 74 TBq (2000 Ci) of [225Ac]Ac per month with the irradiation of 1 g of [226Ra]Ra target at a higher current and with the use of a different irradiation geometry. Nonetheless, the main challenge that faces production via this procedure is the difficulty for the irradiated target to bear a 100 kW beam, as the available ARIEL ISOL targets can only survive up to 50 kW. In ALTO, theoretically, [225Ac]Ac can be produced with a radioactivity level of approximately 56 GBq (1.5 Ci) per month with the use of an electron energy of 50 MeV and 10 A as a beam current [45].
The main advantages of this production method are the generation of pure [225Ac]Ac radionuclide and the absence of co-production of any other actinium isotopes. [224Ra]Ra is a production by-product and will decay to inert [220Rn]Rn (t1/2 = 55.6 s). Moreover, the production yield can be elevated with better optimization of the irradiation parameters. However, the large-scale production of [225Ac]Ac by medical linacs still faces some difficulties, such as high cost and the lack of suitable infrastructure [39].
[225Ac]Ac Production via Proton Accelerators
Production of [225Ac]Ac by low-energy proton accelerators was first proposed in 2005 according to the nuclear reaction 226Ra(p,2n)225Ac [46]. This reaction is characterized by its high cross-section (710 mb) at 16.8 MeV, which facilitates its application at various low-energy cyclotrons that are already available worldwide and used to produce a large number of medical radioisotopes. Globally, over 550 cyclotrons have energy greater than 16 MeV, and many of them can be operated at a higher beam current up to 500 µA [47]. Based on the fact that the proton bombardment of [226Ra]Ra targets possesses a high cross-section, this approach can be used for large-scale production of [225Ac]Ac, which can cover the prospective future medical demand of [225Ac]Ac with the sole use of one production site. Theoretically, 4 TBq (108 Ci) of [225Ac]Ac can be produced monthly by irradiating 1 g of [226Ra]Ra target with a single 20 MeV, 500 µA proton beam [46]. This route does not generate [227Ac]Ac (t1/2 = 21.77 y) as a side-product, which is a highly toxic actinium isotope. Furthermore, while the (p,n) reaction is expected to produce some [226Ac]Ac, the co-production measurements of [226Ac]Ac have not been reported [46]. A simple experiment using the FLUKA model shows the appearance of [226Ac]Ac (t1/2 = 29 h) at the end of bombardment (EOB) with an activity level close to 11% of the total [225Ac]Ac yield. Nevertheless, this undesired amount of [226Ac]Ac is expected to decrease with time due to the differences in their half-lives [48]. Moreover, the co-production of [225Ra]Ra through the 226Ra(p,pn)225Ra reaction is negligible at optimal energies designed for direct [225Ac]Ac production.
[225Ac]Ac Production via Nuclear Reactors
Lately, nuclear reactors have become the primary source for medical radioisotope production [49]. Nonetheless, [225Ac]Ac production using nuclear reactors is still limited. They can be employed for the production of 225Ra by bombarding [226Ra]Ra targets with high-energy neutrons (>6.4 MeV) according to the following nuclear reaction.
226Ra(n,2n)225Ra
However, these high-energy neutrons are only present at the tail of the neutron energy spectrum in a typical breeder reactor. Calculated estimations indicate the possibility of producing MBq to GBq (mCi to Ci) quantities of [225Ac]Ac every month through the irradiation of 1 g of [226Ra]Ra target at a single reactor facility [50]. However, it is worth mentioning that the presence of lower-energy neutrons may result in the co-production of [227Ra]Ra, which, because of its toxic effects, can limit the practical use of [225Ac]Ac as a medical radioisotope. Moreover, the low [225Ra]Ra outcome is not satisfactory, especially with the high cost and the difficulties of safe handling of large [226Ra]Ra targets. Due to these critical difficulties, this method has not been experimentally implemented as of yet.
Challenges Encountered in the Use of Radium Targets
A profitable radioisotope production procedure strongly necessitates a stable, naturally occurring target material, or even a long-lived isotope, to cover the production cost. For this reason, long-lived [226Ra]Ra radionuclide (t1/2 = 1600 y) was chosen as a target material for [225Ac]Ac supply in all previously mentioned production techniques.
Despite the potential impact of using [226Ra]Ra targets, their utilization is accompanied by considerable challenges. These limitations are related to its abundance and underlying safety measures that complicate the production, processing, irradiation, and recycling of [226Ra]Ra targets. The handling of [226Ra]Ra targets is strongly associated with a number of safety hazards, as [226Ra]Ra is a highly radiotoxic isotope, very reactive with both air and water, and its decay generates the gaseous [222Rn]Rn daughter radionuclide [51]. Moreover, radium targets are most commonly used as sealed radioactive sources (SRS) wrapped in platinum. The [226Ra]Ra decay chain involves five subsequent alpha decays that produce helium, which allows gaseous build-up inside the [226Ra]Ra sealed source and may result in source rupture. In addition, high radiation exposure doses are estimated at 8.1 mSv/h at 1 m from 37 GBq (1 Ci, 1 g) of [226Ra]Ra, which are caused by hard gamma rays that are also emitted. These limitations represent the primary roadblocks towards the regular use of [226Ra]Ra targets.
In 1996, the International Atomic Energy Agency (IAEA) set some regulations for the disposal of [226Ra]Ra sources in protracted geological repositories [51]. These regulations strongly restricted the attainability of large [226Ra]Ra quantities. To cover the increasing [225Ac]Ac demand, new [226Ra]Ra sources are required, which may be excavated from the waste of uranium quarrying processes. Nearly 257 mg of [226Ra]Ra can be extracted from one ton of uranium ore, which could be upgraded to 12.85 kg of [226Ra]Ra annually.
The optimum use of [226Ra]Ra sources for a large-scale [225Ac]Ac production still requires far more infrastructure than that applied for routine medical isotope production. In addition, as previously mentioned, the necessary safety measures for production would require complementary infrastructure during [226Ra]Ra irradiation and processing.

3.2.2. [225Ac]Ac Production Using High Energy Proton Spallation of Thorium

The production of [225Ac]Ac can alternatively be achieved by the irradiation of thorium metal targets with proton energies greater than 70 MeV and, in some cases, >100 MeV [38,52], according to the following nuclear reaction.
232 Th(p,x)225Ac
This technique has proposed yields of micro- to milli-Curie quantities of [225Ac]Ac. These results have been verified by the Institute for Nuclear Research of the Russian Academy of Sciences (INR), American and Russian research groups at Brookhaven National Laboratory (BNL), and the Los Alamos National Laboratory (LANL) [31,32,35,41,42,53,54,55]. Compared to [226Ra]Ra targets, thorium is more radiologically safe and more widely available as a target material. Different thorium sources are currently available, as several countries produce tens of kilograms of thorium. Rare earth mining provides hundreds of tons of thorium annually as a by-product [56]. The availability of thorium sources provides the opportunity for cost savings related to target reprocessing procedures. Moreover, some of the current accelerators are well-equipped with the necessary infrastructure for target production, bombardment, and processing.
The major challenge facing this production method is the generation of additional isotopes other than [225Ac]Ac. In this case, post-irradiation separation processes will be required to recover the [225Ac]Ac radionuclide, which significantly elevates the production cost. For instance, the generation of 227Ac at the end of bombardment with an activity percentage of 0.1–0.2% strongly limits the use of [225Ac]Ac for clinical requirements. An alternative method for producing [225Ac]Ac is the 232Th(p,x)225Ac reaction, which utilizes high-energy protons (>100 MeV). This technique represents a promising avenue for exploring production options beyond the traditional [229Th]Th/[225Ra]Ra/[225Ac]Ac generator method. While only a limited number of facilities currently offer proton beams at the energy levels required for this nuclear reaction, this presents an opportunity to invest in and develop more facilities capable of producing these high-energy beams. Furthermore, the challenges of isolating [225Ac]Ac from the bulk thorium target and the various co-produced spallation products can be addressed through advancements in chemical separation techniques. It is important to consider the co-production of long-lived actinium-227 (227Ac), which has a half-life of 21.77 years. By focusing on these aspects, we can enhance the efficiency and feasibility of this production method [52]. To overcome this problem, most facilities tend to modify the current target processing method [57] by separating radium from the irradiated thorium a few days after the end of bombardment (EOB) to build a [225Ra]Ra/[225Ac]Ac generator. Some radium isotopes are also produced, such as [228Ra]Ra, [226Ra]Ra, [225Ra]Ra, [224Ra]Ra, and [223Ra]Ra, with half-lives of 5.7 y, 1600 y, 14.9 d, 3.6 d, and 11.4 d, respectively. Only [228Ra]Ra and [225Ra]Ra decay to actinium isotopes. Hence, using this combination as a radium/actinium generator can provide [225Ac]Ac free from toxic [227Ac]Ac. [228Ac]Ac (t1/2 = 6.2 h) will appear with an activity percentage of 0.88% after the separation of 225Ra/[225Ac]Ac. After an optimal [225Ac]Ac elution, [228Ac]Ac would normally decline with time [57]. [228Ac]Ac decays to [228Th]Th, from which [225Ac]Ac can be separated with a minor amount of impurities.
3.2.3. [225Ac]Ac Production via Special Facilities
The promising therapeutic applications of [225Ac]Ac have a significant impact on establishing several nuclear facilities to overcome its limited availability. In 2000, the Isotope Separator and Accelerator Facility (ISAC) in TRIUMF, Canada, was commissioned and used to produce radioisotopes for specific research purposes. This facility is based on the bombardment of thorium or uranium targets with protons (480 MeV) to produce radioisotopes that can be extracted in a heterogeneous ion beam. These isotopes are then separated according to their mass by Isotope Separation On-Line (ISOL) techniques, producing a consistent isobaric ion beam [58]. [225Ra]Ra and [225Ac]Ac isotopes are separated by isolating the mass of 225, which are then implanted onto an aluminum target, followed by their extraction on a solid-phase resin that acts as a [225Ra]Ra/[225Ac]Ac generator and provides [225Ac]Ac by the decay of the parent [225Ra]Ra [59]. [225Ac]Ac production by ISAC has covered TRIUMF’s needs for [225Ac]Ac for radiolabeling investigations and preclinical studies. In 2016, the ISAC facility provided 44.4 MBq (~1.2 mCi), which can be theoretically scaled up to 370 MBq (~10 mCi) of [225Ac]Ac monthly.
TRIUMF also launched an isotope production facility (IPF) for [225Ac]Ac production targets by proton (500 MeV, 120 µA) bombardment of thorium targets using its primary beamline (BL1A). In 2017, IPF produced 370 MBq (~10 mCi) of [225Ac]Ac, reaching 3700 MBq (~100 mCi) by 2018 [45].
The ISOL technique to produce [225Ac]Ac has also been applied at ISOLDE at the European Organization for Nuclear Research (CERN) in Geneva, Switzerland. The ISOL technique provides pure [225Ac]Ac free from other actinium isotopes. However, the production yields remain unsatisfactory compared to the quantities produced by [229Th]Th generators. In 2017, CERN also started a new radioisotope production facility called MEDICIS, which produces mass-separated radioisotope beams using offline targets, planning to use uranium targets for [225Ac]Ac production. The expected production capacity of MEDICIS could yield 112 MBq (3 mCi) of [225Ac]Ac monthly.
3.2.4. Transformation of [226Ra]Ra to [229Th]Th via Thermal Neutron Irradiation
Establishing alternative routes to increase [229Th]Th hoards would subsequently resolve the [225Ac]Ac supply shortage. This can be achieved by the irradiation of [226Ra]Ra targets with intense neutron fluxes in a reactor according to the following neutron capture reactions.
226 Ra(n,γ)227Ra
227 Ac(n,γ)228Ac
228 Th(n,γ)229Th
These reaction pathways have been tested in the High-Flux Isotope Reactor (HFIR) at ORNL, focusing on required target development and anticipated production outcomes [60]. The resulting amount of [229Th]Th produced is ~2.8 MBq per one gram of radium in a single day of irradiation [61]. Nevertheless, this procedure generates long-lived 228Th (t1/2 = 1.9116 a), which can be very difficult to separate from [229Th]Th. Moreover, as stated before in Section Challenges Encountered in the Use of Radium Targets, the use of [226Ra]Ra targets results in the evolution of the gaseous 222Rn daughter, causing serious difficulties with handling and managing radium targets [21,62]. All of these described production methods are briefly listed in Table 1.

4. [225Ac]Ac/[213Bi]Bi Radionuclide Generator

[225Ac]Ac/[213Bi]Bi radionuclide generator is the leading source for providing short-lived [213Bi]Bi [52]. Because of the half-life of the parent, [225Ac]Ac (t1/2 = 9.9 d), the working lifetime of the [225Ac]Ac/[213Bi]Bi generator is estimated to be several weeks. The generated daughter, [213Bi]Bi (t1/2 = 45.6 min), is known for its high specific activity regardless of the negligible percentage of the long-lived 209Bi (t1/2 = 2.01 × 1019 y) that may be present in the [213Bi]Bi eluate. Generally, according to the level of radioactivity needed per patient dose, [225Ac]Ac/[213Bi]Bi generators are usually eluted every two to three hours in medical practice, as can be seen in Figure 2, showing that [213Bi]Bi yield can reach 83% to 93% [52].
Many [225Ac]Ac/[213Bi]Bi generators have been developed depending on extraction chromatography or cation and anion exchangers [44,49,63,64,65,66,67,68,69]. Among these generators, [225Ac]Ac/[213Bi]Bi generators that are based on AG MP-50 cation exchange resin are the most commonly used systems [27,44,68,70], and they have been used in all [213Bi]Bi patient studies [30]. The Joint Research Centre (JRC) of the European Commission in Karlsruhe, Germany, has improved a high activity generator system that enables an efficient generator operation, especially when loaded with 4 GBq of [225Ac]Ac [28,71]. A crucial criterion of this generator system is the uniform activity distribution along nearly two-thirds of the generator resin to reduce the probable radiolysis of the organic resin and to ensure stable operation over several weeks. This generator has been effectively applied to prepare clinical doses of ~2.3 GBq [213Bi]Bi-Substance P analog for locoregional therapy for brain cancers [28].
An inverse generator protocol has also been proposed for the development of [225Ac]Ac/[213Bi]Bi generators. It is recommended because of its high radiation stability. In this method, the [213Bi]Bi that is generated is sorbed from [225Ac]Ac solution. Then, it is recovered for use in the intended application. Moreover, the design of these generators limits the co-production of 227Ac decay by-products. Based on this approach, the Pacific Northwest National Laboratory (PNNL) in the USA has improved multi-column selective inversion generators (MSID) [68]. This MSIG acts by the sorption of [213Bi]Bi on an anion exchanger, followed by its desorption from the column using sodium acetate buffer solution. Bray et al. developed a protocol for generator automization using sequential flow injection [66]. The main drawback of this procedure is the high percentage of [225Ac]Ac impurity in the product, which is ~ 0.1% of the total [225Ac]Ac activity. Furthermore, 2–3% of the parent [225Ac]Ac activity comes with each elution.
Betenekovet et al. proposed using annealed hydroxide inorganic sorbents with high radiation resistance [72]. These inorganic materials have shown high selective sorption affinity for [213Bi]Bi while [225Ac]Ac remains in the solution, favoring their use in the inverse generator design [72]. As a result, a novel [225Ac]Ac/[213Bi]Bi inverse generator was proposed and investigated, using a modified inorganic sorbent material consisting of zirconium and yttrium oxide mixtures that are thermally treated at 950 °C. The sorption of [213Bi]Bi was performed using a 0.1 M HNO3 solution, while the separation of [213Bi]Bi can be achieved using 1 M HCl. The optimal yield of [213Bi]Bi reached 98% using a 200 mg column, where almost 90% of [213Bi]Bi activity is concentrated in the first 0.3 mL of the eluate. Additionally, increasing the initial salinity of the [225Ac]Ac solution up to 3 M has increased the product purity to reach 0.01% of its activity.

Quality Control Evaluation of Produced [213Bi]Bi

Due to the short half-life of [213Bi]Bi, its quality control evaluation is preferably carried out directly after elution to prevent radioactivity loss. The radiochemical purity of labeled [213Bi]Bi antibodies is usually performed using ITLC within 5 min, whereas [213Bi]Bi labeled peptides can be examined using ITLC or Radio-HPLC.
Conversely, owing to the low activity ratio of [225Ac]Ac to [213Bi]Bi (~10−6) and the reduced emission probability of gamma photons (˂2%) from [225Ac]Ac decay, quick and direct measurement of the parent [225Ac]Ac breakthrough is not applicable simultaneously with [213Bi]Bi. Hence, the measurement of [225Ac]Ac activity is performed after 24 h by gamma spectrometry, and, consequently, an appropriate determination of [225Ac]Ac radioactivity in the eluted [213Bi]Bi is generally assured indirectly using another generator of the same type. Labeled peptides and antibodies with [213Bi]Bi are allowed to contain less than 1 ppm [225Ac]Ac [73].

5. Radiolabeling Strategies for [225Ac]Ac and [213Bi]Bi

The emission of multiple alpha particles from [225Ac]Ac decay (Figure 1) makes it an effective radioisotope in cancer therapy. However, these emissions make the delivery of [225Ac]Ac and its daughter a huge challenge. Because of the conservation of momentum, the energetic alpha-particle emission imparts high recoil energy to the daughter nucleus (>100 KeV). This energy is 1000 times more than the binding energy of any chemical bond [74,75], resulting in the escape of the daughter nuclide from the delivery vector due to the chemical bonds rupturing (Figure 3) and increasing the toxicity of healthy tissue because of the next α or β emission. The recoil daughter 221Fr was accumulated mainly in the kidneys, while [213Bi]Bi was distributed in the kidneys, urine, and other organs by ratios of 40, 30, and 30%, respectively [76].
The main strategies for overcoming the recoil effect are summarized as follows.

5.1. Fast Uptake in Target Tissues

The rapid uptake of the radiopharmaceutical by the target tissue and its rapid excretion from the body will decrease the recoil phenomena. Antibodies and peptides are beneficial for highly specific cell targeting, but antibodies take a long time to reach the tumor site [77]. Using smaller targeting agents will help to increase the stability of these complexes. Many researchers performed many experiments to make [225Ac]Ac safer during its use in treatment [78,79,80]. [225Ac]Ac-DOTA-J591 was preclinically studied on prostate carcinoma and showed a high affinity to the cancer cell, which decreases the recoil phenomenon at a low dose of administered activity [80]. Many other studies have been evaluated using antibodies and peptides to decrease the presence of [225Ac]Ac daughter nuclides in the healthy tissues.
Nevertheless, in all cases, the renal toxicity must be decreased, which could be accomplished by using a metal chelate and a diuretic. For example, DMPS, 2,3-dimercapto-1-propanesulfonic acid, complexes with [213Bi]Bi and suppresses its uptake in the kidneys. Secondly, chlorothiazide and furosemide, as diuretics, inhibit [221Fr]Fr absorption in the renal tissue [81].

5.2. Encapsulation of the Radionuclide into a Nano-Carrier

It is expected that the encapsulation of alpha-emitting radionuclides into a nanoparticle cavity keeps the daughter nuclide retained inside the cavity or structure of the nano-carrier and only permits the release of alpha particles to perform their therapeutic effect. Nonetheless, there are many harmful effects of using nanoparticles for this purpose. One limitation of this therapeutic dynamic is the accumulation of nanoparticle-containing alpha-emitters inside the liver during the hepatobiliary/intestinal execratory pathway (due to its large size), which may cause hepatic damage [77]. Some of the literature illustrates strategies for the formation of a [225Ac]Ac-nanoparticle system [82,83,84,85,86,87]. Many attempts have been carried out to overcome the recoil effect using the nanotechnology therapeutic idea. Liposomes, for example, have been used in sizes between 100 and 800 nm, giving maximum [213Bi]Bi retention of only 12% [88]. Another study that used polymeric particles with sizes of ~800 nm succeeded in retaining 70% of [213Bi]Bi inside the carrier [86]. Almost all of the [221Fr]Fr and ~70% of [213Bi]Bi were retained in carriers by using multi-walled nanoparticles. The outer shell was comprised of gold, whereas the inner shell was gadolinium phosphate, and the conjugated core is [La0.5Gd0.5([225Ac]Ac)]PO4-mAb-201b. However, the large size of the nano-construct is still an obstacle to proper application. Smaller-sized nanoparticles may help to improve the excretion properties of recoiled daughter radionuclides from the body; however, this will also allow for the recoil daughters to escape in very high amounts [85].

5.3. Local Administration of the Alpha-Emitters

The direct injection of alpha-emitters into the target site helps to overcome the recoil effect of the alpha-emitted radionuclide [45] and a series of studies have been performed using this approach [89,90,91,92,93,94]. Substance P, for example, has been combined with [213Bi]Bi-DOTA complex and injected locally into gliomas in a clinical study [93].

6. Chelating Agents for [225Ac]Ac and [213Bi]Bi

The use of [225Ac]Ac as TAT requires not only consideration for the radiometal but also for the chelating agent and the targeting moiety. The development of a suitable chelating agent for [225Ac]Ac needs careful evaluation of the chemistry of [225Ac]Ac. However, there is a notable lack of understanding of its coordination chemistry as Ac3+. Its large ionic radius generates another challenge for the production of suitable metal complexing ligands. The unique chelating agents must offer fast coordination kinetics under mild conditions and also provide in vivo thermodynamic stability of the formed complex [49,62,63,64,95]. DOTA (1,4,7,10-tetraazacyclododecan-1,4,7,10-tetraacetic acid) and its derivatives are the most suitable chelating agents for [225Ac]Ac as of now. The optimum radiolabeling parameters include the addition of 0.02 M DOTA to NH4OAc at pH 6. Then, the mixture is incubated for 2 h at 37 °C, giving a very high radiochemical yield of 99% [45]. Studies have shown very high stability of the complex in preclinical studies with promising results [49,78,96,97,98,99,100,101,102]. Figure 4 summarizes the most widely used chelating agents for [225Ac]Ac.
[213Bi]Bi can be linked in a stable manner to biomolecules through diethylene triamine penta-acetic acid (DTPA) derivatives or 1,4,7,10-tetraazacyclododecan-1,4,7,10-tetraacetic acid (DOTA). These metal complexing agents are also most suitable for labeling [213Bi]Bi with antibodies because of their fast and stable complexation at ambient temperature [49,63,64,109,110,111]. Due to its high stability, the [213Bi]Bi-DOTA-complex has also been used for conjugation to peptides. The synthetic protocol for the preparation of [213Bi]Bi-DOTA complex provides radiochemical yields (RCY) ≥99% using microwave heating for less than 5 min at 95 °C (pH 8–9). Many preclinical and clinical studies have been performed using [213Bi]Bi as TAT (Table 2).

7. Biological Studies of [225Ac]Ac/[213Bi]Bi-Radiopharmaceuticals

Although α-emitters are considered promising and efficient radiotherapeutic agents, many application barriers restrict their market accessibility, especially their high toxicity [130]. The preclinical evaluation of α-emitters is mandatory, and many more in vitro and in vivo studies are necessary for routine clinical usage [131]. Generally, α-particles with 5–9 MeV emission energy show a mean path length of 40–100 µm and LET of ~80 keV/µm [130]. Thereby, they are considered promising radiotherapeutic candidates for metastatic castration-resistant prostate cancer, relapsed or refractory CD-22-positive non-Hodgkin lymphoma, acute myeloid leukemia, neuroendocrine tumors, ovarian carcinoma, gliomas, intralesional and systemic melanoma, colon cancer, and bone metastases [132]. A complete summary of the current progress of preclinical and clinical evaluations of [225Ac]Ac and [213Bi]Bi for the development of novel targeted alpha therapies for the treatment of different solid tumors is therefore necessary as a section of this body of work. Trials have been conducted to evaluate [225Ac]Ac/[213Bi]Bi radiopharmaceuticals on the preclinical scale, and some of these agents have been moved into phase 0, I, and II clinical investigations. For example, [225Ac]Ac-PSMA-617 was preclinically evaluated for prostate cancer therapy in PC-3/PC-3-PIP tumor-bearing mice [133]. Prostate-specific membrane antigen (PSMA) small molecules have also attracted attention in TAT, including PSMA-617 for prostate cancer therapy. The radio-conjugate [225Ac]Ac-PSMA-617 demonstrated improved therapeutic efficacy in PC-3/PC-3-PIP-tumor-bearing mice in comparison to [177Lu]Lu-PSMA-617 [133]. In addition, [225Ac]Ac-BC8 showed effective tumor control of tumors in a multiple myeloma tumor-bearing mouse model [134,135]. [225Ac]Ac-E4G10 has shown high efficacy for managing tumor growth in glioblastoma-bearing mice [136]. Moreover, [225Ac]Ac-PSMA-617 was clinically evaluated in patients with diffuse red marrow infiltration of mCRPC and peritoneal carcinomatosis with liver metastases. This radiopharmaceutical showed a drop in the PSA level for patients, confirming the effectiveness of the drug and the presence of treatment response [12].

7.1. Pre-Clinical Studies of [225Ac]Ac/[213Bi]Bi-Radiopharmaceuticals

Recent preclinical investigations of [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals indicate the potential of being used as effective TAT agents [137]. A key criterion in designing TAT agents is the proper selection of a molecular platform that achieves the best radionuclide delivery to the target tissue as well as the availability of chemically active groups suitable for conjugation of the chelator. Targeting platform selection relies heavily on some critical factors that include off-target binding and its bio-pharmacokinetics behavior [138]. In addition to the traditional delivery platforms (small molecules, peptides, and antibodies), the new pipeline of nano-sized platforms, such as self-assembled vesicular structures and inorganic nanoparticles, has evolved, resulting in improved targeting ability, inducing high efficacy for a new therapeutic delivery platform [139].

7.1.1. Traditional Delivery Platforms

Table 3 summarizes [225Ac]Ac/[213Bi]Bi conjugates that have undergone preclinical evaluation, demonstrating a therapeutic effect via the traditional delivery platforms that include peptides and antibodies.
Antibodies
Radio-immunotherapy (RIT) demonstrates the utilization of monoclonal antibodies (mAbs) that could be conjugated to chelating moieties in order to deliver radioactive payloads of particulate energy to cellular targets [138]. In vitro cytotoxicity investigations of [225Ac]Ac-DTPA-IgG1 and [213Bi]Bi-DTPA-IgG1 probes have been conducted on the human epidermoid A431 tumor cell line [156]. This study showed that the two probes have the ability to destroy tumor cells with a higher therapeutic propensity for the [213Bi]Bi as compared to the [225Ac]Ac probe. Another [225Ac]Ac-labeled antibody, [225Ac]Ac-DTPA-201B, has been investigated in vivo to target murine lung endothelial thrombomodulin [103]. Although the probe demonstrated potential lung targeting ability, it showed an unsatisfactory biological half-life in the target organ. Subsequently, another study utilized HEHA chelator for designing [225Ac]Ac-HEHA-201B to target lung carcinoma [142]. The in vivo distribution showed significant accumulation within the lung (300% ID/g) at 4 h post-injection and an improved biological half-life (49 h). However, regrettably, cleared [225Ac]Ac was accumulated primarily within the liver, spleen, and bone. Further studies have utilized CC49 antibody to formulate an [225Ac]Ac-HEHA-CC49 probe and its in vivo biodistribution has been investigated in a tumor-bearing mouse model (LS174T xenografts) post-subcutaneous (s.c) as well as -intramuscular (i.m.) delivery [152]. This demonstrated that the probe was accumulated in the tumor tissue with 25 and 8% ID/g at 24 h post s.c. and i.m. injection, respectively. Trastuzumab, an FDA-approved mAb for targeting HER2/neu in breast cancer malignancies, has been labeled with [225Ac]Ac, and its in vitro cytotoxicity was investigated against breast cancer spheroids MCF7, MDA-MB-361 (MDA), and BT-474 (BT) as a potential agent for RIT [157]. Recently, another study has been conducted on [225Ac]Ac-Trastuzumab for the effective treatment of ductal carcinoma [158]. Bioluminescent imaging post-intraductal administration of the drug revealed that [225Ac]Ac-Trastuzumab demonstrated better therapeutic efficacy in comparison to the intravenous route of administration by achieving systemic toxicity bypass. Another FDA-approved mAb, anti-VE-cadherin antibody (E4G10), was investigated as a potential TAT agent ([225Ac]Ac-E4G10) for targeting the vascular endothelium of glioblastoma with tumor growth-controlling ability in a mouse model [159]. Anti-CD45 antibodies have also been employed as targeting ligands for TAT [160]. Another study has utilized [225Ac]Ac-BC8 radio-conjugate as a potent targeting agent into organs of the immune system, especially bone marrow [161]. [213Bi]Bi radionuclide has been effectively used for TAT, where it was conjugated to the anti-CD33 antibody construct (HuM-195), establishing a [213Bi]Bi-HuM-195 radio-conjugate [162]. Another study has utilized 9E7.4 anti-CD138 antibodies as targeting moieties in TAT based on [213Bi]Bi [126]. The study revealed that [213Bi]Bi-9E7.4 anti-CD138 is more effective than [177Lu]Lu-9E7.4 anti-CD138 in mice bearing 5T33 multiple myeloma tumors, where there is a significant increase in the mice survival at 45%. A subsequent study utilized J591 anti-PSMA mAb for designing a TAT agent ([213Bi]Bi-J591) for the treatment of prostate cancer [163]. The study revealed that [213Bi]Bi-J591 is a potential cytotoxic agent against LNCaP spheroids in vitro and in vivo.
Peptides
Since octreotide, an analog of the endogenous peptide hormone somatostatin, was approved by the FDA, peptide receptor radionuclide therapy (PRRT) has gained much focus from many researchers around the globe [138]. However, PRRT agents are facing a heightened limitation in that they are compromised by nephrotoxicity [164]. The first reported literature about integrating octreotide analogs as targeting TAT moieties was demonstrated in 2006 [154]. This study revealed that the octreotide analog (DOTATOC), when radiolabeled with the high-LET α-emitter [213Bi]Bi, showed promise in in vivo distribution, toxicity, safety profile, and therapeutic efficacy in a rat pancreatic carcinoma model. The results demonstrated that [213Bi]Bi-DOTATOC radio-conjugate accumulated specifically in tissues rich with somatostatin receptors. It also showed dose-dependent antitumor efficacy in tumor reduction when rats were treated with >20 MBq of [213Bi]Bi-DOTATOC in comparison to the rodents receiving <11 MBq of conjugate. Interestingly, the study revealed little nephrotoxicity at dosing levels of radioactivity from 4 to 22 MBq. Subsequently, [225Ac]Ac-DOTATOC radio-conjugate was developed, and the in vivo distribution was investigated against pancreatic neuroendocrine tumors in a xenografted mouse model [78]. The results showed good accumulation of [225Ac]Ac-DOTATOC in neuroendocrine xenografted tumors with improved therapeutic efficacy at activity greater than 30 kBq of the drug. Recently, preclinical investigations have been conducted on [213Bi]Bi-DOTATATE to evaluate the therapeutic efficacy in athymic mice pre-treated with and without L-lysine as a renal protecting agent [165]. The results demonstrated a good tumor accumulation in both cases, while the renal uptake in the case of L-lysine pre-treated mice showed half the accumulation compared to the non-treated mice. The maximum tolerated dose (MTD) of [213Bi]Bi-DOTATATE increased from 13.0  ±  1.6 MBq in the non-treated mice to 21.7  ±  1.9 MBq in L-lysine pre-treated mice. An additional investigation has evaluated the therapeutic efficacy of [213Bi]Bi-DOTATATE against different-sized (small and large) tumor lesions utilizing two tumor models: H69 (human small cell lung carcinoma) and CA20948 (rat pancreatic tumor) [166]. The results revealed greater tumor accumulation of conjugate (19.6 ± 6.6% ID/g) in the case of the CA20948 model as compared to 9.8 ± 2.4% ID/g in the case of H69. Nevertheless, the antitumor efficacy of [213Bi]Bi-DOTATATE was greater in the case of the H69 tumor model, with a higher survival rate. Also, it revealed that no tumor regrowth was observed in the case of small tumor groups (H69 model) [165].

7.1.2. Advanced Delivery Platforms

The application of nano-constructs as delivery platforms in TAT has received significant attention from researchers and clinicians around the world. It provides the possibility of overcoming the difficulties associated with traditional approaches, which include the recoil effect of the daughter radionuclide and the insufficiency of efficient chelating agents for complexing α-radionuclides to the molecular target [167]. Table 4 summarizes the preclinical studies of [225Ac]Ac/[213Bi]Bi-nanoradiopharmaceuticals.
Self-Assembled Vesicles
  • Liposomes
Multi-vesicular liposomes (MUVELs) have shown potential usefulness for the appropriate delivery of [225Ac]Ac radionuclide to ovarian cancer cells [75,88]. This study demonstrated the conjugation of MUVELs with HER2/neumAb (Trastuzumab) and evaluated the therapeutic efficacy in vitro against ovarian cancer cells. The results revealed an appropriate reduction in the recoil effect of the daughter [213Bi]Bi radionuclide (17–18% after 20 days). Also, biological investigations revealed high-affinity binding to ovarian cancer cells with improved cell uptake kinetics. Another study has demonstrated the usage of [225Ac]Ac-labelled liposomes for the treatment of prostate cancer [82]. The authors of this study demonstrated appropriate [225Ac]Ac radionuclide loading to J591 mAb and the A10 aptamers conjugated to a liposome construct to target the PSMA protein. The in vitro investigation revealed that [225Ac]Ac-J591-liposomes had exhibited a greater cytotoxic effect than [225Ac]Ac-A10 aptamers-liposomes, encouraging the significance of [225Ac]Ac-J591-liposomes as a PSMA-TAT agent. Another study conducted the utilization of [213Bi]Bi-radiolabeled immunoliposomes as a potential agent for the treatment of early-stage metastases [177]. [213Bi]Bi was conjugated to liposomal-CHX-A-DTPA nano-construct (100 nm), and the therapeutic efficacy was evaluated against a rat/neu transgenic mouse model of mammary carcinoma. The results revealed an increase in the survival time of 38 days as compared to 29 days for the [213Bi]Bi-CHX-A-DTPA radio-conjugate.
  • Polymersomes
Compared to liposomes, polymersomes (self-assembled polymeric amphiphiles) can demonstrate higher stability and reduced permeability in human tissue. Consequently, they have become a potential platform to reduce the recoil effect of the daughter radionuclides [171]. The first attempt to formulate polymersomes encapsulated with [225Ac]Ac radionuclide to reduce the recoiling effect demonstrated that double-layered polymersomes (400 nm) achieved 47% of [213Bi]Bi recoil reduction, while the single-layered derivative achieved only 10–20% [86]. Furthermore, it was postulated that increasing the polymersome size to 800 nm could reduce the [213Bi]Bi recoils by 80%. The therapeutic efficacy of [225Ac]Ac-labeled polymersomes has been investigated against U87 human glioma spheroids. An appropriate distribution of polymersome was achieved after only four days [169]. This study showed that 0.1 kBq [225Ac]Ac-labeled polymersome caused a reduction in spheroidal growth and that increasing the activity to 5 kBq caused complete tumor death after only two days. An additional investigation has compared the recoil effect of [225Ac]Ac-DTPA with [225Ac]Ac-doped-InPO4 immobilized polymersome (200 nm) [170]. The in vivo distribution post-intra-tumor (i.t.) injection in mice bearing MDA-MB-231 tumors revealed high activity accumulation of the polymersome in the tumor lesions, while [225Ac]Ac-DTPA was mainly accumulated and excreted rapidly through the renal–urinary excretion pathway. Recoil studies showed that [213Bi]Bi was retained in the tumor lesion in the case of the polymersome formulation, while more [213Bi]Bi was retained in the kidney in the case of [225Ac]Ac-DTPA complex. Studies conducted using sticky lipid vesicles with clustered HER2-targeting peptides loaded with [225Ac]Ac radionuclide for the treatment of breast cancer [175] have demonstrated that the sticky vesicles have a propensity to target HER2-negative breast cancer cells (MDA-MB-231 and MCF7) with some cytotoxic effect of 42–61%.
Carbon Nano-Construct
As an alternative delivery platform, carbon nanotubes (CNTs) for the immobilization of [225Ac]Ac radionuclide have been employed by Ruggiero et al. [178]. The study developed DOTA-CNTs functionalized with E4G10 antibody for targeting the tumor vascular endothelial-cadherin. The in vivo investigation of [225Ac]Ac- CNT-DOTA-E4G10 against mice induced with LS174T (colon) adenocarcinomas showed appropriate tumor reduction with a two-fold higher survival rate compared to controls. In another study involving CNTs for the targeted delivery of [225Ac]Ac radionuclide to the tumor, a pre-targeting approach was employed using mAb-MORF [174]. [225Ac]Ac-DOTA-CNT modified with mAb-MORF was developed, and in vitro investigations showed high-affinity binding to cancer cells. The in vivo evaluation in a lymphoma xenografted mouse model indicated improved therapeutic efficacy in a two-step approach, where complete tumor regression was achieved with rapid clearance.
Inorganic Nanoparticles
The utilization of lanthanum phosphate nanoparticles as a potential delivery platform for [225Ac]Ac radionuclide to target tumor lung vasculature with a high recoil effect reduction has provided some interesting results as well [75,172]. Here, the authors developed [225Ac]Ac-doped lanthanum phosphate nanoparticles functionalized with mAb 201B. In vivo studies for the i.v. administration of La-[225Ac]Ac-PO4 NPs-mAb201B in mice showed rapid and high accumulation within the lungs (30% of the total injected dose at 1 h post-injection). In addition, the authors declared that the lung had retained about 50% of the recoiled daughter radionuclides.
A study demonstrating the loading of [225Ac]Ac to gold-decorated lanthanide phosphate nanoparticles functionalized with mAb 201b antibody to produce [225Ac]Ac-(La0.5Gd0.5)PO4-mAb 201b aimed to reduce the recoil effect of the daughter radionuclides and to target tumor lung vasculature [85]. The in vivo distribution analysis of this new agent showed high accumulation within lungs, with 69% [213Bi]Bi retention in lung tissue with only 2.8% within the kidney at 1 h post-injection. Cedrowska et al. reported the demonstration of TiO2 nanoparticles functionalized with silane-PEG-SP(5-11) conjugates and labeled with [225Ac]Ac radionuclide to produce [225Ac]Ac-TiO2-silane-PEG-SP(5-11) as a potential TAT-nano-construct for brain tumors. This study showed appropriate cytotoxic action against glioblastoma cancer cells as revealed during in vitro investigations [173].
Gold nanoparticles are another inorganic delivery platform that has been used to design [225Ac]Ac-radiopharmaceuticals. Recently, Salvanou et al. have produced [225Ac]Ac-Au-TADOTAGA for localized cancer treatment [176]. This in vitro investigation against U87 glioblastoma cancer cells showed that [225Ac]Ac-Au-TADOTAGA demonstrated a cytotoxic effect with more than 70% cell killing after 48 h. Additionally, in vivo investigation revealed high accumulation (60.67 ± 3.87% ID/g) of drug within the tumor lesion at 2 h post i.t. injection, which decreased slowly over time.

7.2. Clinical Trials of [225Ac]Ac/[213Bi]Bi-Radiopharmaceuticals

Table 5 summarizes most of the [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals that have been evaluated preclinically or under clinical phase investigations [11,14,22,36,106,109,114,121,122,123,133,134,135,136,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194]. Clinical trial phase investigations for [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals have begun and have been evaluated in ~1000 patients suffering from many different tumors. These studies have been largely based upon radiolabeling of cell-targeting antibodies, although other platforms, as described herein, have also taken precedence. Examples of several of these agents are listed below.
  • [225Ac]Ac-PSMA -617 was evaluated clinically in ~500 prostate tumor patients. It was used with an average dose of 100 kBq/kg body weight. It showed excellent radiotherapeutic action with an extensive decrease in prostate-specific antigen (PSA) [121,187,188,191,193,195,196].
  • [213Bi]Bi-anti-CD33-mAb was therapeutically efficient in clinical trials in ~150 leukemia patients. It reached phase I and II stages, and showed a safe dose of up to 3 µCi/kg with no acute toxicity effects [106].
  • [213Bi]Bi- Substance P has been used in ~100 glioma patients, resulting in promising extensive radiotherapeutic action with a dose of about 2 GBq in 2-month intervals [93,190,194,197].
  • [213Bi]Bi-anti-MCSP-mAb was investigated in ~65 melanoma patients to evaluate its ability to deliver an appropriate radiation load to the tumor. However, its pharmacokinetic behavior failed to reach the required therapeutic dose [122,123].
  • [213Bi]Bi-DOTATOC was tested in ~50 neuroendocrine tumor patients to evaluate its toxicity. It showed safe usage with radioactivities of 18.5 MBq in two-month dosing intervals. However, hepatotoxicity was a limiting parameter for use [14].
  • [225Ac]Ac/[213Bi]Bi-anti-EGFR-mAb was used via the intra-vesical route of administration in ~12 bladder cancer patients. It extensively reduced the toxicity to neighboring tissues with excellent therapeutic activity at the site of application [109].

8. Conclusions

This article underlines the potential role of alpha emitters, especially [225Ac]Ac and [213Bi]Bi radionuclides, as potential leaders for future TAT. Undoubtedly, because of their short penetration length in human tissues and their high LET, alpha emitters offer a therapeutic opportunity for patients who do not respond to conventional therapies. Preliminary studies have shown the advantageous utilization of [225Ac]Ac and [213Bi]Bi radiopharmaceuticals for cancer therapy. However, some challenges still remain and need to be carefully explored and resolved. For example, new and competitive production techniques have to be designed and tested to provide a sufficient supply of both radionuclides. In addition, their chemical and coordination behavior still requires resolution via extensive research considerations. Moreover, the scientific community must conduct additional clinical and preclinical studies on a larger scale in order to provide for more reliable and consistent results for using [225Ac]Ac and [213Bi]Bi labeled radiopharmaceuticals if routine usage for cancer treatments is to become relevant. Clinical applications of [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals are still in the early phases of evaluation but still face many challenging issues that scientists are working intensively to resolve. The success of solving these complex research problems and achieving these goals will support TAT, introducing high LET radiation to tumor cells, producing a high payload of cytotoxic effect on them. Lastly, it is worth mentioning that the use of [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals is still in its infancy, which means that growth is needed in terms of the production capacity and clinical application scales necessary for success. The scientific community has overcome and achieved many of these milestones in recent years. However, undoubtedly, there is still room to continue seeking a high level of research effort. It is pertinent to point out that continuous research cooperation on a global scale can exponentially increase the growth of using [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals as influential TAT leaders in the very near future.

Funding

This work was supported by VA Merit Award 1I01BX003392 and VA Research Career Scientist Award IK6BX006810, Fulbright Fellowship.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

All authors declare no conflict of interest.

Abbreviations

AcActinium
PoPolonium
TlThallium 
RnRadon
UUranium
BqBecquerel 
keVKilo electron volt
Yyear
Hhour
Aalpha
t1/2Half-life
WWatt 
NNeuton
CK-CENNuclear- Energy Research Centre
IPFIsotope production facility
KIRAMSKorean Institute for Radiological and Medical Sciences
CNEAArgentine National Atomic Energy Commission
MSIGmulticolumn selective inversion generators 
INRInstitute for Nuclear Research of the Russian Academy of Sciences
BNLBrookhaven National Laboratory
linaclinear electron accelerator
ITUInstitute for Trans-Uranium Elements
ISOLIsotope Separation On-Line
SRSsealed radioactive sources
USAUnited States of America 
ITLCInstant thin-layer chromatography
SPECTSingle-photon emission computed tomography
TATTargeted alpha-particle therapy
LETLinear energy transfer
DTPADiethylenetriaminepenta acetic acid
DMPS2,3-Dimercapto-1-propanesulfonic acid
PEPA1,4,7,10,13-pentaazacyclopentadecane-N,N′,N′′,N′′′,N′′′′-pentaacetic acid
DOTPA1,4,7,10-tetraazacyclododecan-1,4,7,10-tetrapropionic acid
p-SCN-Bn-DOTAS-2-(4-Isothiocyanatobenzyl)-1,4,7,10-tetraazacyclododecane tetraacetic acid
DTPA-p-Bn-NCSS-2-(4-Isothiocyanatobenzyl)-diethylenetriamine pentaacetic acid
2B-DOTA-NCS2-(p-isothiocyanatobenzyl)-1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid
SCN-CHX-A-DTPA2-(4-isothiocyanatobenzyl)diethylenetriamine pentaacetic acid
CHX-A″-DTPAA′′ isomer of N-[(R)-2-amino-3-(4-nitrophenyl)propyl]-trans-(S,S-cyclohexane-1,2-diamine-N,N,N′,N′′,N′′-pentaacetic acid
MeO-DOTA-NCSα-(5-isothiocyanato-2-methoxyphenyl)-1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid
HEHA-NCS2-(4-isothiocyanatobenzyl)-1,4,7,10,13, 16-hexaazacyclohexadecane- 1,4,7,10,13,16-hexaacetic acid
Bispa2(6,6′-((9- hydroxy-1,5-bis(methoxycarbonyl)-2,4-di(pyridin-2-yl)-3,7-diazabicyclo [3.3.1]nonane3,7-diyl)bis(methylene))dipicolinic acid)
SCN-CHX-A′′-DTPA or CHX-A′′-DTPA-p-Bn-NCS[(R)-2-Amino-3-(4-isothiocyanatophenyl)propyl]-trans-(S,S)-cyclohexane1,2-diamine-pentaacetic acid
BiBismuth 
FrFrancium
PbLead
RaRadium
NPINuclear Physics Institute
CiCurie
MeVMega electron volt
dDay
sSecond
γGamma
βBeta
ppmPart per million
pProton 
JRCJoint Research Centre
PNNLPacific Northwest National Laboratory
ISACIsotope Separator and Accelerator Facility
CERNEuropean Organization for Nuclear Research
QSTNational Institutes for Quantum and Radiological Sciences and Technology
IPPELipinski Institute for Physics and Power Engineering
LANLLos Alamos National Lab
ARIELAdvanced Rare IsotopE Laboratory
ORNLOak Ridge National Laboratory
IAEAInternational Atomic Energy Agency
HFIRHigh-Flux Isotope Reactor
EOBend of bombardment
HPLCHigh-pressure liquid chromatography
PCPaper chromatography
RCYRadiochemical-yield
EDTAEthylenediaminetetra acetic acid
DMSADimercapto succinic acid
DOTA(1,4,7,10-tetraazacyclododecan-1,4,7,10-tetraacetic acid)
DO3A1,4,7,10-Tetraazacyclododecane-1,4,7-triacetic acid
cDTPACyclic diethylenetriaminepenta acetic acid
TETPA1,4,8,11-tetraazacyclotetradecan-1,4,8,11-tetrapropionic acid
DOTATOC1,4,7,10-tetra-azacylododecane N,N′,N′′,N′′′-J-tetraacetic acid-Tyr3-octreotide
DOTMP(1,4,7,10-tetraazacyclododecan-1,4,7,10-tetramethylene-phosphonic acid)
DOTAGA2,2′,2′′-(10-(2,6-dioxotetrahydro-2H-pyran-3-yl)-1,4,7,10-tetraazacyclododecane-1,4,7-triyl)triacetic acid
HEHA1,4,7,10,13,16-hexaaazacyclohexadecane-N,N′,N′′,N′′′,N′′′′, N′′′′′-hexaacetic acid
Lpy(2R)-2-hydroxy-3-[[(S)-hydroxy[[(1S,2R,3R,4S,5S,6R)-2,3,4,5,6-pentahydroxy cyclohexyl]oxy]phosphoryl]oxy]propyl tetradecanoate

References

  1. Baidoo, K.E.; Yong, K.; Brechbiel, M.W. Molecular Pathways: Targeted α-Particle Radiation Therapy. Clin. Cancer Res. 2013, 19, 530–537. [Google Scholar] [CrossRef]
  2. Brechbiel, M.W. Targeted alpha-therapy: Past, present, future? Dalton Trans. 2007, 4918–4928. [Google Scholar] [CrossRef]
  3. Couturier, O.; Supiot, S.; Degraef-Mougin, M.; Faivre-Chauvet, A.; Carlier, T.; Chatal, J.F.; Davodeau, F.; Cherel, M. Cancer radioimmunotherapy with alpha-emitting nuclides. Eur. J. Nucl. Med. Mol. Imaging 2005, 32, 601–614. [Google Scholar] [CrossRef]
  4. Elgqvist, J.; Frost, S.; Pouget, J.P.; Albertsson, P. The potential and hurdles of targeted alpha therapy—Clinical trials and beyond. Front. Oncol. 2014, 3, 324. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, Y.S.; Brechbiel, M.W. An overview of targeted alpha therapy. Tumour Biol. 2012, 33, 573–590. [Google Scholar] [CrossRef]
  6. Maucksch, U.; Runge, R.; Oehme, L.; Kotzerke, J.; Freudenberg, R. Radiotoxicity of alpha particles versus high and low energy electrons in hypoxic cancer cells. Nuklearmedizin 2018, 57, 56–63. [Google Scholar] [CrossRef] [PubMed]
  7. Mulford, D.A.; Scheinberg, D.A.; Jurcic, J.G. The promise of targeted alpha-particle therapy. J. Nucl. Med. 2005, 46 (Suppl. 1), 199S–204S. [Google Scholar] [PubMed]
  8. Wilbur, D.S. Chemical and radiochemical considerations in radiolabeling with alpha-emitting radionuclides. Curr. Radiopharm. 2011, 4, 214–247. [Google Scholar] [CrossRef]
  9. McDevitt, M.R.; Sgouros, G.; Finn, R.D.; Humm, J.L.; Jurcic, J.G.; Larson, S.M.; Scheinberg, D.A. Radioimmunotherapy with alpha-emitting nuclides. Eur. J. Nucl. Med. 1998, 25, 1341–1351. [Google Scholar] [CrossRef]
  10. Sgouros, G.; Roeske, J.C.; McDevitt, M.R.; Palm, S.; Allen, B.J.; Fisher, D.R.; Brill, A.B.; Song, H.; Howell, R.W.; Akabani, G.; et al. MIRD Pamphlet No. 22 (abridged): Radiobiology and dosimetry of alpha-particle emitters for targeted radionuclide therapy. J. Nucl. Med. 2010, 51, 311–328. [Google Scholar] [CrossRef]
  11. Roncali, L.; Hindre, F.; Samarut, E.; Lacoeuille, F.; Rousseau, A.; Lemee, J.M.; Garcion, E.; Cherel, M. Current landscape and future directions of targeted-alpha-therapy for glioblastoma treatment. Theranostics 2025, 15, 4861–4889. [Google Scholar] [CrossRef]
  12. Kratochwil, C.; Bruchertseifer, F.; Giesel, F.L.; Weis, M.; Verburg, F.A.; Mottaghy, F.; Kopka, K.; Apostolidis, C.; Haberkorn, U.; Morgenstern, A. 225Ac-PSMA-617 for PSMA-Targeted alpha-Radiation Therapy of Metastatic Castration-Resistant Prostate Cancer. J. Nucl. Med. 2016, 57, 1941–1944. [Google Scholar] [CrossRef] [PubMed]
  13. van der Doelen, M.J.; Mehra, N.; van Oort, I.M.; Looijen-Salamon, M.G.; Janssen, M.J.R.; Custers, J.A.E.; Slootbeek, P.H.J.; Kroeze, L.I.; Bruchertseifer, F.; Morgenstern, A.; et al. Clinical-Prostate cancer Clinical outcomes and molecular profiling of advanced metastatic castration-resistant prostate cancer patients treated with Ac-PSMA-617 targeted alpha-radiation therapy. Urol. Oncol. Semin. Orig. Investig. 2021, 39, 729.e7–729.e16. [Google Scholar] [CrossRef]
  14. Kratochwil, C.; Giesel, F.L.; Bruchertseifer, F.; Mier, W.; Apostolidis, C.; Boll, R.; Murphy, K.; Haberkorn, U.; Morgenstern, A. (2)(1)(3)Bi-DOTATOC receptor-targeted alpha-radionuclide therapy induces remission in neuroendocrine tumours refractory to beta radiation: A first-in-human experience. Eur. J. Nucl. Med. Mol. Imaging 2014, 41, 2106–2119. [Google Scholar] [CrossRef] [PubMed]
  15. Friesen, C.; Glatting, G.; Koop, B.; Schwarz, K.; Morgenstern, A.; Apostolidis, C.; Debatin, K.M.; Reske, S.N. Breaking chemoresistance and radioresistance with [213Bi]anti-CD45 antibodies in leukemia cells. Cancer Res. 2007, 67, 1950–1958. [Google Scholar] [CrossRef]
  16. Diamond, W.T.; Ross, C.K. Actinium-225 production with an electron accelerator. J. Appl. Phys. 2021, 129, 104901. [Google Scholar] [CrossRef]
  17. Rahman, A.K.M.R.; Babu, M.H.; Ovi, M.K.; Zilani, M.M.; Eithu, I.S.; Chakraborty, A. Actinium-225 in Targeted Alpha Therapy. J. Med. Phys. 2024, 49, 137–147. [Google Scholar] [CrossRef]
  18. Miederer, M.; Benesová-Schäfer, M.; Mamat, C.; Kästner, D.; Pretze, M.; Michler, E.; Brogsitter, C.; Kotzerke, J.; Kopka, K.; Scheinberg, D.A.; et al. Alpha-Emitting Radionuclides: Current Status and Future Perspectives. Pharmaceuticals 2024, 17, 76. [Google Scholar] [CrossRef]
  19. Mourtada, F.; Tomiyoshi, K.; Sims-Mourtada, J.; Mukai-Sasaki, Y.; Yagihashi, T.; Namiki, Y.; Murai, T.; Yang, D.J.; Inoue, T. Actinium-225 Targeted Agents: Where Are We Now? Brachytherapy 2023, 22, 697–708. [Google Scholar] [CrossRef]
  20. Miederer, M.; Scheinberg, D.A.; McDevitt, M.R. Realizing the potential of the Actinium-225 radionuclide generator in targeted alpha particle therapy applications. Adv. Drug Deliv. Rev. 2008, 60, 1371–1382. [Google Scholar] [CrossRef]
  21. Bruchertseifer, F.; Kellerbauer, A.; Malmbeck, R.; Morgenstern, A. Targeted alpha therapy with bismuth-213 and actinium-225: Meeting future demand. J. Label. Compd. Radiopharm. 2019, 62, 794–802. [Google Scholar] [CrossRef]
  22. Ferrier, M.G.; Batista, E.R.; Berg, J.M.; Birnbaum, E.R.; Cross, J.N.; Engle, J.W.; La Pierre, H.S.; Kozimor, S.A.; Lezama Pacheco, J.S.; Stein, B.W.; et al. Spectroscopic and computational investigation of actinium coordination chemistry. Nat. Commun. 2016, 7, 12312. [Google Scholar] [CrossRef] [PubMed]
  23. Kulikov, E.V.; Novgorodov, A.F.; Schumann, D. Hydrolysis of Actinium-225 Trace Quantities. J. Radioanal. Nucl. Chem. 1992, 164, 103–108. [Google Scholar] [CrossRef]
  24. Thiele, N.A.; Wilson, J.J. Actinium-225 for Targeted alpha Therapy: Coordination Chemistry and Current Chelation Approaches. Cancer Biother. Radiopharm. 2018, 33, 336–348. [Google Scholar] [CrossRef] [PubMed]
  25. Ramogida, C.F.; Orvig, C. Tumour targeting with radiometals for diagnosis and therapy. Chem. Commun. 2013, 49, 4720–4739. [Google Scholar] [CrossRef]
  26. Pomme, S.; Marouli, M.; Suliman, G.; Dikmen, H.; Van Ammel, R.; Jobbagy, V.; Dirican, A.; Stroh, H.; Paepen, J.; Bruchertseifer, F.; et al. Measurement of the 225Ac half-life. Appl. Radiat. Isot. 2012, 70, 2608–2614. [Google Scholar] [CrossRef]
  27. Morgenstern, A.; Apostolidis, C.; Kratochwil, C.; Sathekge, M.; Krolicki, L.; Bruchertseifer, F. An Overview of Targeted Alpha Therapy with (225)Actinium and (213)Bismuth. Curr. Radiopharm. 2018, 11, 200–208. [Google Scholar] [CrossRef]
  28. Boll, R.A.; Malkemus, D.; Mirzadeh, S. Production of actinium-225 for alpha particle mediated radioimmunotherapy. Appl. Radiat. Isot. 2005, 62, 667–679. [Google Scholar] [CrossRef]
  29. Morgenstern, A.; Apostolidis, C.; Bruchertseifer, F. Supply and Clinical Application of Actinium-225 and Bismuth-213. Semin. Nucl. Med. 2020, 50, 119–123. [Google Scholar] [CrossRef] [PubMed]
  30. Ahenkorah, S.; Cassells, I.; Deroose, C.M.; Cardinaels, T.; Burgoyne, A.R.; Bormans, G.; Ooms, M.; Cleeren, F. Bismuth-213 for Targeted Radionuclide Therapy: From Atom to Bedside. Pharmaceutics 2021, 13, 599. [Google Scholar] [CrossRef]
  31. Aliev, R.A.; Ermolaev, S.V.; Vasiliev, A.N.; Ostapenko, V.S.; Lapshina, E.V.; Zhuikov, B.L.; Zakharov, N.V.; Pozdeev, V.V.; Kokhanyuk, V.M.; Myasoedov, B.F.; et al. Isolation of Medicine-Applicable Actinium-225 from Thorium Targets Irradiated by Medium-Energy Protons. Solvent Extr. Ion Exch. 2014, 32, 468–477. [Google Scholar] [CrossRef]
  32. Radchenko, V.; Engle, J.W.; Wilson, J.J.; Maassen, J.R.; Nortier, F.M.; Taylor, W.A.; Birnbaum, E.R.; Hudston, L.A.; John, K.D.; Fassbender, M.E. Application of ion exchange and extraction chromatography to the separation of actinium from proton-irradiated thorium metal for analytical purposes. J. Chromatogr. A 2015, 1380, 55–63. [Google Scholar] [CrossRef]
  33. Zhuikov, B.L. Successes and problems in the development of medical radioisotope production in Russia. Phys. Usp. 2016, 59, 481–486. [Google Scholar] [CrossRef]
  34. Apostolidis, C.; Molinet, R.; Rasmussen, G.; Morgenstern, A. Production of Ac-225 from Th-229 for targeted α therapy. Anal. Chem. 2005, 77, 6288–6291. [Google Scholar] [CrossRef]
  35. Harvey, J.; Nolen, J.A.; Kroc, T.; Gomes, I.; Horwitz, E.P.; Mcalister, D.R. Production of Actinium-225 Via High Energy Proton Induced Spallation of Thorium-232. In Proceedings of the Workshop on Applications of High Intensity Proton Accelerators, Batavia, IL, USA, 19–21 October 2009; pp. 321–326. [Google Scholar]
  36. Bidkar, A.P.; Zerefa, L.; Yadav, S.; VanBrocklin, H.F.; Flavell, R.R. Actinium-225 targeted alpha particle therapy for prostate cancer. Theranostics 2024, 14, 2969–2992. [Google Scholar] [CrossRef]
  37. Radchenko, V.; Engle, J.W.; Thisgaard, H. Status and future perspectives of Meitner-Auger and low energy electron-emitting radionuclides for targeted radionuclide therapy. Nucl. Med. Biol. 2021, 94–95, 106. [Google Scholar] [CrossRef] [PubMed]
  38. Eychenne, R.; Chérel, M.; Haddad, F.; Guérard, F.; Gestin, J.F. Overview of the Most Promising Radionuclides for Targeted Alpha Therapy: The “Hopeful Eight”. Pharmaceutics 2021, 13, 906. [Google Scholar] [CrossRef] [PubMed]
  39. Melville, G.; Allen, B.J. Cyclotron and linac production of Ac-225. Appl. Radiat. Isot. 2009, 67, 549–555. [Google Scholar] [CrossRef] [PubMed]
  40. Koch, L.; Apostolidis, C.; Janssens, W.; Molinet, R.; Van Geel, J. Production of Ac-225 and application of the Bi-213 daughter in cancer therapy. Czechoslov. J. Phys. 1999, 49, 817–822. [Google Scholar] [CrossRef]
  41. Engle, J.W.; Mashnik, S.G.; Weidner, J.W.; Wolfsberg, L.E.; Fassbender, M.E.; Jackman, K.; Couture, A.; Bitteker, L.J.; Ullmann, J.L.; Gulley, M.S.; et al. Cross sections from proton irradiation of thorium at 800 MeV. Phys. Rev. C 2013, 88, 014604. [Google Scholar] [CrossRef]
  42. Weidner, J.W.; Mashnik, S.G.; John, K.D.; Ballard, B.; Birnbaum, E.R.; Bitteker, L.J.; Couture, A.; Fassbender, M.E.; Goff, G.S.; Gritzo, R.; et al. 225Ac and 223Ra production via 800 MeV proton irradiation of natural thorium targets. Appl. Radiat. Isot. 2012, 70, 2590–2595. [Google Scholar] [CrossRef]
  43. Melville, G.; Meriarty, H.; Metcalfe, P.; Knittel, T.; Allen, B.J. Production of Ac-225 for cancer therapy by photon-induced transmutation of Ra-226. Appl. Radiat. Isot. 2007, 65, 1014–1022. [Google Scholar] [CrossRef] [PubMed]
  44. Ma, D.; McDevitt, M.R.; Finn, R.D.; Scheinberg, D.A. Breakthrough of 225Ac and its radionuclide daughters from an 225Ac/213Bi generator: Development of new methods, quantitative characterization, and implications for clinical use. Appl. Radiat. Isot. 2001, 55, 667–678. [Google Scholar] [CrossRef]
  45. Robertson AK, R.C.; Schaffer, P.; Radchenko, V. Development of 225Ac radiopharmaceuticals: TRIUMF perspectives and experiences. Curr. Radiopharm. 2018, 11, 156–172. [Google Scholar] [CrossRef]
  46. Apostolidis, C.; Molinet, R.; McGinley, J.; Abbas, K.; Möllenbeck, J.; Morgenstern, A. Cyclotron production of Ac-225 for targeted alpha therapy. Appl. Radiat. Isot. 2005, 62, 383–387. [Google Scholar] [CrossRef]
  47. Schaffer, P.; Bénard, F.; Bernstein, A.; Buckley, K.; Celler, A.; Cockburn, N.; Corsaut, J.; Dodd, M.; Economou, C.; Eriksson, T.; et al. Direct Production Of 99mTc via 100Mo(p,2n) on Small Medical Cyclotrons. Phys. Procedia 2015, 66, 383–395. [Google Scholar] [CrossRef]
  48. Böhlen, T.T.; Cerutti, F.; Chin, M.P.W.; Fossò, A.; Ferrari, A.; Ortega, P.G.; Mairani, A.; Sala, P.R.; Smirnov, G.; Vlachoudisl, V. The FLUKA Code: Developments and Challenges for High Energy and Medical Applications. Nucl. Data Sheets 2014, 120, 211–214. [Google Scholar] [CrossRef]
  49. Jalilian, A.R.; Kleynhans, J.; Bouziotis, P.; Bruchertseifer, F.; Chakraborty, S.; De Blois, E.; Denecke, M.; Elboga, U.; Gizawyi, M.; Horak, C.; et al. IAEA activities to support the member states in the production of targeted alpha therapy radiopharmaceuticals. Nucl. Med. Biol. 2025, 144, 109008. [Google Scholar] [CrossRef] [PubMed]
  50. Chadwick, M.B.; Herman, M.; Oblozinsky, P.; Dunn, M.E.; Danon, Y.; Kahler, A.C.; Smith, D.L.; Pritychenko, B.; Arbanas, G.; Arcilla, R.; et al. ENDF/B-VII.1 Nuclear Data for Science and Technology: Cross Sections, Covariances, Fission Product Yields and Decay Data. Nucl. Data Sheets 2011, 112, 2887–2996. [Google Scholar] [CrossRef]
  51. IAEA. IAEA-TECDOC-886; IAEA: Vienna, Austria, 1996. [Google Scholar]
  52. Tosato, M.; Favaretto, C.; Kleynhans, J.; Burgoyne, A.R.; Gestin, J.F.; van der Meulen, N.P.; Jalilian, A.; Koster, U.; Asti, M.; Radchenko, V. Alpha Atlas: Mapping global production of alpha-emitting radionuclides for targeted alpha therapy. Nucl. Med. Biol. 2025, 142–143, 108990. [Google Scholar] [CrossRef]
  53. Ermolaev, S.V.; Zhuikov, B.L.; Kokhanyuk, V.M.; Matushko, V.L.; Kalmykov, S.N.; Aliev, R.A.; Tananaev, I.G.; Myasoedov, B.F. Production of actinium, thorium and radium isotopes from natural thorium irradiated with protons up to 141 MeV. Radiochim. Acta 2012, 100, 223–229. [Google Scholar] [CrossRef]
  54. Griswold, J.R.; Medvedev, D.G.; Engle, J.W.; Copping, R.; Fitzsimmons, J.M.; Radchenko, V.; Cooley, J.C.; Fassbender, M.E.; Denton, D.L.; Murphy, K.E.; et al. Large scale accelerator production of 225AC: Effective cross sections for 78-192 MeV protons incident on 232Th targets. Appl. Radiat. Isot. 2016, 118, 366–374. [Google Scholar] [CrossRef] [PubMed]
  55. Weidner, J.W.; Mashnik, S.G.; John, K.D.; Hemez, F.; Ballard, B.; Bach, H.; Birnbaum, E.R.; Bitteker, L.J.; Couture, A.; Dry, D.; et al. Proton-induced cross sections relevant to production of Ac and Ra in natural thorium targets below 200 MeV. Appl. Radiat. Isot. 2012, 70, 2602–2607. [Google Scholar] [CrossRef] [PubMed]
  56. Englert, M.; Krall, L.; Ewing, R.C. Is nuclear fission a sustainable source of energy? MRS Bull. 2012, 37, 417–424. [Google Scholar] [CrossRef]
  57. Robertson, A.K.H.; Ladouceur, K.; Nozar, M.; Moskven, L.; Ramogida, C.F.; D’Auria, J.; Sossi, V.; Schaffer, P. Design and Simulation of a Thorium Target for 225AC Production. AIP Conf. Proc. 2017, 1845, 020019. [Google Scholar] [CrossRef]
  58. Laxdal, R.E.; Morton, A.C.; Schaffer, P. Radioactive Ion Beams and Radiopharmaceuticals. Rev. Accel. Sci. Technol. 2013, 6, 37–57. [Google Scholar] [CrossRef]
  59. Ziegler, J.F.; Ziegler, M.D.; Biersack, J.P. SRIM—The stopping and range of ions in matter. Nucl. Instrum. Meth B 2010, 268, 1818–1823. [Google Scholar] [CrossRef]
  60. Copping, R.; Denton, D.; Murphy, K.; Hickman, E.; Marcus, C.; Stracener, D.; Mirzadeh, S. Radium targets for the reactor production of alpha-emitting medical radioisotopes. Abstr. Pap. Am. Chem. S 2018, 256, S1–S42. [Google Scholar]
  61. Boll, R.A.; Garland, M.; Mirzadeh, S. Reactor production of thorium-229. AIP Conf. Proc. 2005, 769, 1674–1675. [Google Scholar] [CrossRef]
  62. Baimukhanova, A.; Engudar, G.; Marinov, G.; Kurakina, E.; Dadakhanov, J.; Karaivanov, D.; Yang, H.; Ramogida, C.F.; Schaffer, P.; Magomedbekov, E.P.; et al. An alternative radiochemical separation strategy for isolation of Ac and Ra isotopes from high energy proton irradiated thorium targets for further application in Targeted Alpha Therapy (TAT). Nucl. Med. Biol. 2022, 112-113, 35–43. [Google Scholar] [CrossRef]
  63. Franchi, S.; Di Marco, V.; Tosato, M. Bismuth chelation for targeted alpha therapy: Current state of the art. Nucl. Med. Biol. 2022, 114, 168–188. [Google Scholar] [CrossRef]
  64. Forrester, R.; Dutech, G.; Akin, A.; Fassbender, M.E.; Mastren, T. An electrochemical generator for the continual supply of Bi from Ac for use in targeted alpha therapy applications. Nucl. Med. Biol. 2024, 136, 108941. [Google Scholar] [CrossRef]
  65. McAlister, D.R.; Horwitz, E.P. Automated two column generator systems for medical radionuclides. Appl. Radiat. Isot. 2009, 67, 1985–1991. [Google Scholar] [CrossRef] [PubMed]
  66. Bray, L.A.; Tingey, J.M.; DesChane, J.R.; Egorov, O.B.; Tenforde, T.S.; Wilbur, D.S.; Hamlin, D.K.; Pathare, P.M. Development of a unique bismuth (Bi-213) automated generator for use in cancer therapy. Ind. Eng. Chem. Res. 2000, 39, 3189–3194. [Google Scholar] [CrossRef]
  67. Wu, C.; Brechbiel, M.W.; Gansow, O.A. An improved generator for the production of Bi-213 from Ac-225. Abstr. Pap. Am. Chem. S 1996, 212, 61. [Google Scholar]
  68. McDevitt, M.R.; Finn, R.D.; Sgouros, G.; Ma, D.S.; Scheinberg, D.A. An Ac/Bi generator system for therapeutic clinical applications:: Construction and operation. Appl. Radiat. Isot. 1999, 50, 895–904. [Google Scholar] [CrossRef]
  69. Mirzadeh, S. Generator-produced alpha-emitters. Appl. Radiat. Isot. 1998, 49, 345–349. [Google Scholar] [CrossRef]
  70. Vasiliev, A.N.; Ermolaev, S.V.; Lapshina, E.V.; Zhuikov, B.L.; Betenekov, N.D. Ac/Bi generator based on inorganic sorbents. Radiochim. Acta 2019, 107, 1203–1211. [Google Scholar] [CrossRef]
  71. Zielinska, B.; Apostolidis, C.; Bruchertseifer, F.; Morgenstern, A. An improved method for the production of Ac-225/Bi-213 from Th-229 for targeted alpha therapy. Solvent Extr. Ion Exch. 2007, 25, 339–349. [Google Scholar] [CrossRef]
  72. Betenekov, N.D.; Denisov, E.I.; Vasiliev, A.N.; Ermolaev, S.V.; Zhuikov, B.L. Prospects for the Development of an 225Ac/213Bi Generator Using Inorganic Hydroxide Sorbents. Radiochemistry 2019, 61, 211–219. [Google Scholar] [CrossRef]
  73. Bruchertseifer, F.; Comba, P.; Martin, B.; Morgenstern, A.; Notni, J.; Starke, M.; Wadepohl, H. First-Generation Bispidine Chelators for Bi Radiopharmaceutical Applications. ChemMedChem 2020, 15, 1591–1600. [Google Scholar] [CrossRef]
  74. Hassan, M.; Bokhari, T.H.; Lodhi, N.A.; Khosa, M.K.; Usman, M. A review of recent advancements in Actinium-225 labeled compounds and biomolecules for therapeutic purposes. Chem. Biol. Drug Des. 2023, 102, 1276–1292. [Google Scholar] [CrossRef] [PubMed]
  75. Mdanda, S.; Ngema, L.M.; Mdlophane, A.; Sathekge, M.M.; Zeevaart, J.R. Recent Innovations and Nano-Delivery of Actinium-225: A Narrative Review. Pharmaceutics 2023, 15, 1719. [Google Scholar] [CrossRef]
  76. Jaggi, J.S.; Kappel, B.J.; McDevitt, M.R.; Sgouros, G.; Flombaum, C.D.; Cabassa, C.; Scheinberg, D.A. Efforts to control the errant products of a targeted in vivo generator. Cancer Res. 2005, 65, 4888–4895. [Google Scholar] [CrossRef]
  77. de Kruijff, R.M.; Wolterbeek, H.T.; Denkova, A.G. A Critical Review of Alpha Radionuclide Therapy-How to Deal with Recoiling Daughters? Pharmaceuticals 2015, 8, 321–336. [Google Scholar] [CrossRef]
  78. Miederer, M.; Henriksen, G.; Alke, A.; Mossbrugger, I.; Quintanilla-Martinez, L.; Senekowitsch-Schmidtke, R.; Essler, M. Preclinical evaluation of the α-particle generator nuclide Ac for somatostatin receptor radiotherapy of neuroendocrine tumors. Clin. Cancer Res. 2008, 14, 3555–3561. [Google Scholar] [CrossRef]
  79. Essler, M.; Gartner, F.C.; Neff, F.; Blechert, B.; Senekowitsch-Schmidtke, R.; Bruchertseifer, F.; Morgenstern, A.; Seidl, C. Therapeutic efficacy and toxicity of 225Ac-labelled vs. 213Bi-labelled tumour-homing peptides in a preclinical mouse model of peritoneal carcinomatosis. Eur. J. Nucl. Med. Mol. Imaging 2012, 39, 602–612. [Google Scholar] [CrossRef]
  80. McDevitt, M.R.; Ma, D.; Lai, L.T.; Simon, J.; Borchardt, P.; Frank, R.K.; Wu, K.; Pellegrini, V.; Curcio, M.J.; Miederer, M.; et al. Tumor therapy with targeted atomic nanogenerators. Science 2001, 294, 1537–1540. [Google Scholar] [CrossRef] [PubMed]
  81. Sofou, S.; Kappel, B.J.; Jaggi, J.S.; McDevitt, M.R.; Scheinberg, D.A.; Sgouros, G. Enhanced retention of the α-particle-emitting daughters of actinium-225 by liposome carriers. Bioconjugate Chem. 2007, 18, 2061–2067. [Google Scholar] [CrossRef] [PubMed]
  82. Bandekar, A.; Zhu, C.; Jindal, R.; Bruchertseifer, F.; Morgenstern, A.; Sofou, S. Anti-prostate-specific membrane antigen liposomes loaded with 225Ac for potential targeted antivascular alpha-particle therapy of cancer. J. Nucl. Med. 2014, 55, 107–114. [Google Scholar] [CrossRef]
  83. Fitzsimmons, J.; Atcher, R.; Cutler, C. Development of a prelabeling approach for a targeted nanochelator. J. Radioanal. Nucl. Chem. 2015, 305, 161–167. [Google Scholar] [CrossRef]
  84. Matson, M.L.; Villa, C.H.; Ananta, J.S.; Law, J.J.; Scheinberg, D.A.; Wilson, L.J. Encapsulation of α-Particle-Emitting 225Ac3+ Ions Within Carbon Nanotubes. J. Nucl. Med. 2015, 56, 897–900. [Google Scholar] [CrossRef]
  85. McLaughlin, M.F.; Woodward, J.; Boll, R.A.; Wall, J.S.; Rondinone, A.J.; Kennel, S.J.; Mirzadeh, S.; Robertson, J.D. Gold coated lanthanide phosphate nanoparticles for targeted alpha generator radiotherapy. PLoS ONE 2013, 8, e54531. [Google Scholar] [CrossRef] [PubMed]
  86. Wang, G.; de Kruijff, R.M.; Rol, A.; Thijssen, L.; Mendes, E.; Morgenstern, A.; Bruchertseifer, F.; Stuart, M.C.; Wolterbeek, H.T.; Denkova, A.G. Retention studies of recoiling daughter nuclides of 225Ac in polymer vesicles. Appl. Radiat. Isot. 2014, 85, 45–53. [Google Scholar] [CrossRef]
  87. Zhu, C.; Bandekar, A.; Sempkowski, M.; Banerjee, S.R.; Pomper, M.G.; Bruchertseifer, F.; Morgenstern, A.; Sofou, S. Nanoconjugation of PSMA-Targeting Ligands Enhances Perinuclear Localization and Improves Efficacy of Delivered Alpha-Particle Emitters against Tumor Endothelial Analogues. Mol. Cancer Ther. 2016, 15, 106–113. [Google Scholar] [CrossRef]
  88. Sofou, S.; Thomas, J.L.; Lin, H.Y.; McDevitt, M.R.; Scheinberg, D.A.; Sgouros, G. Engineered liposomes for potential alpha-particle therapy of metastatic cancer. J. Nucl. Med. 2004, 45, 253–260. [Google Scholar]
  89. Arazi, L.; Cooks, T.; Schmidt, M.; Keisari, Y.; Kelson, I. Treatment of solid tumors by interstitial release of recoiling short-lived alpha emitters. Phys. Med. Biol. 2007, 52, 5025–5042. [Google Scholar] [CrossRef] [PubMed]
  90. Confino, H.; Hochman, I.; Efrati, M.; Schmidt, M.; Umansky, V.; Kelson, I.; Keisari, Y. Tumor ablation by intratumoral Ra-224-loaded wires induces anti-tumor immunity against experimental metastatic tumors. Cancer Immunol. Immunother. 2015, 64, 191–199. [Google Scholar] [CrossRef]
  91. Cooks, T.; Schmidt, M.; Bittan, H.; Lazarov, E.; Arazi, L.; Kelson, I.; Keisari, Y. Local control of lung derived tumors by diffusing alpha-emitting atoms released from intratumoral wires loaded with radium-224. Int. J. Radiat. Oncol. Biol. Phys. 2009, 74, 966–973. [Google Scholar] [CrossRef] [PubMed]
  92. Cooks, T.; Tal, M.; Raab, S.; Efrati, M.; Reitkopf, S.; Lazarov, E.; Etzyoni, R.; Schmidt, M.; Arazi, L.; Kelson, I.; et al. Intratumoral 224Ra-loaded wires spread alpha-emitters inside solid human tumors in athymic mice achieving tumor control. Anticancer Res. 2012, 32, 5315–5321. [Google Scholar]
  93. Cordier, D.; Forrer, F.; Bruchertseifer, F.; Morgenstern, A.; Apostolidis, C.; Good, S.; Muller-Brand, J.; Macke, H.; Reubi, J.C.; Merlo, A. Targeted alpha-radionuclide therapy of functionally critically located gliomas with 213Bi-DOTA-[Thi8,Met(O2)11]-substance P: A pilot trial. Eur. J. Nucl. Med. Mol. Imaging 2010, 37, 1335–1344. [Google Scholar] [CrossRef] [PubMed]
  94. Horev-Drori, G.; Cooks, T.; Bittan, H.; Lazarov, E.; Schmidt, M.; Arazi, L.; Efrati, M.; Kelson, I.; Keisari, Y. Local control of experimental malignant pancreatic tumors by treatment with a combination of chemotherapy and intratumoral 224radium-loaded wires releasing alpha-emitting atoms. Transl. Res. 2012, 159, 32–41. [Google Scholar] [CrossRef] [PubMed]
  95. Dhiman, D.; Vatsa, R.; Sood, A. Challenges and opportunities in developing Actinium-225 radiopharmaceuticals. Nucl. Med. Commun. 2022, 43, 970–977. [Google Scholar] [CrossRef]
  96. Escorcia, F.E.; Henke, E.; McDevitt, M.R.; Villa, C.H.; Smith-Jones, P.; Blasberg, R.G.; Benezra, R.; Scheinberg, D.A. Selective Killing of Tumor Neovasculature Paradoxically Improves Chemotherapy Delivery to Tumors. Cancer Res. 2010, 70, 9277–9286. [Google Scholar] [CrossRef]
  97. Singh Jaggi, J.; Henke, E.; Seshan, S.V.; Kappel, B.J.; Chattopadhyay, D.; May, C.; McDevitt, M.R.; Nolan, D.; Mittal, V.; Benezra, R.; et al. Selective alpha-particle mediated depletion of tumor vasculature with vascular normalization. PLoS ONE 2007, 2, e267. [Google Scholar] [CrossRef]
  98. Maguire, W.F.; McDevitt, M.R.; Smith-Jones, P.M.; Scheinberg, D.A. Efficient 1-step radiolabeling of monoclonal antibodies to high specific activity with 225Ac for alpha-particle radioimmunotherapy of cancer. J. Nucl. Med. 2014, 55, 1492–1498. [Google Scholar] [CrossRef]
  99. Majkowska-Pilip, A.; Rius, M.; Bruchertseifer, F.; Apostolidis, C.; Weis, M.; Bonelli, M.; Laurenza, M.; Krolicki, L.; Morgenstern, A. In vitro evaluation of (225) Ac-DOTA-substance P for targeted alpha therapy of glioblastoma multiforme. Chem. Biol. Drug Des. 2018, 92, 1344–1356. [Google Scholar] [CrossRef]
  100. Miederer, M.; McDevitt, M.R.; Sgouros, G.; Kramer, K.; Cheung, N.K.V.; Scheinberg, D.A. Pharmacokinetics, dosimetry, and toxicity of the targetable atomic generator, Ac-HuM195, in nonhuman primates. J. Nucl. Med. 2004, 45, 129–137. [Google Scholar]
  101. Pruszynski, M.; D’Huyvetter, M.; Bruchertseifer, F.; Morgenstern, A.; Lahoutte, T. Evaluation of an Anti-HER2 Nanobody Labeled with (225)Ac for Targeted alpha-Particle Therapy of Cancer. Mol. Pharm. 2018, 15, 1457–1466. [Google Scholar] [CrossRef]
  102. Borchardt, P.E.; Yuan, R.R.; Miederer, M.; McDevitt, M.R.; Scheinberg, D.A. Targeted actinium-225 generators for therapy of ovarian cancer. Cancer Res. 2003, 63, 5084–5090. [Google Scholar]
  103. Deal, K.A.; Davis, I.A.; Mirzadeh, S.; Kennel, S.J.; Brechbiel, M.W. Improved in vivo stability of actinium-225 macrocyclic complexes. J. Med. Chem. 1999, 42, 2988–2992. [Google Scholar] [CrossRef]
  104. Chappell, L.L.; Deal, K.A.; Dadachova, E.; Brechbiel, M.W. Synthesis, conjugation, and radiolabeling of a novel bifunctional chelating agent for (225)Ac radioimmunotherapy applications. Bioconjug Chem. 2000, 11, 510–519. [Google Scholar] [CrossRef]
  105. Davis, I.A.; Glowienka, K.A.; Boll, R.A.; Deal, K.A.; Brechbiel, M.W.; Stabin, M.; Bochsler, P.N.; Mirzadeh, S.; Kennel, S.J. Comparison of 225actinium chelates: Tissue distribution and radiotoxicity. Nucl. Med. Biol. 1999, 26, 581–589. [Google Scholar] [CrossRef]
  106. Rosenblat, T.L.; McDevitt, M.R.; Mulford, D.A.; Pandit-Taskar, N.; Divgi, C.R.; Panageas, K.S.; Heaney, M.L.; Chanel, S.; Morgenstern, A.; Sgouros, G.; et al. Sequential cytarabine and alpha-particle immunotherapy with bismuth-213-lintuzumab (HuM195) for acute myeloid leukemia. Clin. Cancer Res. 2010, 16, 5303–5311. [Google Scholar] [CrossRef]
  107. Cao, X.; Li, Q.; Moritz, A.; Xie, Z.; Dolg, M.; Chen, X.; Fang, W. Density functional theory studies of actinide(III) motexafins (An-Motex2+, An = Ac, Cm, Lr). Structure, stability, and comparison with lanthanide(III) motexafins. Inorg. Chem. 2006, 45, 3444–3451. [Google Scholar] [CrossRef]
  108. Comba, P.; Jermilova, U.; Orvig, C.; Patrick, B.O.; Ramogida, C.F.; Ruck, K.; Schneider, C.; Starke, M. Octadentate Picolinic Acid-Based Bispidine Ligand for Radiometal Ions. Chemistry 2017, 23, 15945–15956. [Google Scholar] [CrossRef]
  109. Autenrieth, M.E.; Seidl, C.; Bruchertseifer, F.; Horn, T.; Kurtz, F.; Feuerecker, B.; D’Alessandria, C.; Pfob, C.; Nekolla, S.; Apostolidis, C.; et al. Treatment of carcinoma in situ of the urinary bladder with an alpha-emitter immunoconjugate targeting the epidermal growth factor receptor: A pilot study. Eur. J. Nucl. Med. Mol. Imaging 2018, 45, 1364–1371. [Google Scholar] [CrossRef]
  110. Wu, C.; Gansow, O.A.; Brechbiel, M.W. Evaluation of methods for large scale preparation of antibody ligand conjugates. Nucl. Med. Biol. 1999, 26, 339–342. [Google Scholar] [CrossRef]
  111. Xu, H.; Baidoo, K.E.; Wong, K.J.; Brechbiel, M.W. A novel bifunctional maleimido CHX-A” chelator for conjugation to thiol-containing biomolecules. Bioorganic Med. Chem. Lett. 2008, 18, 2679–2683. [Google Scholar] [CrossRef][Green Version]
  112. Schwartz, J.; Jaggi, J.S.; O’Donoghue, J.A.; Ruan, S.; McDevitt, M.; Larson, S.M.; Scheinberg, D.A.; Humm, J.L. Renal uptake of bismuth-213 and its contribution to kidney radiation dose following administration of actinium-225-labeled antibody. Phys. Med. Biol. 2011, 56, 721–733. [Google Scholar] [CrossRef]
  113. Jurcic, J.G.; Larson, S.M.; Sgouros, G.; McDevitt, M.R.; Finn, R.D.; Divgi, C.R.; Ballangrud, A.M.; Hamacher, K.A.; Ma, D.; Humm, J.L.; et al. Targeted alpha particle immunotherapy for myeloid leukemia. Blood 2002, 100, 1233–1239. [Google Scholar] [CrossRef]
  114. Park, S.I.; Shenoi, J.; Pagel, J.M.; Hamlin, D.K.; Wilbur, D.S.; Orgun, N.; Kenoyer, A.L.; Frayo, S.; Axtman, A.; Back, T.; et al. Conventional and pretargeted radioimmunotherapy using bismuth-213 to target and treat non-Hodgkin lymphomas expressing CD20: A preclinical model toward optimal consolidation therapy to eradicate minimal residual disease. Blood 2010, 116, 4231–4239. [Google Scholar] [CrossRef] [PubMed]
  115. Roscher, M.; Hormann, I.; Leib, O.; Marx, S.; Moreno, J.; Miltner, E.; Friesen, C. Targeted alpha-therapy using [Bi-213]anti-CD20 as novel treatment option for radio- and chemoresistant non-Hodgkin lymphoma cells. Oncotarget 2013, 4, 218–230. [Google Scholar] [CrossRef]
  116. Song, E.Y.; Rizvi, S.M.A.; Qu, C.F.; Raja, C.; Brechbiel, M.W.; Morgenstern, A.; Apostolidis, C.; Allen, B.J. Pharmacokinetics and toxicity ofBi-labeled PAI2 in preclinical targeted alpha therapy for cancer. Cancer Biol. Ther. 2007, 6, 898–904. [Google Scholar] [CrossRef] [PubMed]
  117. Qu, C.F.; Song, Y.J.; Rizvi, S.M.A.; Li, Y.; Smith, R.; Perkins, A.C.; Morgenstern, A.; Brechbiel, M.; Allen, B.J. In vivo and in vitro inhibition of pancreatic cancer growth by targeted alpha therapy using Bi-CHX. A”-C595. Cancer Biol. Ther. 2005, 4, 848–853. [Google Scholar] [CrossRef]
  118. Allen, B.J.; Tian, Z.; Rizvi, S.M.; Li, Y.; Ranson, M. Preclinical studies of targeted alpha therapy for breast cancer using 213Bi-labelled-plasminogen activator inhibitor type 2. Br. J. Cancer 2003, 88, 944–950. [Google Scholar] [CrossRef] [PubMed]
  119. Song, E.Y.; Qu, C.F.; Rizvi, S.M.A.; Raja, C.; Beretov, J.; Morgenstern, A.; Apostolidis, C.; Bruchertseifer, F.; Perkins, A.; Allen, B.J. Bismuth-213 radioimmunotherapy with C595 anti-MUC1 monoclonal antibody in an ovarian cancer ascites model. Cancer Biol. Ther. 2008, 7, 76–80. [Google Scholar] [CrossRef]
  120. Morgenstern, A.; Krolicki, L.; Kunikowska, J.; Koziara, H.; Krolicki, B.; Jakucinski, M.; Apostolidis, C.; Bruchertseifer, F. Targeted alpha therapy of glioblastoma multiforme: First clinical experience with 213Bi-substance P. J. Nucl. Med. 2014, 55, 390. [Google Scholar]
  121. Allen, B.J.; Raja, C.; Rizvi, S.; Li, Y.; Tsui, W.; Graham, P.; Thompson, J.F.; Reisfeld, R.A.; Kearsley, J. Intralesional targeted alpha therapy for metastatic melanoma. Cancer Biol. Ther. 2005, 4, 1318–1324. [Google Scholar] [CrossRef]
  122. Allen, B.J.; Singla, A.A.; Rizvi, S.M.; Graham, P.; Bruchertseifer, F.; Apostolidis, C.; Morgenstern, A. Analysis of patient survival in a Phase I trial of systemic targeted alpha-therapy for metastatic melanoma. Immunotherapy 2011, 3, 1041–1050. [Google Scholar] [CrossRef]
  123. Raja, C.; Graham, P.; Abbas Rizvi, S.M.; Song, E.; Goldsmith, H.; Thompson, J.; Bosserhoff, A.; Morgenstern, A.; Apostolidis, C.; Kearsley, J.; et al. Interim analysis of toxicity and response in phase 1 trial of systemic targeted alpha therapy for metastatic melanoma. Cancer Biol. Ther. 2007, 6, 846–852. [Google Scholar] [CrossRef]
  124. Li, Y.; Wang, J.; Rizvi, S.M.; Jager, M.J.; Conway, R.M.; Billson, F.A.; Allen, B.J.; Madigan, M.C. In vitro targeting of NG2 antigen by 213Bi-9.2.27 alpha-immunoconjugate induces cytotoxicity in human uveal melanoma cells. Investig. Ophthalmol. Vis. Sci. 2005, 46, 4365–4371. [Google Scholar] [CrossRef]
  125. Rizvi, S.M.; Allen, B.J.; Tian, Z.; Goozee, G.; Sarkar, S. In vitro and preclinical studies of targeted alpha therapy (TAT) for colorectal cancer. Color. Dis. 2001, 3, 345–353. [Google Scholar] [CrossRef]
  126. Fichou, N.; Gouard, S.; Maurel, C.; Barbet, J.; Ferrer, L.; Morgenstern, A.; Bruchertseifer, F.; Faivre-Chauvet, A.; Bigot-Corbel, E.; Davodeau, F.; et al. Single-Dose Anti-CD138 Radioimmunotherapy: Bismuth-213 is More Efficient than Lutetium-177 for Treatment of Multiple Myeloma in a Preclinical Model. Front. Med. 2015, 2, 76. [Google Scholar] [CrossRef]
  127. Li, Y.; Tian, Z.; Rizvi, S.M.A.; Bander, N.H.; Allen, B.J. and preclinical targeted alpha therapy of human prostate cancer with Bi-213 labeled J591 antibody against the prostate specific membrane antigen. Prostate Cancer Prostatic Dis. 2002, 5, 36–46. [Google Scholar] [CrossRef]
  128. Adams, G.P.; Shaller, C.C.; Chappell, L.L.; Wu, C.; Horak, E.M.; Simmons, H.H.; Litwin, S.; Marks, J.D.; Weiner, L.M.; Brechbiel, M.W. Delivery of the α-emitting radioisotope bismuth-213 to solid tumors via single-chain Fv and diabody molecules. Nucl. Med. Biol. 2000, 27, 339–346. [Google Scholar] [CrossRef]
  129. Pfost, B.; Seidl, C.; Autenrieth, M.; Saur, D.; Bruchertseifer, F.; Morgenstern, A.; Schwaiger, M.; Senekowitsch-Schmidtke, R. Intravesical α-Radioimmunotherapy with Bi-Anti-EGFR-mAb Defeats Human Bladder Carcinoma in Xenografted Nude Mice. J. Nucl. Med. 2009, 50, 1700–1708. [Google Scholar] [CrossRef] [PubMed]
  130. Poty, S.; Francesconi, L.C.; McDevitt, M.R.; Morris, M.J.; Lewis, J.S. α-Emitters for Radiotherapy: From Basic Radiochemistry to Clinical Studies-Part 2. J. Nucl. Med. 2018, 59, 1020–1027. [Google Scholar] [CrossRef]
  131. Yeong, C.H.; Cheng, M.H.; Ng, K.H. Therapeutic radionuclides in nuclear medicine: Current and future prospects. J. Zhejiang Univ. Sci. B 2014, 15, 845–863. [Google Scholar] [CrossRef]
  132. Sollini, M.; Marzo, K.; Chiti, A.; Kirienko, M. The five “W”s and “How” of Targeted Alpha Therapy: Why? Who? What? Where? When? and How? Rend. Lincei-Sci. Fis. 2020, 31, 231–247. [Google Scholar] [CrossRef]
  133. Current, K.; Meyer, C.; Magyar, C.E.; Mona, C.E.; Almajano, J.; Slavik, R.; Stuparu, A.D.; Cheng, C.; Dawson, D.W.; Radu, C.G.; et al. Investigating PSMA-Targeted Radioligand Therapy Efficacy as a Function of Cellular PSMA Levels and Intratumoral PSMA Heterogeneity. Clin. Cancer Res. 2020, 26, 2946–2955. [Google Scholar] [CrossRef] [PubMed]
  134. Ludwig, D.L.; Bryan, R.A.; Dawicki, W.; Geoghegan, E.M.; Liang, Q.; Gokhale, M.; Reddy, V.; Garg, R.; Allen, K.J.H.; Berger, M.S.; et al. Preclinical Development of an Actinium-225-Labeled Antibody Radio-Conjugate Directed Against CD45 for Targeted Conditioning and Radioimmunotherapy. Biol. Blood Marrow Tr. 2020, 26, S160–S161. [Google Scholar] [CrossRef]
  135. Marcu, L.; Bezak, E.; Allen, B.J. Global comparison of targeted alpha vs targeted beta therapy for cancer: In vitro, in vivo and clinical trials. Crit. Rev. Oncol. Hemat 2018, 123, 7–20. [Google Scholar] [CrossRef]
  136. Behling, K.; Maguire, W.F.; Lopez Puebla, J.C.; Sprinkle, S.R.; Ruggiero, A.; O’Donoghue, J.; Gutin, P.H.; Scheinberg, D.A.; McDevitt, M.R. Vascular Targeted Radioimmunotherapy for the Treatment of Glioblastoma. J. Nucl. Med. 2016, 57, 1576–1582. [Google Scholar] [CrossRef]
  137. Makvandi, M.; Dupis, E.; Engle, J.W.; Nortier, F.M.; Fassbender, M.E.; Simon, S.; Birnbaum, E.R.; Atcher, R.W.; John, K.D.; Rixe, O.; et al. Alpha-Emitters and Targeted Alpha Therapy in Oncology: From Basic Science to Clinical Investigations. Target. Oncol. 2018, 13, 189–203. [Google Scholar] [CrossRef] [PubMed]
  138. Tafreshi, N.K.; Doligalski, M.L.; Tichacek, C.J.; Pandya, D.N.; Budzevich, M.M.; El-Haddad, G.; Khushalani, N.I.; Moros, E.G.; McLaughlin, M.L.; Wadas, T.J.; et al. Development of Targeted Alpha Particle Therapy for Solid Tumors. Molecules 2019, 24, 4314. [Google Scholar] [CrossRef] [PubMed]
  139. McDevitt, M.R.; Sgouros, G.; Sofou, S. Targeted and Nontargeted alpha-Particle Therapies. Annu. Rev. Biomed. Eng. 2018, 20, 73–93. [Google Scholar] [CrossRef] [PubMed]
  140. Miederer, M.; McDevitt, M.R.; Borchardt, P.; Bergman, I.; Kramer, K.; Cheung, N.K.V.; Scheinberg, D.A. Treatment of neuroblastoma meningeal carcinomatosis with intrathecal application of α-emitting atomic nanogenerators targeting disialo-ganglioside GD2. Clin. Cancer Res. 2004, 10, 6985–6992. [Google Scholar] [CrossRef]
  141. Song, H.; Hobbs, R.F.; Vajravelu, R.; Huso, D.L.; Esaias, C.; Apostolidis, C.; Morgenstern, A.; Sgouros, G. Radioimmunotherapy of Breast Cancer Metastases with α-Particle Emitter Ac: Comparing Efficacy withBi and Y. Cancer Res. 2009, 69, 8941–8948. [Google Scholar] [CrossRef] [PubMed]
  142. Kennel, S.J.; Chappell, L.L.; Dadachova, K.; Brechbiel, M.W.; Lankford, T.K.; Davis, I.A.; Stabin, M.; Mirzadeh, S. Evaluation of Ac for vascular targeted radioimmunotherapy of lung tumors. Cancer Biother. Radiopharm. 2000, 15, 235–244. [Google Scholar] [CrossRef]
  143. Sattiraju, A.; Sai, K.K.S.; Xuan, A.; Pandya, D.N.; Almaguel, F.G.; Wadas, T.J.; Herpai, D.M.; Debinski, W.; Mintz, A. IL13RA2 targeted alpha particle therapy against glioblastomas. Oncotarget 2017, 8, 42997–43007. [Google Scholar] [CrossRef] [PubMed]
  144. Tafreshi, N.K.; Tichacek, C.J.; Pandya, D.N.; Doligalski, M.L.; Budzevich, M.M.; Kil, H.; Bhatt, N.B.; Kock, N.D.; Messina, J.L.; Ruiz, E.E.; et al. Melanocortin 1 Receptor-Targeted α-Particle Therapy for Metastatic Uveal Melanoma. J. Nucl. Med. 2019, 60, 1124–1133. [Google Scholar] [CrossRef] [PubMed]
  145. Pfannkuchen, N.; Pektor, S.; Miederer, M.; Roesch, F. In vivo evaluation of [Ac]Ac-DOTA for α-therapy of bone metastases. J. Nucl. Med. 2018, 11, 223–230. [Google Scholar]
  146. Nakamae, H.; Wilbur, D.S.; Hamlin, D.K.; Thakar, M.S.; Santos, E.B.; Fisher, D.R.; Kenoyer, A.L.; Pagel, J.M.; Press, O.W.; Storb, R.; et al. Biodistributions, Myelosuppression, and Toxicities in Mice Treated with an Anti-CD45 Antibody Labeled with the α-Emitting Radionuclides Bismuth-213 or Astatine-211. Cancer Res. 2009, 69, 2408–2415. [Google Scholar] [CrossRef] [PubMed]
  147. Chérel, M.; Gouard, S.; Gaschet, J.; Saï-Maurel, C.; Bruchertseifer, F.; Morgenstern, A.; Bourgeois, M.; Gestin, J.F.; Bodéré, F.K.; Barbet, J.; et al. 213Bi Radioimmunotherapy with an Anti-mCD138 Monoclonal Antibody in a Murine Model of Multiple Myeloma. J. Nucl. Med. 2013, 54, 1597–1604. [Google Scholar] [CrossRef]
  148. Seidl, C.; Zöckler, C.; Beck, R.; Quintanilla-Martinez, L.; Bruchertseifer, F.; Senekowitsch-Schmidtke, R. Lu-immunotherapy of experimental peritoneal carcinomatosis shows comparable effectiveness to Bi-immunotherapy, but causes toxicity not observed with Bi. Eur. J. Nucl. Med. Mol. I 2011, 38, 312–322. [Google Scholar] [CrossRef]
  149. Behr, T.M.; Behe, M.; Stabin, M.G.; Wehrmann, E.; Apostolidis, C.; Molinet, R.; Strutz, F.; Fayyazi, A.; Wieland, E.; Gratz, S.; et al. High-linear energy transfer (LET) alpha versus low-LET beta emitters in radioimmunotherapy of solid tumors: Therapeutic efficacy and dose-limiting toxicity of 213Bi- versus 90Y-labeled CO17-1A Fab’ fragments in a human colonic cancer model. Cancer Res. 1999, 59, 2635–2643. [Google Scholar]
  150. Nikula, T.K.; McDevitt, M.R.; Finn, R.D.; Wu, C.; Kozak, R.W.; Garmestani, K.; Brechbiel, M.W.; Curcio, M.J.; Pippin, C.G.; Tiffany-Jones, L.; et al. Alpha-emitting bismuth cyclohexylbenzyl DTPA constructs of recombinant humanized anti-CD33 antibodies: Pharmacokinetics, bioactivity, toxicity and chemistry. J. Nucl. Med. 1999, 40, 166–176. [Google Scholar]
  151. Bloechl, S.; Beck, R.; Seidl, C.; Morgenstern, A.; Schwaiger, M.; Senekowitsch-Schmidtke, R. Fractionated locoregional low-dose radioimmunotherapy improves survival in a mouse model of diffuse-type gastric cancer using a 213Bi-conjugated monoclonal antibody. Clin. Cancer Res. 2005, 11, 7070s–7074s. [Google Scholar] [CrossRef]
  152. Kennel, S.J.; Brechbiel, M.W.; Milenic, D.E.; Schlom, J.; Mirzadeh, S. Actinium-225 conjugates of MAb CC49 and humanized ΔCHCC49. Cancer Biother. Radiopharm. 2002, 17, 219–231. [Google Scholar] [CrossRef]
  153. Knor, S.; Sato, S.; Huber, T.; Morgenstern, A.; Bruchertseifer, F.; Schmitt, M.; Kessler, H.; Senekowitsch-Schmidtke, R.; Magdolen, V.; Seidl, C. Development and evaluation of peptidic ligands targeting tumour-associated urokinase plasminogen activator receptor (uPAR) for use in alpha-emitter therapy for disseminated ovarian cancer. Eur. J. Nucl. Med. Mol. Imaging 2008, 35, 53–64. [Google Scholar] [CrossRef] [PubMed]
  154. Norenberg, J.P.; Krenning, B.J.; Konings, I.R.H.M.; Kusewitt, D.F.; Nayak, T.K.; Anderson, T.L.; de Jong, M.; Garmestani, K.; Brechbiel, M.W.; Kvols, L.K. Bi-[DOTA,Tyr]octreotide peptide receptor radionuclide therapy of pancreatic tumors in a preclinical animal model. Clin. Cancer Res. 2006, 12, 897–903. [Google Scholar] [CrossRef] [PubMed]
  155. Wild, D.; Frischknecht, M.; Zhang, H.; Morgenstern, A.; Bruchertseifer, F.; Boisclair, J.; Provencher-Bolliger, A.; Reubi, J.C.; Maecke, H.R. Alpha- versus beta-particle radiopeptide therapy in a human prostate cancer model (213Bi-DOTA-PESIN and 213Bi-AMBA versus 177Lu-DOTA-PESIN). Cancer Res. 2011, 71, 1009–1018. [Google Scholar] [CrossRef] [PubMed]
  156. Kaspersen, F.M.; Bos, E.; Doornmalen, A.V.; Geerlings, M.W.; Apostolidis, C.; Molinet, R. Cytotoxicity of 213Bi- and 225Ac-immunoconjugates. Nucl. Med. Commun. 1995, 16, 468–476. [Google Scholar] [CrossRef]
  157. Ballangrud, Å.M.; Yang, W.H.; Palm, S.; Enmon, R.; Borchardt, P.E.; Pellegrini, V.A.; McDevitt, M.R.; Scheinberg, D.A.; Sgouros, G. Alpha-particle emitting atomic generator (actinium-225)-labeled trastuzumab (herceptin) targeting of breast cancer spheroids:: Efficacy HER2/expression. Clin. Cancer Res. 2004, 10, 4489–4497. [Google Scholar] [CrossRef]
  158. Yoshida, T.; Jin, K.; Song, H.; Park, S.; Huso, D.L.; Zhang, Z.; Han, L.F.; Zhu, C.; Bruchertseifer, F.; Morgenstern, A.; et al. Effective treatment of ductal carcinoma in situ with a HER-2-targeted alpha-particle emitting radionuclide in a preclinical model of human breast cancer. Oncotarget 2016, 7, 33306–33315. [Google Scholar] [CrossRef]
  159. Behling, K.; Maguire, W.F.; Di Gialleonardo, V.; Heeb, L.E.M.; Hassan, I.F.; Veach, D.R.; Keshari, K.R.; Gutin, P.H.; Scheinberg, D.A.; McDevitt, M.R. Remodeling the Vascular Microenvironment of Glioblastoma with α-Particles. J. Nucl. Med. 2016, 57, 1771–1777. [Google Scholar] [CrossRef]
  160. Nelson, B.J.B.; Andersson, J.D.; Wuest, F. Targeted Alpha Therapy: Progress in Radionuclide Production, Radiochemistry, and Applications. Pharmaceutics 2021, 13, 49. [Google Scholar] [CrossRef]
  161. Ludwig, D.; Bryan, R.; Dawicki, W.; Geoghegan, E.M.; Liang, Q.; Seth, S.; Gokhale, M.; Berger, M.S.; Reddy, V.; Garg, R.; et al. Preclinical Development of an Actinium-225-Labeled Antibody Radio-Conjugate Directed Against CD45 for Targeted Conditioning and Radioimmunotherapy. Blood 2019, 134, 5601. [Google Scholar] [CrossRef]
  162. McDevitt, M.R.; Finn, R.D.; Ma, D.; Larson, S.M.; Scheinberg, D.A. Preparation of alpha-emitting 213Bi-labeled antibody constructs for clinical use. J. Nucl. Med. 1999, 40, 1722–1727. [Google Scholar]
  163. McDevitt, M.R.; Barendswaard, E.; Ma, D.; Lai, L.; Curcio, M.J.; Sgouros, G.; Ballangrud, A.M.; Yang, W.H.; Finn, R.D.; Pellegrini, V.; et al. An α-particle emitting antibody ([213Bi]J591) for radioimmunotherapy of prostate cancer. Cancer Res. 2000, 60, 6095–6100. [Google Scholar] [PubMed]
  164. Sabet, A.; Ezziddin, K.; Pape, U.F.; Reichman, K.; Haslerud, T.; Ahmadzadehfar, H.; Biersack, H.J.; Nagarajah, J.; Ezziddin, S. Accurate assessment of long-term nephrotoxicity after peptide receptor radionuclide therapy withLu-octreotate. Eur. J. Nucl. Med. Mol. I 2014, 41, 505–510. [Google Scholar] [CrossRef]
  165. Chan, H.S.; Konijnenberg, M.W.; de Blois, E.; Koelewijn, S.; Baum, R.P.; Morgenstern, A.; Bruchertseifer, F.; Breeman, W.A.; de Jong, M. Influence of tumour size on the efficacy of targeted alpha therapy withBi-[DOTA,Tyr]-octreotate. Ejnmmi Res. 2016, 6, 6. [Google Scholar] [CrossRef]
  166. Chan, H.S.; Konijnenberg, M.W.; Daniels, T.; Nysus, M.; Makvandi, M.; de Blois, E.; Breeman, W.A.; Atcher, R.W.; de Jong, M.; Norenberg, J.P. Improved safety and efficacy of Bi-DOTATATE-targeted alpha therapy of somatostatin receptor-expressing neuroendocrine tumors in mice pre-treated with L-lysine. Ejnmmi Res. 2016, 6, 83. [Google Scholar] [CrossRef] [PubMed]
  167. Majkowska-Pilip, A.; Gaweda, W.; Zelechowska-Matysiak, K.; Wawrowicz, K.; Bilewicz, A. Nanoparticles in Targeted Alpha Therapy. Nanomaterials 2020, 10, 1366. [Google Scholar] [CrossRef] [PubMed]
  168. Chang, M.Y.; Seideman, J.; Sofou, S. Enhanced loading efficiency and retention of Ac in rigid liposomes for potential targeted therapy of micrometastases. Bioconjugate Chem. 2008, 19, 1274–1282. [Google Scholar] [CrossRef] [PubMed]
  169. de Kruijff, R.M.; van der Meer, A.; Windmeijer, C.A.A.; Kouwenberg, J.J.M.; Morgenstern, A.; Bruchertseifer, F.; Sminia, P.; Denkova, A.G. The therapeutic potential of polymersomes loaded with (225)Ac evaluated in 2D and 3D in vitro glioma models. Eur. J. Pharm. Biopharm. 2018, 127, 85–91. [Google Scholar] [CrossRef]
  170. Kruijff, R.M.; Raave, R.; Kip, A.; Molkenboer-Kuenen, J.; Morgenstern, A.; Bruchertseifer, F.; Heskamp, S.; Denkova, A.G. The in vivo fate of (225)Ac daughter nuclides using polymersomes as a model carrier. Sci. Rep. 2019, 9, 11671. [Google Scholar] [CrossRef]
  171. Thijssen, L.; Schaart, D.R.; de Vries, D.; Morgenstern, A.; Bruchertseifer, F.; Denkova, A.G. Polymersomes as nano-carriers to retain harmful recoil nuclides in alpha radionuclide therapy: A feasibility study. Radiochim. Acta 2012, 100, 473–481. [Google Scholar] [CrossRef]
  172. Woodward, J.; Kennel, S.J.; Stuckey, A.; Osborne, D.; Wall, J.; Rondinone, A.J.; Standaert, R.F.; Mirzadeh, S. LaPO Nanoparticles Doped with Actinium-225 that Partially Sequester Daughter Radionuclides. Bioconjugate Chem. 2011, 22, 766–776. [Google Scholar] [CrossRef] [PubMed]
  173. Cedrowska, E.; Pruszynski, M.; Majkowska-Pilip, A.; Meczynska-Wielgosz, S.; Bruchertseifer, F.; Morgenstern, A.; Bilewicz, A. Functionalized TiO(2) nanoparticles labelled with (225)Ac for targeted alpha radionuclide therapy. J. Nanoparticle Res. 2018, 20, 83. [Google Scholar] [CrossRef] [PubMed]
  174. Mulvey, J.J.; Villa, C.H.; McDevitt, M.R.; Escorcia, F.E.; Casey, E.; Scheinberg, D.A. Self-assembly of carbon nanotubes and antibodies on tumours for targeted amplified delivery. Nat. Nanotechnol. 2013, 8, 763–771. [Google Scholar] [CrossRef]
  175. Sempkowski, M.; Zhu, C.; Menzenski, M.Z.; Kevrekidis, I.G.; Bruchertseifer, F.; Morgenstern, A.; Sofou, S. Sticky Patches on Lipid Nanoparticles Enable the Selective Targeting and Killing of Untargetable Cancer Cells. Langmuir 2016, 32, 8329–8338. [Google Scholar] [CrossRef]
  176. Salvanou, E.A.; Stellas, D.; Tsoukalas, C.; Mavroidi, B.; Paravatou-Petsotas, M.; Kalogeropoulos, N.; Xanthopoulos, S.; Denat, F.; Laurent, G.; Bazzi, R.; et al. A Proof-of-Concept Study on the Therapeutic Potential of Au Nanoparticles Radiolabeled with the Alpha-Emitter Actinium-225. Pharmaceutics 2020, 12, 188. [Google Scholar] [CrossRef]
  177. Lingappa, M.; Song, H.; Thompson, S.; Bruchertseifer, F.; Morgenstern, A.; Sgouros, G. Immunoliposomal Delivery of Bi for α-Emitter Targeting of Metastatic Breast Cancer. Cancer Res. 2010, 70, 6815–6823. [Google Scholar] [CrossRef] [PubMed]
  178. Ruggiero, A.; Villa, C.H.; Holland, J.P.; Sprinkle, S.R.; May, C.; Lewis, J.S.; Scheinberg, D.A.; McDevitt, M.R. Imaging and treating tumor vasculature with targeted radiolabeled carbon nanotubes. Int. J. Nanomed. 2010, 5, 783–802. [Google Scholar] [CrossRef]
  179. Sathekge, M.; Morgenstern, A. Advances in targeted alpha therapy of cancer. Eur. J. Nucl. Med. Mol. I 2024, 51, 1205–1206. [Google Scholar] [CrossRef]
  180. Zhang, J.J.; Qin, S.S.; Yang, M.D.; Zhang, X.Y.; Zhang, S.H.; Yu, F. Alpha-emitters and targeted alpha therapy in cancer treatment. Iradiology 2023, 1, 245–261. [Google Scholar] [CrossRef]
  181. Jang, A.; Kendi, A.T.; Johnson, G.B.; Halfdanarson, T.R.; Sartor, O. Targeted Alpha-Particle Therapy: A Review of Current Trials. Int. J. Mol. Sci. 2023, 24, 1626. [Google Scholar] [CrossRef] [PubMed]
  182. Sun, J.; Yang, J.; Guo, J.; Tao, L.; Xu, B.; Wang, G.; Meng, F.; Zhong, Z. Dual-targeted alpha therapy mitigates prostate cancer and boosts immune checkpoint blockade therapy. J. Control Release 2025, 382, 113686. [Google Scholar] [CrossRef] [PubMed]
  183. Tulik, M.; Kulinski, R.; Tabor, Z.; Brzozowska, B.; Laba, P.; Bruchertseifer, F.; Morgenstern, A.; Krolicki, L.; Kunikowska, J. Quantitative SPECT/CT imaging of actinium-225 for targeted alpha therapy of glioblastomas. EJNMMI Phys. 2024, 11, 41. [Google Scholar] [CrossRef]
  184. Besiroglu, H.; Kadihasanoglu, M. The Safety and Efficacy of Targeted Alpha Therapy, Ac-225 Prostate-Specific Membrane Antigen, in Patients with Metastatic Castration-Resistant Prostate Cancer: A Systematic Review and Meta-Analysis. Prostate 2025, 85, 541–557. [Google Scholar] [CrossRef] [PubMed]
  185. Zhang, J.; Kulkarni, H.R.; Baum, R.P. Peptide Receptor Radionuclide Therapy Using 225Ac-DOTATOC Achieves Partial Remission in a Patient With Progressive Neuroendocrine Liver Metastases After Repeated beta-Emitter Peptide Receptor Radionuclide Therapy. Clin. Nucl. Med. 2020, 45, 241–243. [Google Scholar] [CrossRef]
  186. Sathekge, M.M.; Bruchertseifer, F.; Lawal, I.O.; Vorster, M.; Knoesen, O.; Lengana, T.; Boshomane, T.G.; Mokoala, K.K.; Morgenstern, A. Treatment of brain metastases of castration-resistant prostate cancer with (225)Ac-PSMA-617. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 1756–1757. [Google Scholar] [CrossRef]
  187. Sathekge, M.; Bruchertseifer, F.; Knoesen, O.; Reyneke, F.; Lawal, I.; Lengana, T.; Davis, C.; Mahapane, J.; Corbett, C.; Vorster, M.; et al. (225)Ac-PSMA-617 in chemotherapy-naive patients with advanced prostate cancer: A pilot study. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 129–138. [Google Scholar] [CrossRef]
  188. Popovtzer, A.; Rosenfeld, E.; Mizrachi, A.; Bellia, S.R.; Ben-Hur, R.; Feliciani, G.; Sarnelli, A.; Arazi, L.; Deutsch, L.; Kelson, I.; et al. Initial Safety and Tumor Control Results From a “First-in-Human” Multicenter Prospective Trial Evaluating a Novel Alpha-Emitting Radionuclide for the Treatment of Locally Advanced Recurrent Squamous Cell Carcinomas of the Skin and Head and Neck. Int. J. Radiat. Oncol. Biol. Phys. 2020, 106, 571–578. [Google Scholar] [CrossRef]
  189. Naya, M.; Murthy, V.L.; Taqueti, V.R.; Foster, C.R.; Klein, J.; Garber, M.; Dorbala, S.; Hainer, J.; Blankstein, R.; Resnic, F.; et al. Preserved coronary flow reserve effectively excludes high-risk coronary artery disease on angiography. J. Nucl. Med. 2014, 55, 248–255. [Google Scholar] [CrossRef]
  190. Krolicki, L.; Bruchertseifer, F.; Kunikowska, J.; Koziara, H.; Krolicki, B.; Jakucinski, M.; Pawlak, D.; Apostolidis, C.; Mirzadeh, S.; Rola, R.; et al. Prolonged survival in secondary glioblastoma following local injection of targeted alpha therapy with (213)Bi-substance P analogue. Eur. J. Nucl. Med. Mol. Imaging 2018, 45, 1636–1644. [Google Scholar] [CrossRef]
  191. Kratochwil, C.; Bruchertseifer, F.; Rathke, H.; Hohenfellner, M.; Giesel, F.L.; Haberkorn, U.; Morgenstern, A. Targeted alpha-Therapy of Metastatic Castration-Resistant Prostate Cancer with (225)Ac-PSMA-617: Swimmer-Plot Analysis Suggests Efficacy Regarding Duration of Tumor Control. J. Nucl. Med. 2018, 59, 795–802. [Google Scholar] [CrossRef] [PubMed]
  192. Belli, M.L.; Sarnelli, A.; Mezzenga, E.; Cesarini, F.; Caroli, P.; Di Iorio, V.; Strigari, L.; Cremonesi, M.; Romeo, A.; Nicolini, S.; et al. Targeted Alpha Therapy in mCRPC (Metastatic Castration-Resistant Prostate Cancer) Patients: Predictive Dosimetry and Toxicity Modeling of (225)Ac-PSMA (Prostate-Specific Membrane Antigen). Front. Oncol. 2020, 10, 531660. [Google Scholar] [CrossRef]
  193. Hadaschik, B. Re:(225)Ac-PSMA-617 for PSMA-Targeting Alpha-radiation Therapy of Patients with Metastatic Castration-resistant Prostate Cancer. Eur. Urol. 2016, 70, 1080–1081. [Google Scholar] [CrossRef] [PubMed]
  194. Kneifel, S.; Cordier, D.; Good, S.; Ionescu, M.C.S.; Ghaffari, A.; Hofer, S.; Kretzschmar, M.; Tolnay, M.; Apostolidis, C.; Waser, B.; et al. Local targeting of malignant gliomas by the diffusible peptidic vector 1,4,7,10-tetraazacyclododecane-1-glutaric acid-4,7,10-triacetic acid-substance P. Clin. Cancer Res. 2006, 12, 3843–3850. [Google Scholar] [CrossRef] [PubMed]
  195. Sathekge, M.; Bruchertseifer, F.; Vorster, M.; Lawal, I.O.; Knoesen, O.; Mahapane, J.; Davis, C.; Reyneke, F.; Maes, A.; Kratochwil, C.; et al. Predictors of Overall and Disease-Free Survival in Metastatic Castration-Resistant Prostate Cancer Patients Receiving (225)Ac-PSMA-617 Radioligand Therapy. J. Nucl. Med. 2020, 61, 62–69. [Google Scholar] [CrossRef]
  196. Kratochwil, C.; Bruchertseifer, F.; Rathke, H.; Bronzel, M.; Apostolidis, C.; Weichert, W.; Haberkorn, U.; Giesel, F.L.; Morgenstern, A. Targeted alpha-Therapy of Metastatic Castration-Resistant Prostate Cancer with (225)Ac-PSMA-617: Dosimetry Estimate and Empiric Dose Finding. J. Nucl. Med. 2017, 58, 1624–1631. [Google Scholar] [CrossRef]
  197. Krolicki, L.; Bruchertseifer, F.; Kunikowska, J.; Koziara, H.; Krolicki, B.; Jakucinski, M.; Pawlak, D.; Apostolidis, C.; Mirzadeh, S.; Rola, R.; et al. Safety and efficacy of targeted alpha therapy with (213)Bi-DOTA-substance P in recurrent glioblastoma. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 614–622. [Google Scholar] [CrossRef]
Figure 1. 233U decay chain.
Figure 1. 233U decay chain.
Cancers 17 03055 g001
Figure 2. Growth of [213Bi]Bi activity from [225Ac]Ac decay after generator elution.
Figure 2. Growth of [213Bi]Bi activity from [225Ac]Ac decay after generator elution.
Cancers 17 03055 g002
Figure 3. Instability of the daughter radionuclide/chelating agent complex as a result of [225Ac]Ac α-particle emission (recoil effect).
Figure 3. Instability of the daughter radionuclide/chelating agent complex as a result of [225Ac]Ac α-particle emission (recoil effect).
Cancers 17 03055 g003
Figure 4. Chelating agents for [225Ac]Ac. (a) DOTA [38,80]. (b) MeO-DOTA-NCS [80]. (c) 2B-DOTA-NCS [80]. (d) DO3A [80]. (e) DOTPA [80]. (f) TETPA [80]. (g) DOTMP [80]. (h) DTPA [80]. (i) DTPA-p-Bn-NCS [80]. (j) PEPA [103]. (k) HEHA [104]. (l) HEHA-NCS [104]. (m) CHX-A′′-DTPA [103,105]. (n) CHX-A′′-DTPA-p-Bn-NCS [103,105]. (o) EDTA [105]. (p) t-Bucalix[4]arene-tetracarboxylic acid [24,27,106]. (q) Motexafin [24,107]. (r) Lpy [24,107]. (s) Macropid [24,108]. (t) Macropa [24,108]. (u) Macropa-NCS [24,108]. (v) Bispa2 [24,108]. (w) EuK-106 [24,108].
Figure 4. Chelating agents for [225Ac]Ac. (a) DOTA [38,80]. (b) MeO-DOTA-NCS [80]. (c) 2B-DOTA-NCS [80]. (d) DO3A [80]. (e) DOTPA [80]. (f) TETPA [80]. (g) DOTMP [80]. (h) DTPA [80]. (i) DTPA-p-Bn-NCS [80]. (j) PEPA [103]. (k) HEHA [104]. (l) HEHA-NCS [104]. (m) CHX-A′′-DTPA [103,105]. (n) CHX-A′′-DTPA-p-Bn-NCS [103,105]. (o) EDTA [105]. (p) t-Bucalix[4]arene-tetracarboxylic acid [24,27,106]. (q) Motexafin [24,107]. (r) Lpy [24,107]. (s) Macropid [24,108]. (t) Macropa [24,108]. (u) Macropa-NCS [24,108]. (v) Bispa2 [24,108]. (w) EuK-106 [24,108].
Cancers 17 03055 g004aCancers 17 03055 g004b
Table 1. Current and prospective production approaches for [225Ac]Ac.
Table 1. Current and prospective production approaches for [225Ac]Ac.
[225Ac]Ac ProductionProduction ModeProduction FacilityRefs.
 Current [229Th]Th[229Th]Th generators:
Natural decay of 233U hoards
ORNL, USA
I.T.U., Germany
IPPE, Russia
[27,28,31,32,33,35]
Developed[226Ac]AcElectron Accelerators:
226Ra(γ,n)225Ra → 225Ac
Medical linacs:
ARIEL (TRIUMF, Canada)
[39,43]
Proton Accelerators
226Ra(p,2n) 225Ac
Medium-sized cyclotrons:
JRC cyclotron, Germany
1 (QST), Japan
2 NPI, Czech Republic
3 CNEA, Argentina
4 SCK-CEN, Belgium
5 KIRAMS, South Korea
[21,40,41,46]
Nuclear Reactors:
226Ra(n,2n)225Ra → 225Ac
Fast breeder reactor[50]
[232Th]ThThorium Spallation:232Th(p,x) 225AcINR, Russia
BNL, USA
LANL, New Mexico
[31,32,35,41,42]
Special facilitiesISOL, ISAC (TRIUMF, Canada)
ISOL, ISOLDE (CERN, Geneva)
MEDICIS (CERN, Switzerland)
IPF (TRIUMF, Canada)
[46,58,59]
Transformation of [226Ra]Ra to [229Th]Th226Ra(n,γ)227Ra,227Ac(n,γ)228Ac,
228Th(n,γ) 229Th
High-Flux Isotope Reactor (HFIR), ORNL, USA[60,61]
1 (QST.): National Institutes for Quantum and Radiological Sciences and Technology; 2 (NPI): Nuclear Physics Institute; 3 (CNEA): Argentine National Atomic Energy Commission; 4 (CK-CEN): Nuclear Energy Research Centre; 5 (KIRAMS): Korean Institute for Radiological and Medical Sciences.
Table 2. An overview of the most-used chelating agents for [225Ac]Ac/[213Bi]Bi-labeled molecules.
Table 2. An overview of the most-used chelating agents for [225Ac]Ac/[213Bi]Bi-labeled molecules.
Targeting AgentChelating AgentRefs.
Anti-CD33 IgG (HuM195)DMSA, DMPS, Ca-DTPA, SCNCHX-A-DTPA[106,112,113]
Anti-CD20 IgG (rituximab)DMPS, CHX-A′′-DTPA[114,115]
Plasminogen activator inhibitor type 2CHX-A′′-DTPA, cDTPA[116,117,118]
Anti-MUC1 IgG (C595 IgG)cDTPA, CHX.A′′[117,119]
Substance P DOTAGA-/DOTA[93,120]
Anti-NG2 IgG (9.2.27 IgG)cDTPA[121,122,123,124,125]
Anti-CD138 IgGp-SCN-Bn-DOTA, SCN-CHX-A′′-DTPA[126]
Anti-PSMA IgG (J591 IgG)cDTPAa[127]
C6.5K-A scFv, C6.5K-A diabodyCHX-A′′-DTPA[128]
Anti-EGFR-mAbp-SCN-Bn-CHX-A′′-DTPA, CHX-A′′-DTPA[109,129]
DOTATOC DOTA[14]
Table 3. [225Ac]Ac/[213Bi]Bi conjugates in preclinical evaluation that demonstrate the traditional delivery platforms.
Table 3. [225Ac]Ac/[213Bi]Bi conjugates in preclinical evaluation that demonstrate the traditional delivery platforms.
IsotopeCompoundTargeting MoietyCancer TypeResultRef.
[225Ac]Ac[225Ac]Ac–DOTA–HuM195HuM195 AbCynomolgus monkey leukemia.Improved efficacy with less renal toxicity.[100]
[225Ac]Ac–DOTA-trastuzumabTrastuzumabSKOV3 human ovarian cancer.Improved efficacy with less toxicity.[68]
[225Ac]Ac–DOTA–J591J591 mAbsHuman LNCaP prostate.Improved efficacy with no toxicity.[70]
[225Ac]Ac–DOTA–3F83F8 AbNMB-7 human neuroblastoma xenografts in nude mice.Increased survival with low toxicity.[140]
[225Ac]Ac-7.16.47.16.4 mAbneu-N transgenic mouse model with rat HER-2/neu expression and spontaneous lung metastases.Improved efficacy with slight renal toxicity.[141]
[225Ac]Ac–DOTA–E4G10E4G10 AbLS174T human colon xenografts in female nude mice.High affinity targeting with appropriate efficacy.[96]
[225Ac]Ac-HEHA-Mab 201BMab 201BEMT-6 mammary carcinoma.Appropriate efficacy with some toxicity.[142]
[225Ac]Ac-DOTATOCDOTATOC peptideAR42J rat pancreatic exocrine.Appropriate efficacy with less toxicity.[78]
[225Ac]Ac–DOTA–F3F3 peptideMDA-MB-435 human peritoneal carcinomatosis in SCID mice.Improved targeting and efficacy with minor renal toxicity.[79]
[225Ac]Ac-Pep-1LPep-1L peptideU8251 human glioblastoma orthotopic xenografts in male nude mice.U8251 human glioblastoma.
Orthotopic xenografts in male nude mice.
[143]
[225Ac]Ac-DOTA–MC1RLMC1RL peptideMEL270 human uveal melanoma xenografts in SCID mice.Decreased metastasis and improved survival.[144]
[225Ac]Ac-DOTAZOLZoledronic acidWistar rats.Improved bone/blood ratio with some renal toxicity. [145]
[213Bi]Bi[213Bi]Bi-30F11-CHX-A30F11 AbFemale BALB/c mice.High renal toxicity.[146]
[213Bi]Bi-CD138Anti-CD138 Ab5T33 mouse multiple myeloma cell culturedinC57BL/KaLwRij mice.Improved survival with moderate toxicity.[147]
[213Bi]Bi-d9Mabd9MabHSC45-M2 human gastric cancer cells in nude mice.Improved efficacy with low toxicity.[148]
[213Bi]Bi-matuzumabMatuzumabEJ28 human orthotopic bladder xenografts in nude mice.Increased efficacy with some renal toxicity.[129]
[213Bi]Bi-FabCO17-1A FabGW-39 human colon cancer in xenograft mice.High efficacy with low toxicity.[149]
[213Bi]Bi-CHX-A-DTPA HuM195HuM195 mAbNormal BALB/c mice.Improved pharmacokinetic parameters.[150]
[213Bi]Bi-d9-E-cad mAbd9-E-cad mAbHSC45-M2 human gastric xenografts with d9-E-cad mutation in female nude mice.Improved efficacy upon double administration compared to a single one, with no toxicity.[151]
[213Bi]Bi- HuCC49DCH2Humanized CC49
mAb
LS-174T human colon xenografts in female nude mice.Improved efficacy.[152]
[213Bi]Bi-P-P4DP-P4D peptideOV-MZ-6 human ovarian xenografts in female nude mice.High tumor accumulation with low nephrotoxicity.[153]
[213Bi]Bi-DOTATOCDOTATOC peptideCA20948 rat pancreatic adenocarcinoma tumors in Lewis rats.Improved efficacy with low toxicity.[154]
[213Bi]Bi–DOTA–PESIN, or
[213Bi]Bi-AMBA
PESIN and AMBA
Peptides
PC-3 human prostate xenografts in female nude mice.[213Bi]Bi–DOTA–PESIN had lower nephrotoxicity than [213Bi]Bi-AMBA.[155]
Table 4. Summary of the preclinical studies of [225Ac]Ac/[213Bi]Bi-Nanoradiopharmaceuticals.
Table 4. Summary of the preclinical studies of [225Ac]Ac/[213Bi]Bi-Nanoradiopharmaceuticals.
IsotopeNano-PlatformTargeting MoietyCancer TypeObjectiveRef.
[225Ac]AcPEGylated liposomePSMA J591 antibody(PSMA)- expressing cellsReduction in the recoil effect[82]
PEGylated liposomeTrastuzumabSKOV-3 ovarian cellsRecoiling[168]
PEGylated liposome--Recoiling[88]
PEGylatedMUVELTrastuzumabSKOV-3 ovarian cellsRecoiling[81]
Polymersomes--Recoiling[169]
Polymersomes--Recoiling[170]
Polymersomes--Recoiling[171]
(La0.5Gd0.5)PO4MAb 201bEMT-6 lung tumor cellsRecoiling[85]
La ([225Ac]Ac)PO4MAb 201bEMT-6 lung tumor cellsRecoiling[172]
TiO2Substance P (5-11)NK1 glioma receptorRecoiling[173]
Carbon nanotubes-DOTALintuzumabRituximabanti-A33CD20 + B-cell lymphoma
C33 + myelocytic leukemia
A33 + colon adenocarcinoma
Targeted delivery and
rapid clearance
[174]
Lipid vehicleTrastuzumabBT-474, MDA-MB-231,
MCF7 breast carcinoma cell
Improved targeting of low HER2-expression cells[175]
Carbon nanotubes-DOTATumor
neovascular-targeting
antibody
LS174T xeno-graft tumor
model
Targeted delivery and
rapid clearance
[136]
Gold NPs-DOTA-U87 glioblastoma cancer cellsLocalized cancer treatment[176]
[213Bi]BiImmunoliposomesCHX-A-DTPARat/neu transgenic mouse model of mammary carcinomaImprove the therapeutic efficacy[177]
Table 5. [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals on preclinical and clinical phases.
Table 5. [225Ac]Ac/[213Bi]Bi-radiopharmaceuticals on preclinical and clinical phases.
Tumor[225Ac]Ac Agent[213Bi]Bi Agent
Neuroendocrine[225Ac]Ac-DOTATOC[213Bi]Bi-DOTATOC
Multiple myeloma[225Ac]Ac-BC8[213Bi]Bi-labeled 9.E7.4 anti-CD138 mAb
[225Ac]Ac-lintuzumab[213Bi]Bi-labeled Anti-CD138 IgG
Metastatic castration
resistant prostate cancer
[225Ac]Ac-PSMA617[213Bi]Bi-DOTA-PESIN
Glioblastoma[225Ac]Ac-E4G10[213Bi]Bi-DOTA-substance P
Lymphoma[225Ac]Ac-anti-CD33 HUM195[213Bi]Bi-DOTA-biotin
[213Bi]Bi-anti-CD20-mAb 12
Leukemia[225Ac]Ac-lintuzumab[213Bi]Bi-lintuzumab
[225Ac]Ac-anti-CD33-mAb[213Bi]Bi-HuM195
[213Bi]Bi-anti-CD33-mAb 49
Breast cancer[225Ac]Ac-7.16.4 anti-rat HER-2/neu[213Bi]Bi-CHX-A″DTPA-C6.5K-A scFv
[213Bi]Bi-Plasminogen activator inhibitor type 2
Melanoma[225Ac]Ac-crown-_MSH[213Bi]Bi-anti-MCSP-mAb 54
[213Bi]Bi-Cdtpa-Anti-NG2 IgG (9.2.27 IgG)
Glioma[225Ac]Ac-Substance P 20[213Bi]Bi-Substance P 68
Neuroendocrine
Tumors
[225Ac]Ac-DOTATOC 39[213Bi]Bi-DOTATOC 25
Prostate cancer[225Ac]Ac-PSMA617 > 400[213Bi]Bi- cDTPAa- Anti-PSMA IgG (J591 IgG)
Advanced refractory
Solid tumors
[225Ac]Ac-FPI-1434Information is not available
Colorectal cancerInformation is not available[213Bi]Bi-labeled CO-1A Fab’
Bladder carcinomaInformation is not available[213Bi]Bi-anti-EGFR mAb
Peritoneal carcinomaInformation is not available[213Bi]Bi-d9MAb
Ovarian cancerInformation is not available[213Bi]Bi-CHX-A″DTPA-C6.5K-A scFv
[213Bi]Bi-CHX-A″DTPA- C6.5K-A diabody
[213Bi]Bi-Anti-MUC1 IgG (C595 IgG)
Non-Hodgkin lymphomaInformation is not available[213Bi]Bi-Anti-CD20 IgG (rituximab)
Pancreatic cancerInformation is not available[213Bi]Bi-Plasminogen activator inhibitor type 2
[213Bi]Bi-Anti-MUC1 IgG (C595 IgG)
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

Nawar, M.F.; Selim, A.A.; Essa, B.M.; El-Daoushy, A.F.; Swidan, M.M.; Chambers, C.G.; Al Qahtani, M.H.; Smith, C.J.; Sakr, T.M. Actinium-225/Bismuth-213 as Potential Leaders for Targeted Alpha Therapy: Current Supply, Application Barriers, and Future Prospects. Cancers 2025, 17, 3055. https://doi.org/10.3390/cancers17183055

AMA Style

Nawar MF, Selim AA, Essa BM, El-Daoushy AF, Swidan MM, Chambers CG, Al Qahtani MH, Smith CJ, Sakr TM. Actinium-225/Bismuth-213 as Potential Leaders for Targeted Alpha Therapy: Current Supply, Application Barriers, and Future Prospects. Cancers. 2025; 17(18):3055. https://doi.org/10.3390/cancers17183055

Chicago/Turabian Style

Nawar, Mohamed F., Adli A. Selim, Basma M. Essa, Alaa F. El-Daoushy, Mohamed M. Swidan, Claudia G. Chambers, Mohammed H. Al Qahtani, Charles J. Smith, and Tamer M. Sakr. 2025. "Actinium-225/Bismuth-213 as Potential Leaders for Targeted Alpha Therapy: Current Supply, Application Barriers, and Future Prospects" Cancers 17, no. 18: 3055. https://doi.org/10.3390/cancers17183055

APA Style

Nawar, M. F., Selim, A. A., Essa, B. M., El-Daoushy, A. F., Swidan, M. M., Chambers, C. G., Al Qahtani, M. H., Smith, C. J., & Sakr, T. M. (2025). Actinium-225/Bismuth-213 as Potential Leaders for Targeted Alpha Therapy: Current Supply, Application Barriers, and Future Prospects. Cancers, 17(18), 3055. https://doi.org/10.3390/cancers17183055

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