Next Article in Journal
Low Water Absorption, High-Strength Polyamide 6 Composites Blended with Sustainable Bamboo-Based Biochar
Next Article in Special Issue
Hydroxyapatite and Titanium Dioxide Nanoparticles: Radiolabelling and In Vitro Stability of Prospective Theranostic Nanocarriers for 223Ra and 99mTc
Previous Article in Journal
Effects of Different TiO2 Nanoparticles Concentrations on the Physical and Antibacterial Activities of Chitosan-Based Coating Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nanoparticles in Targeted Alpha Therapy

by
Agnieszka Majkowska-Pilip
*,
Weronika Gawęda
,
Kinga Żelechowska-Matysiak
,
Kamil Wawrowicz
and
Aleksander Bilewicz
Centre of Radiochemistry and Nuclear Chemistry, Institute of Nuclear Chemistry and Technology, Dorodna 16, 03-195 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Submission received: 20 June 2020 / Revised: 4 July 2020 / Accepted: 9 July 2020 / Published: 13 July 2020
(This article belongs to the Special Issue Isotopes Labeled Nanoparticles)

Abstract

:
Recent advances in the field of nanotechnology application in nuclear medicine offer the promise of better therapeutic options. In recent years, increasing efforts have been made on developing nanoconstructs that can be used as carriers for immobilising alpha (α)-emitters in targeted drug delivery. In this publication, we provide a comprehensive overview of available information on functional nanomaterials for targeted alpha therapy. The first section describes why nanoconstructs are used for the synthesis of α-emitting radiopharmaceuticals. Next, we present the synthesis and summarise the recent studies demonstrating therapeutic applications of α-emitting labelled radiobioconjugates in targeted therapy. Finally, future prospects and the emerging possibility of therapeutic application of radiolabelled nanomaterials are discussed.

1. Introduction

Radiation therapy is an important component of cancer treatment with approximately 50% of all cancer patients receiving radiation therapy during their course of illness [1]. The key objective in cancer radiotherapy is to achieve a high therapeutic efficacy by maximising damage to the tumour while minimising damage to surrounding healthy tissue. In contrast to radiation therapy methods that use an external ion beam source, internal radiotherapy is performed by the direct administration of radionuclides conjugated to a targeting vector. This can be achieved by coupling suitable radionuclides to antibodies, antibody fragments, nanobodies or small peptides that bind cell surface receptors or other proteins specifically overexpressed by cancer cells.
At the beginning, beta (β)-emitting radionuclides were widely used in cancer therapy. There are two registered therapeutic antibodies: tositumomab labelled with 131I (BEXXARTM) that is used to treat follicular lymphoma, and ibritumomab labelled with 90Y (Zevalin®) is used to treat B cell non-Hodgkin’s lymphoma [2]. Furthermore, 90Y and 177Lu-labelled peptides, like somatostatin and bombesin analogues, showed promising results in current clinical trials [3]. In February 2018, the U.S. Food and Drug Administration (FDA) approved DOTATATE labelled with 177Lu (Lutathera®) for the treatment of certain neuroendocrine tumours [4]. One of the limitations of β- emitting radionuclide therapies is their inability to treat small clusters of cancer cells like micrometastatic cancers or single leukaemia cells. This is because long ranges of the β- particles cause death in normal healthy cells [5]. For example, β--particles from 90Y (Eβ(max)) = 2.3 MeV deposit their energy over a range of 12 mm. The effective tissue range of β--particles is not optimal for the treatment of tumours as small clusters of cells or of single cells and micrometastases, because much of the decay energy is deposited outside the boundary of the tumour. On the other hand, targeted radiotherapy using α-particles is a promising alternative to that based on β--particles, because α-particles deposit whole of their energy within a few cell diameters (50–100 µm) [6]. The α-particle, a 4He nucleus, is relatively heavier than other subatomic particles emitted from decaying radionuclides and has a much shorter range in tissues. Compared with β--particles, α-particles provide a very high relative biological effectiveness, killing more cells with less radioactivity. Accordingly, α-particles have roughly 500 times higher cytotoxic potency than β--particles [7]. Just 15 α tracks through the nucleus of a cell are sufficient to cause apoptosis due to their high energy deposited per unit distance travelled (approximately 80 keV/mm) [8].
Recently, it has been found that in some cancers, such as in the case with leukaemia, breast and brain cancer, small subpopulations of tumour cells are able to self-renew and reconstitute the heterogeneous tumour cell population [9,10]. These stem-like cancer cells are also thought be involved in the widespread metastatic dissemination of cancer [11]. These findings suggest that failure in cancer treatment may be associated with the failure to eradicate cancer stem cells [12]. Cancer stem cells are usually resistant to chemotherapy, as well as external and internal radiotherapy including β--emitters [13]. For this reason, effective targeted therapies for the complete eradication of these cells are urgently needed. Given the properties outlined above, tumour stem cells are ideal targets for targeted α particle therapy [14]. For example, Substance P labelled with α-emitter 225Ac is highly cytotoxic to chemo- and radioresistant glioblastoma multiforme (GBM) stem cells [15].
As with any medical methods, there are some limitations using α-particle therapy. First, is the availability. There are more than 100 radionuclides that decay by α-emission, but the majority of these radionuclides have half-lives either too short or too long for any therapeutic use, their production is not economically viable or chemical properties do not allow their use in medicine. Currently, the main α-emitters used for targeted therapy are 223,224Ra, 211At, 225Ac, 212Pb, 227,226Th and 212,213Bi. Table 1 summarises the main nuclear properties of medically used α-emitters.
However, these radionuclides have shortcomings. In the case of 212Bi, 213Bi and 226Th, the short half-life often limits the application of these nuclides to situations where tumour cells are rapidly accessible to the targeting agent. Moreover, 212Bi shows high-energy α-emission with 32% abundance that is absent in 213Bi. Therefore, the latter is generally considered to be a more attractive candidate for α-radiotherapy. In the case of 225Ac, 223Ra and 227Th, recoiling of α-emitting decay products from radiobioconjugates can accumulate in critical organs. 225Ac decays directly to 221Fr (alkali metal) that has a half-life of 4.9 min and then escaping from the 225Ac-radiobioconjugate. A similar situation appears in the case of 227Th and 223Ra where the decay product, the gaseous 219Rn, easily liberates itself from 227Th and 223Ra radiobioconjugates. It is worth mentioning that this problem is reduced in the case of 223Ra series because 75% of its total alphas are delivered within a few seconds (t1/2 = 4 s) after the 223Ra decay [16].
Additionally, the lack of appropriate bifunctional ligands is the reason why radium radionuclides did not find application in a receptor targeted therapy. The Ra2+ cation, like other cations of group 2, forms weak complexes. Therefore, labelling the biomolecules with 223Ra is a very difficult task. Until now, only 223Ra in the simple form of RaCl2, with its natural affinity to build into bones, finds application in the treatment of bone metastases from breast and prostate cancer [17]. The second alpha emitter whose binding to the biomolecule is a challenging is 211At.
In the last two years, a number of review articles on targeted α therapy [18,19,20,21] as well as several reviews concerned radioactive nanoparticles and their use in diagnostics and therapy have been published [22,23,24]. These reviews contain also some information on nanoparticles labelled with α emitters; however, this issue is briefly described. Below we summarise existing works on α emitter radiolabelled functional nanomaterials and provide illustrative examples of their application in nuclear nanomedicine. The first section focuses on the problem of the reasons for the use of nanostructures in targeted α therapy. Next various nanomaterials investigated thus far and biomedical results of these studies are briefly described. Finally, future perspectives and the emerging potential for medical application of alpha emitter labelled nanomaterials are discussed.

2. Why Nanoconstructs Are Used for α Radionuclide Therapy?

There are three important reasons why nanoparticles are used in α radionuclide therapy. The first two are related to the difficulty of attachment of the radionuclides to the targeting vectors.
  • Release of daughters from the radioisotopes, that are the mother radionuclides for the decay chains (225Ac, 223Ra, 227Th and 212Pb)
  • Lack of appropriate bifunctional ligands for effective binding of α–emitters to targeted molecules (211At and 223Ra)
  • Application of α-emitter-labelled nanoparticles in new therapeutic approach–targeted nanobrachytherapy.
These reasons are described in more detail below.

2.1. Controlling the Recoil of the Daughter Radionuclides

Radionuclides 225Ac, 227Th and 223Ra are good candidates for α therapy because they have a relatively long half-life and high cumulative decay energy (>28 MeV). However, all of these α-emitters have four or five unstable daughter nuclides that often emit α-particles as well. Figure 1 presents decay chains for 225Ac and 227Th-223Ra radionuclides. Important limitation for the application of these α-emitters in radionuclide therapy is the escape of daughter radionuclides from radiopharmaceuticals.
From the simple formula of the conservation of angular momentum, the decay energy is divided between the nucleus and the α particle:
E r = Q m α M r
where Er is the recoil energy, mα the rest mass of α particle, Mr mass of the recoil nucleus and Q is the decay energy. Because the mass of an α-particle is 4 amu (atomic mass units) and the mass of recoiled nucleus is ~210–220 amu, the energy distribution between the α-particle and the recoiling atom is typically as 1 to 50 [25].
As the energy of α decay is usually between 4 and 8 MeV, the daughter nuclide typically receives ~0.1 MeV of recoil energy. This is ~1000 times higher than the chemical binding energy, meaning the daughter radionuclide cannot be held by a chemical bond. Therefore, the sequestration of daughter radionuclides in chelating ligands, such as cyclic or linear polyamino carboxylate chelators like 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (DOTA) or diethylenetriaminepentaacetic acid (DTPA), is not possible. In this case, all daughter nuclides produced from α-emitting radionuclides are released from their ligand in vivo, limiting the dose that can be delivered to the target cells. After dissociating from the radiobioconjugate, free daughter radionuclides may cause harm to healthy tissue and initiate secondary tumourigenesis. This is particularly true when the daughter nuclides themselves are α-emitters. The transfer of the daughter radionuclide depends on its half-life, diffusion and affinity for certain organs. 213Bi from 225Ac decay migrates to the kidney causing renal toxicity. This renal toxicity can be partially moderated through the use of scavengers or addition of non-radioactive Bi3+. However, kidney toxicity remains a main limitation to application of 225Ac in radiotherapy [26,27]. Schwartz et al. [28] evaluated the contribution of nonequilibrium 213Bi to kidney dose in mice via γ-ray spectroscopy. The average absorbed dose into the kidneys was 0.77 Gy·kBq−1, where 60% was attributed to nonequilibrium 213Bi excess. There is less of a problem is with the 223Ra series because 75% of its total alphas are delivered within a few seconds (t1/2 = 4 s) after the 223Ra decay. This problem is more pronounced with the 225Ac series, because the 225Ac decays directly to 221Fr that has a t1/2 = 4.9 min. Additionally, as an alkali metal cation, it can be transported over a relatively long distance [16].
As discussed in the recent review by de Kruijff et al. [29], there are three different approaches to deal with this recoil problem: cell internalisation, local administration or encapsulation of α-emitters in nanocarriers.
Cell internalisation approach assumes the accumulation of radiopharmaceuticals inside cancer cells and keeps all daughter nuclides in the target cells. The remaining not adsorbed part of the radioconjugate is excreted fast from the body. The volume of the cell is usually large enough to keep inside most recoiling daughter radionuclides. This can be achieved only when the blood circulation time of radiobioconjugates is short and radiopharmaceutical rapidly accumulates inside the cancer cells. This strategy was applied to α-emitter-labelled internalised peptides like vascular tumour-homing peptide F3 [30], octreotide [31] and small fragments of monoclonal antibodies such as nanobodies [32]. However, this is particularly problematic in the case of 225Ac-labelled radiopharmaceuticals where 221Fr (t1/2 = 4.9 min) is first decay product. Francium as potassium analogue excreted from the cell by the Na+/K+ pump with subsequent decay (generation of the rest of the 3 α-emissions) happening outside the target cells [23].
The next approach is injecting the α-emitting radionuclides locoregionally in or near the tumour tissue, or in the cavity after tumour resection. The radiobioconjugate must be applied in a region with no or slow exchange with the surrounding tissue in order to ensure that no daughter radionuclides may infiltrate blood circulation [23]. Such strategy has been tested in Phase I clinical studies with 213Bi-DOTA-substance P locally injected in gliomas by Cordier et al. [33] and Królicki et al. [34]. More recently, a pilot study on the locoregional treatment of bladder cancer (carcinoma in situ) using the 213Bi-labelled anti-EGFR monoclonal antibody cetuximab was conducted in collaboration of Joint Research Center Karlsruhe and Technical University Munich, Germany [35]. The therapy was found to be safe and without any side effects as no activity of 213Bi was detected outside the bladder. Królicki et al. initiated a dose escalation study investigating the intratumoural/intercavitary injection of 225Ac-DOTAGA-[Thi8, Met(O2)11]-substance P [36]. The patients were treated with activities ranging from 10 to 42 MBq 225Ac-DOTAGA-[Thi8, Met(O2)11]-substance P. The treatment was well tolerated and the analysis of therapeutic efficacy and patient recruitment is ongoing. Preliminary results indicate negligible diffusion of 213Bi from the injection site.
The third option for preventing escape of daughter radionuclides from the target site is encapsulate the mother radionuclide in a nanoparticle that is big enough to physically sequester all recoils in the decay chain in its structure. The major advantage of incorporating 225Ac, 223Ra and 227Th within the nanoparticles is that the daughter radionuclides are sequestered at the site of targeting, preventing the nonspecific radiotoxicity. The size, shape and type of material needed to fully encapsulate 225Ac decay products have been comprehensively analysed by Holzwarth et al. [37]. They calculated the probability of retention 225Ac decay products in three-dimensional space after α disintegration of completely random orientations. They found that 12 nm gold layer or 39 nm graphite layer may prevent the release of 221Fr, the first decay product of 225Ac. To sequester all radionuclides up to the third daughter, a layer of 35 nm gold or 143 nm of graphite is necessary. However, it should be noted that for these calculations the authors assumed that the mother radionuclides of the decay chains must be localised at the centre of the spherical nanoparticles. We describe this problem in more detail presenting experimental results below.

2.2. Nanoparticles as Agent for Binding of Alpha with Biomolecules

Another important limitation that affects the use of targeted alpha therapy is the availability and price of the radionuclides. 211At and 223Ra radionuclides are readily available. Unfortunately, the difficulty of stable attachment of these radionuclides to the carrier biomolecule substantially limits their use in targeted α therapy. 223Ra decays to stable lead and bismuth through a cascade of short-lived α- and β--particle emitters, releasing a total energy of ~28 MeV (Figure 1). Only 223Ra in the simple ionic form of RaCl2 has application in radiotherapy for the treatment of bone metastasis. Ra2+ cations, like other cations of the 2 group Mendeleev Table, form very weak complexes. Therefore, labelling biomolecules with 223Ra is a very difficult task. The stability constant of the Ra-DOTA complex is unknown, but using log K value for Ba2+ (12.6) [38] and difference between K values of Ba-EDTA = 9.88 ± 0.11 and Ra-EDTA = 9.11 ± 0.09 [39] can be estimated as 12 and is 10 orders of magnitude lower than for the Ac-DOTA complex.
Previous studies on binding of 223Ra to biomolecules by complexation of Ra2+ by tetraazacarboxylic acids, cryptands and calixarenes were unsuccessfully. Henriksen et al. [40] compared the most suitable ligands for complexation of 223Ra, cyclic compounds 1,4,7,10-tetraazacyclododecane- 1,4,7,10 tetraacetic acid (DOTA), 4,7,13,16,21,24-hexaoxa-1,10-diazabicyclo8.8.8-hexacosane (Kryptofix2.2.2), 5,11,17,23-tetra-tert-butyl-25,26,27,28-tetrakis(carboxymethoxy), calix[4] arene-tetraacetic acid (calix[4]-tetraacetic acid) and the open-chain chelator diethylene triamine-N;N′;N″-pentaacetic acid (DTPA). Based on the relative extraction constants, calix[4]-tetraacetic is the most promising of the compounds tested. However, rapid dissociation indicates that the calix[4]-derivative is not suitable for in vivo application. Recently, Gott et al. [41] employed an innovative approach using polyoxopalladates to stably incorporate radium radionuclides as central cations. During the synthesis, the polyanion is formed with Ra2+ as central and counter ions. Unfortunately, after chromatographic purification, a portion of the unincorporated Ra2+ remains on the polyanion surface. This hinders the use of this complex. Therefore, the use of nanoparticles to joint 223Ra with biological vectors can be a best way for the application of 223Ra in targeted α therapy.
Cyclotron obtained 211At is also a promising candidate for targeted α-radiotherapy because is 7.2 h half-life assures sufficient time for its transportation, synthetic chemistry, multistep labelling, quality control and clinical application without problems caused by the daughter emitting α-particles. Another merit of this radionuclide is the simplicity of its production in cyclotrons in 209Bi(α,2n)211At nuclear reaction, followed by simple dry distillation and isolation from irradiated bismuth target. A variety of 211At labelled biomolecules have been examined in preclinical models; two have reached the stage of initial clinical trials [42,43]. Unfortunately, initial results were not positive due to instability of the astatine–biomolecule bond in biological fluids. Astatine is a member of the halogen group and shows some metallic character in the + 1 oxidation state [44]. The energy carbon–halogen bond for astatine is significantly lower than for iodine, which excludes the use of elaborated iodination methods for labelling biomolecules with 211At [45]. More stable astatinated proteins have been prepared by acylation of astatobenzoic acid derivatives prepared from trialkylstannyl precursors [46]. Unfortunately, biomolecules labelled by this method is not always stable with respect to in vivo deastatination [47]. Because of these difficulties, non-traditional solutions labelling carboranes [48], calixarenes [49], astatine hypervalent compounds [50] and 211At-Rh(III) [51] complexes have been reported with little success. Therefore, we believe that the use of nanostructures can solve this problem. The attempts of such an approach are discussed below in the discussion of individual radionuclides.

2.3. Application of α-Emitter Labelled Nanoparticles in New Therapeutic Approach—Targeted Nanobrachytherapy

Systemic radiotherapy involves the delivery of a soluble substance into the body, usually by injecting a radiopharmaceutical, which accumulates in the pathologically changed tissues. Another type of internal radiation therapy called brachytherapy involves placing closed radiation sources within or near the tumour using minimally invasive procedures. However, they require a complicated implantation technique under general anaesthesia. Furthermore, seed migration may also occur after implantation, and seed removal is required [52].
Recently, Reilly et al. proposed a novel targeted nanobrachytherapy approach for the treatment of locally-advanced breast cancer. This strategy uses intratumoural injection of 30 nm diameter gold nanoparticles modified with polyethylene glycol (PEG) chains with DOTA ligand that complex the therapeutic radionuclide 177Lu and linked to panitumumab that bind the AuNP to epidermal growth factor receptor (EGFR)-positive tumour cells or linked to Trastuzumab that bind to HER2 receptors [53,54]. Their studies on organ distribution shows that Au-Trastuzumab delivered intratumourally are retained (~30% ID/g) with minimal uptake by the liver and spleen [55]. This approach has also been proposed for 111In radiobioconjugates for application in Auger electron therapy [56].
There are several advantages of this type of injectable approach compared with conventional brachytherapy. Administration is easier and less invasive. Radiolabelled nanoparticles are in a microscopically dispersed in liquid form and can be injected intra- or peritumourally by syringe and needle. This strategy is less invasive than inserting solid seeds into the body. Additionally, their nanometer size permits local diffusion from the injection site, thus further homogenising the radiation dose deposition in the tumour [52]. Multiple small injections can achieve a more homogeneous distribution of the gold nanoseeds. Recently, a number of studies have extended this method to α-emitters such as 211At and 225Ac. This will be discussed in Section 3.

3. Nanoconstructs Labelled with α-Emitters

In Section 2, we presented the reasons why various nanoconstructs were used to immobilise α-emitters. Table 2 shows the data for labelled nanomaterials, summarised from the available literature.

3.1. Radium-223

223Ra in the simple form of radium dichloride ([223Ra]RaCl2) is the first α-particle emitting therapeutic agent approved by the FDA for bone metastatic castration-resistant cancers [102]. However, as mentioned before, Ra2+ does not form stable complexes with conventional bifunctional chelators, similar to most alkali earth metals. To extend the use of 223Ra to other than bone metastases applications, several nanomaterials have been tested for stable incorporation of radium and linking to the targeting vector. The first studies used liposomes. As presented on Figure 2 liposomes are spherical vesicles having at least one lipid bilayer.
Liposomes have an aqueous solution core surrounded by a hydrophobic membrane. Hydrophilic solutes that are dissolved in the core cannot readily pass through the bilayer membrane [103]. Liposome drug delivery systems have produced substantial results in cancer therapy with various products reaching the marketing phase. In 1995, the first FDA-approved nanodrug was Doxil®, a pegylated liposome containing doxorubicin. It was reported that Doxil® liposomes preferentially accumulate in mouse model tumours [104] as well as in patients with primary and metastatic disease [105]. In the first studies on immobilisation of 223Ra in liposomes, authors demonstrated that liposome-encapsulated radium could be prepared from preformed liposomes by ionophore-mediated loading [16]. 223Ra was incorporated with a good loading yield and was stably retained for several days when incubated at 37 °C in serum. In the next studies, Larsen et al. [57] used pegylated liposomal doxorubicin Caelyx®/Doxil® drug to synthesise radiobioconjugates containing 223Ra that were functionalised by two vectors: folic acid and fragment of monoclonal antibody F(ab’)2(IgG1). These biodistribution studies showed the blood clearance of injected liposomal radium was much slower than the free 223Ra, as expected for pegylated liposomes. Among soft-tissue organs, the highest uptake was observed in the liver and spleen. Bone uptake also increased with time, most likely because 223Ra liposomes are metabolised by macrophages in the reticuloendothelial system. This process generates free cationic 223Ra, which is eliminated either by intestinal or renal clearance, or is incorporated onto bone surfaces. In the case of 223Ra decay product, biodistribution data show less than expected from radioactive equilibrium, activity of 211Pb and 211Bi in urine and spleen, and some accumulation of 211Pb and 211Bi in the kidneys. This indicates incomplete retention of decay product inside the liposome. In another report, the same group investigated the distribution and tumour-targeting properties of 223Ra encapsulated in the same liposome in a human osteosarcoma xenograft mice model and in a dog model with spontaneous osteosarcoma [58]. In the xenograft model they found higher retention activity in the tumour in comparison to soft tissue. In the dog, the uptake was considerably higher in both calcified and non-calcified tumour metastases of different organs compared to normal tissue. Unfortunately, these promising studies did not continue after 2006.
Further studies focused on the possibility of using inorganic nanocarriers to bind 223Ra on the surface of nanoparticles or by incorporating into the nanoparticle structures. Few studies have considered the high similarity between Ba2+ and Ra2+ cations when incorporating 223Ra into the crystal structure of insoluble barium salts. The ionic radii of both cations are nearly identical, 142 pm and 148 pm, respectively [106]. Therefore, it is easy to exchange the Ba2+ for Ra2+ cations. Simple co-precipitation of Ra2+ with BaSO4 is widely used in analytical procedures for environmental determination of 226Ra. Because BaSO4 is not toxic and is easily synthesised for the use of small nanoparticles, Reissig et al. [59] proposed BaSO4 nanoparticles as carriers to stably bind radium radionuclides to a targeting molecule. In a one-pot synthesis, radium-doped alendronate-functionalised BaSO4 nanoparticles were obtained from [224Ra]Ra(NO3)2, (NH4)2SO4, BaCl2 and alendronate. Alendronate, a bisphosphonate containing an amine group, was used as a linker to attach phosphonate groups to the surface of BaSO4. The amino group can be used to form a peptide bond with the targeting biomolecule. In optimal conditions, the authors obtained [224Ra]BaSO4 nanoparticles with a hydrodynamic radius of 140 nm. Unfortunately, it was found that only approximately 20% of 224Ra were incorporated into the nanoparticles, whereas 80% of the activity remained in the supernatant solution. The radiolabelled product showed very high stability with <5% activity released. In the next publication, the same group synthesised much smaller 224Ra-labelled BaSO4 nanoparticles [60]. Moreover, labelling efficiency was better and was ~30%. This was sufficient for future therapeutic applications. In the next step, the authors plan conducting in vitro and in vivo studies to examine the diagnostic and therapeutic properties of targeted bioconjugated [224Ra]BaSO4.
In another work, Gawęda et al. [69] proposed incorporating 223Ra into barium ferrite nanoparticles. Nanoparticles with a diameter of 15–30 nm are obtained in a one-pot hydrothermal synthesis (210 °C, 6.5 h). The labelling efficiency exceeded 90% due to the high similarity of Ra2+ and Ba2+. No leakage of 223Ra was observed in the human serum stability test. Leakage of 211Pb (decay product of 223Ra) was ~15%. A ethylphosphonoacetate linker was used to attach the targeting biomolecule (Trastuzumab).
In a series of publications, the possibility of using hydroxyapatite (HAP) nanoparticles to immobilise 223Ra is described [61,62,63,64]. HAP is a naturally occurring mineral form of calcium phosphate with the formula Ca5(PO4)3(OH). HAP is the main inorganic constituent of bones and teeth and is well tolerated by the living organism. In nuclear medicine, HAP microparticles are used to transport 177Lu for liver cancer therapy [107] and can be labelled with 99mTc for the diagnosis of bone cancer [108]. Kozempel et al. [63] tested HAP nanoparticles as a carrier of 223Ra. They proposed two strategies for 223Ra-HAP nanoparticles preparation: The first method was based on the surface sorption of 223Ra on ready-made HAP nanoparticles. The latter was based on the intrinsic incorporation of 223Ra into HAP structures during synthesis. In both methods, labelling yields were high and stability was acceptable, though the internal strategy gave slightly better stability results. Surprisingly, no significant release of activity was detected in the stability tests probably due to readsorption of liberated 211Bi and 211Pb. A similar situation was observed in other nanomaterials like nanozeolites [65] and barium ferrites [69] where readsorption was also the source of large retention of 223Ra decay products, 211Bi, and 211Pb. In subsequent studies devoted to labelling HAP nanoparticles with 223Ra, Vasiliev et al. obtained similar results [62]. They found that the optimal labelling of HAP nanoparticles is obtained by the intrinsic incorporation of 223Ra into HAP structure during synthesis at pH 4–7, followed by annealing at 900 °C. When synthesised under these conditions, 223Ra-HAP nanoparticles exhibited high stability while retaining 95% Ra activity. The same group presented the dynamics of 223Ra penetration into porous HAP granules and its redistribution in sorption–desorption processes.
Suchankova et al. [64] tested the labelling of two HAP nanoparticles with 223Ra in a Britton–Robinson buffer solution within a pH range of 2 to 12. Both nanomaterials >pH 6 showed a sorption higher than 95% of 223Ra. Using the applied chemical equilibrium model, they postulated the most important species playing a role in sorption were RaCO3, RaPO4, RaHPO4 and Ra(Ac)2 on the edge sites, and Ra2+ and RaH2PO4+ on layer sites.
A promising approach was proposed by Piotrowska et al., using nanozeolites NaA as 223Ra carriers [65,66]. Unlike other proposed solutions, 223Ra labelling occurs by a Na+ for Ra2+ ion exchange process after synthesis of the bioconjugate. In other systems, radioactive nanoparticles are first prepared and then modified, which causes necessity of working with radioactive materials and loss of radioactivity by 223Ra decay. It was shown that NaA nanozeolite strongly binds radium radionuclides and its decay products [65]. NaA nanozeolites were conjugated by silan linker to substance P(5-11)-peptide exhibited high affinity to NK1 receptor on glioma cell [66]. The 223RaA–silane–PEG–SP(5–11) bioconjugate successfully retained 99% of 223Ra and 95% of the daughter radionuclides without compromising the tumouricidal radiation properties. This retention was higher than expected for the size of the nanozeolites and was explained by resorption of decay products 219Rn and 211Pb on the nanoparticle due to high affinity of zeolite for Rn and Pb2+ [66]. However, according to Holzwart et al. [37], these results were obtained using nanozeolite labelled with 223Ra equilibrated with human serum and cannot be transferred to in vivo models. Blood flow may rapidly dislocate the decay products from the surface of the nanozeolite particles and reduce the resorption probability. However, considering the internalisation rate of the bioconjugate inside the cell, perhaps after reaching the target cells the resorption process will play an important role in preventing of the release free 211Pb and 211Bi from the cells. Nanozeolite radiobioconjugates have high receptor affinity towards the NK-1 receptor expressing glioma cells in vitro, and exhibits properties suitable for the treatment of glioma cancer cells by intratumoural or post-resection injection [65,66]. Intravenous injection of the 223RaA–silane–PEG–SP(5–11) radiobioconjugate for glioma treatment is excluded due to its relatively large size and high hydrophilicity preventing it from crossing the blood–brain barrier.
After successful attempts to immobilise 225Ac inside LnPO4 nanoparticles [82,83,84,85], a team from the University of Missouri explored using LnPO4 nanoparticles to encapsulate two radium radionuclides: 223Ra and 225Ra [67]. Ac3+ can co-crystalise with lanthanum phosphate to form one phase of (Ac)LnPO4. However, Ra2+ cations most likely form a phosphate mixture, and we observe a weaker binding of 223,225Ra compared to 225Ac. The 223Ra-labelled LnPO4 retained ~88% of the 223Ra activity over a period of 35 days. For reference, (225Ac)LnPO4 retained >99% over 21 days [109]. Covering of (223Ra)LnPO4 core by two cold LaPO4 shells reduced the release of 223Ra and its daughter, 211Pb, to 0.1% over 27 days.
Carbon nanostructures are also being studied as carriers of 223Ra. In 2020, Kazakov et al. [70] tested various materials as carriers for selected radionuclides, including 226Ra radionuclide, as surrogate of 223Ra. They reported that Ra2+ sorption occurred only on reduced graphite oxide samples, and reaching 60% activity. However, in PBS and PBS + BSA buffers the desorption went from 35 to 70% in 30 min. Recently, the same team presented the results of 223Ra sorption on nanodiamond modified with derivatives of amino acids and EDTA [71]. Unfortunately, the maximum labelling was only 10% at the nanodiamond concentration 380 µg mL−1.

3.2. Actinum-225

The half-life of 9.92 d and decay energy of ~28 MeV (emission of 4 α-particles) makes the 225Ac radionuclide an attractive isotope in targeted α therapy. The half-life of 225Ac allows for long-distance distribution, as well as comfortable labelling and administration. Another important advantage of 225Ac in nuclear medicine is the emission of 440 keV γ-ray after the decay of the daughter radionuclide—213Bi. This can be used for imaging to determine the biodistribution of the radiopharmaceutical in the body. However, similar to 223Ra, the release of α-particle-emitting decay products from 225Ac-chelate complexes with high recoil energy can cause severe toxic effects in healthy organs and tissues. The use of nanoparticles seems to be very promising therapeutic strategy to entrap potentially radiotoxic daughter nuclides at the tumour site.
Sofou et al. developed pegylated liposomes with different membrane charges (zwitterionic and cationic) and sizes to encapsulate 225Ac and daughter nuclides [72]. Zwitterionic liposomes retained more than 88% of 225Ac for over 30 days, whereas cationic liposomes only retained 54%. Furthermore, the liposome size is critical in determining daughter nuclide retention. Based on theoretical calculations, the authors suggested that large-sized liposomes, ~650 nm, are required to yield >50% retention of 213Bi. However, these predictions did not consider the experiments where 225Ac localisation at the phospholipid membrane reduced the 213Bi (last daughter) retention up to 7% after 10 days.
To overcome this problem, new complex nanoconstructs were synthesised such as multivesicular liposomes (MUVELs), which are larger liposomes that contain small lipid vesicles [75]. These MUVELs were conjugated to the monoclonal anti-HER2/neu antibody, Trastuzumab, and evaluated in vitro for targeted delivery of 225Ac to ovarian cancer cells. The results showed better retention of 213Bi (17–18% after 20 days) generated by 225Ac in MUVELs compared to the aforementioned liposome strategy [72]. Their biological studies demonstrated higher binding of radiolabelled immunoliposomes to ovarian cancer cells than nontargeted liposomes, but lower immunoreactivity than the radiolabelled antibody. Moreover, the cell uptake kinetics was slower for immunolabelled MUVELs in comparison to free antibody due to different diffusibilities of each structure. Despite some improvements, the maximum entrapment efficiency of 225Ac in multivesicular liposomes did not exceed 10% of the total initial activity leading to low specific activities of produced radioactive nanoparticles.
In order to increase the encapsulated radioactivity of 225Ac, the same authors applied higher temperature (65 °C) to label the liposome (120 nm nanoconstructs containing 1,2-dinonadecanoyl-sn-glycero-3-phosphocholine lipids and cholesterol) and entrapped DOTA [73]. At this temperature, they observed significant membrane transport of the radionuclide/ionophore complexes and chelation of 225Ac by the encapsulated DOTA occurred with minimal thermal hydrolysis of the lipids comprising the membrane. In addition, they tested the maximised radionuclide loading and retention of radioactivity in two types of ionophores (oxine and calcium ionophore A23187) and buffers (citrate and acetate) that play a significant role in enabling cations through lipid membranes. This study showed identical 225Ac loading efficacy (~55–73%) of PEGylated liposomes for both ionophores in acetate buffer. By contrast, the radioactivity loading in liposomes was extremely low due to slow kinetics of complex formation between 225Ac and the encapsulated DOTA in the presence of citrate buffer. The main fraction release of the entrapped radioactive contents in PBS and 10% serum supplemented media at 37 °C occurred after the first two hours and was stably retained for 30 d. The highest retention of 225Ac was observed in PBS (~81 ± 7% of the initially encapsulated radioactivity). The conjugation of Trastuzumab allowed specifically targeting liposomes to SKOV-3-NMP2 cells expressing HER2 receptors. Unfortunately, these immunoliposomes exhibited lower binding capacity in comparison to radiolabelled Trastuzumab.
The next article of liposome encapsulation of 225Ac was published by Bandekar et al. in 2014 [8]. The authors developed 225Ac-labelled liposomes for targeted antivascular radiotherapy of prostate cancer. High radioactive 225Ac liposomes were conjugated to the J591 monoclonal antibody and the A10 aptamers that recognise the extracellular domain of prostate-specific membrane antigen (PSMA) protein. The loading efficiency ranged from 58% to 86%, which was dependent on the terminated PEG chains. In vitro biological studies showed that radioactive J591-liposomes and 225Ac labelled J591antibody exhibit similar cytotoxicity. These radiobioconjugates were more cytotoxic than A10 aptamer-labelled liposomes. Both J591- and A10-labelled liposomes were internalised similarly, ranging between 25 and 36%. These studies demonstrated the dominance of anti-PSMA liposomes loaded with 225Ac over both A10-labelled liposomes and nontargeted nanoparticles, signifying the potential of this radiobioconjugate in antivascular α-radiotherapy. The usefulness of liposomes as carriers was also studied by Henriksen et al. [16], where 120 nm liposomes that were loaded with 228Ac showed good labelling yield (~61%) and 95% stability in serum after 24 h. Polymersomes are spherical vesicles consisted of polymers and have been proposed for 225Ac encapsulation (Figure 3). These nanocarriers are similar to liposomes (Figure 2). However, they consist of amphiphilic block copolymers whilst liposomes are composed of lipid layers. The membrane thickness of polymersomes can vary between 5 and 50 nm, whereas liposomes are thinner (3–4 nm) [109]. In addition, polymersomes exhibit greater stability and less permeability, suggesting they may prevent or at least reduce the release of the recoiling daughter nuclides [79].
The first attempt using polymersomes to encapsulate 225Ac focused on the theoretical retention recoiling atoms using Monte Carlo simulations. The results showed that the size and number of block copolymer layers play an important role in the retention of recoiling atoms. Double-layered polymersomes with 400 nm retained 47% of 213Bi recoils, whereas retention of single layer polymersomes retained 10–20% lower. Increasing the nanovesicle diameter to 800 nm completely retained 211Fr and 80% of the third radionuclide daughter 213Bi. Based on Monte Carlo calculations, larger double-layered polymersomes are the most effective in 225Ac encapsulation and the retention of 211Fr and 213Bi recoil daughter nuclides. Furthermore, theoretical computer simulations were verified experimentally [110]. The encapsulation of 225Ac radionuclide into polymersomes and its transport through the hydrophobic bilayer into the aqueous cavity was achieved by applying calcium ionophore (A23187) and tropolone. The maximum encapsulation efficiency of 225Ac in combination with calcium ionophore capped at 68% and was similar to the previous results presented for lyposomes using the same method [16]. The percentage of retention for 211Fr and 213Bi in 800 nm vesicles was 69.7 ± 1.5% and 53.7 ± 4%, respectively. These vesicles exhibited the same tendency. However, the retention for 211Fr was lower in comparison to theoretical simulations [79]. In the biological studies of 100 nm polymersomes, the internalisation in HeLa cell lines showed that these nanocarriers highly accumulate around the cell’s nucleus, suggesting endocytosis as an actively transporting these molecules that may contribute to the reduction of escaping daughter nuclides.
To improve the recoil retention, 225Ac was coprecipitated with InPO4 to form small metal-phosphate nanoparticles inside polymersomes [77]. The loading efficiency of 225Ac in these nanoparticles was 90%. The results showed 20% higher retention of the first daughter nuclide. Improvement with 211Fr was seen in all size vesicles, whereas 213Bi retention was improved by 10% in smaller polymersomer sizes in comparison to previous studies with 225Ac-DTPA-containing polymersomes. The 100 nm nanocarriers exhibited better retention of both α-emitting daughter nuclides (~20% higher) in comparison to polymersomes without indium phosphate carrier [110]. Next, the therapeutic potential of 225Ac incorporated polymersomes in U87 spheroids mimicking tumour was determined [76]. The distribution of radioactive nanocarriers in 3D cell culture was not rapid. It took 4 days to spread throughout the spheroid, whereas after 7 days the distribution of polymersomes was scarcely homogenous. Distribution of nanocarriers in the tumour not only depends on the size of administered vesicles, but also on the shape, surface charge, solubility as well as the type of the cancer [111]. The cytotoxicity studies performed on 3D cell culture using 1 kBq of 225Ac-labelled polymersomes showed significant reduction in spheroid size. Interestingly, only 0.1 kBq of activity was sufficient to inhibit spheroid growth, indicating that very low activity of 225Ac can inhibit tumour growth. These results demonstrated the high potential of 225Ac vesicles in future therapy of glioblastoma. Another publication compared the recoil retention of 225Ac DTPA complexes and 225Ac coprecipitated with InPO4 immobilised in 100 nm polymersomes [78]. In in vivo experiments, synthesised 225Ac-nanocarriers were injected intravenously through the tail vein in healthy mice and intratumourally in xenografted mice bearing MDA-MB-231 well-vascularised tumours. After 4 h of the intravenous injection, the recoil retention studies showed that more 213Bi was retained in the blood and kidney for metal-phosphate nanoconstructs containing polymersomes than for DTPA nanoparticles, showing the potency of nanocarriers with InPO4. Spleen retention was similar for both nanostructures. High 213Bi activity was found in the tumour tissue where 225Ac-containing DTPA polymersomes were injected intratumourally. 213Bi was also completely retained in the tumour when entrapped in 200 nm vesicles. Biodistribution studies in the mouse models showed high uptake of polymersomes in the cancer tissue, whereas 225Ac-DOTA complex (used as a control) was rapidly excreted through the kidneys. Additionally, 225Ac-polymersomes and 225Ac-DOTA significantly inhibited tumour growth and caused γ-H2AX foci (double-stranded breaks), indicating the effectiveness of this α radionuclide therapy. Nevertheless, before 225Ac-polymersomes can be used for metastatic cancer therapy, additional studies are needed to find the appropriate size of nanocarriers, circulation time in the blood, retention of daughter nuclides at the therapy site and tumour uptake.
An interesting paper published by Sempkowski et al. [90] investigated sticky lipid nanoparticles loaded with 225Ac radionuclide. These vesicles clustered with HER2-targeting peptides on their surface exhibited high reactivity to breast cancer cells with a low expression of receptors like MDA-MB-231 or MCF7 compared to uniformly functionalised nanoparticles (nanoparticles with PEG and HER2 targeting monoclonal antibody). Furthermore, sticky 225Ac vesicles caused the death of 42–61% of MDA-MB-231 and MCF7 cancer cells at the extracellular pH of 6.5. Uniform functionalised nanoparticles did not affect cell viability. Upon entering the cell, the sticky vesicles localise fast to the perinuclear area, improving the toxic efficacy of 225Ac and the retention of recoiled atoms. These data show that nanoparticles could be a potent therapy for breast cancer patients with low HER2 receptor density, particularly because cytotoxicity for normal breast cells was not observed.
Looking for other nanocarriers for 225Ac immobilisation, Mwakisege et al. applied carbon-based nanostructures, fullerenes, to encapsulate 225Ac radionuclides because of their chemical and thermodynamic stability [80]. 225Ac was entrapped in fullerenes by direct current (DC) arc discharge-catcher method in a He atmosphere. After coupling the fullerene surface with organic adducts, only ~45% of original activity was retained in the cage, and unfortunately, 221Fr leakage was also observed. Next, the synthesis of 225Ac metallofullerenes and its electronic properties were studied by radiochromatography [81]. 225Ac@C82 (85 π electrons) species were suggested as the best potential candidates as determined by high-performance liquid chromatography (HPLC). The authors did not perform any retention studies.
The first use of carbon nanotubes as carriers for 225Ac was proposed by Ruggiero et al. [88]. Single-wall carbon nanotubes (SWCNT) appended with amines were conjugated to a DOTA ligand and a neovascular-targeting antibody E4G10 that has an affinity for the monomeric vascular endothelial-cadherin (VE-cad) epitope presented in tumour angiogenic vessels. In vivo studies using mice bearing LS174T (colon) adenocarcinomas were treated with radiolabelled high specific activity (SA) SWCNT constructs (851 GBq/g of nanoparticles) and showed two times higher survival and significant tumour regression in comparison to control groups (treated with saline and low SA analogue—1.9 GBq/g) These results demonstrated high specificity of the targeting E4G10 antibody when in combination with α particles and the unique properties of SWCNT, such as rapid renal excretion, enzymatic degradation and nontoxicity. Further studies presented a two-step approach to target tumours with the use of carbon nanotubes [87]. The external sidewall of SWCNTs (350 nm in length and with a diameter of ~1.2 nm) with primary amines was modified with the bifunctional chelator DOTA, attached to a morpholino oligonucleotide complementary to a functionalised antibody (cMORF), and then finally labelled with 225Ac. The in vitro results of studies showed that SWNT-cMORF self-assembled onto cancer cells with high specificity. The in vivo mice experiments revealed that multistep therapy, with the use of mAb-MORF followed by SWNT-cMORF-(225Ac)DOTA, was very effective and caused complete elimination of the lymphoma tumour. In addition, the rapid clearance of SWNT-cMORF-(225Ac)DOTA construct significantly reduced the toxicity five- to tenfold in mice compared to labelled mAb, free 225Ac and mAb labelled with MORF. The two-step targeted alpha therapy was found to be feasible and effective.
The application of carbon nanotubes in 225Ac3+ encapsulation was demonstrated by Matson et al. [89]. These studies were performed with the use of modified SWNT (sidewall defects) known as ultrashort tubes (US-tubes). These nanoparticles were loaded with 225Ac3+. Three different loading methods were examined: 225Ac alone, 225Ac3+ loaded with Gd3+ ions and 225Ac3+ loaded after the Gd3+ ions. The results demonstrated that after 225Ac3+ loading alone, 95% of the initial activity remained in US-tubes. Simultaneous and sequential adding of ions caused 50% encapsulation of α radionuclide due to the competition of large excess of Gd3+. Challenge experiments in human serum showed that only 40% of single-loaded 225Ac3+ was present inside the nanoparticles. Simultaneous and sequential techniques retained 77% and 80% of this radionuclide, respectively. Unfortunately, for all these 225Ac3+ studies, retention of 225Ac decay products was not determined. These authors provided a novel method to encapsulate two ions which can be used in imaging by MR and α therapy. However, further studies are needed to optimise the process and modify the nanotubes as carriers for targeted treatment.
The synthesis of 225Ac-labelled lanthanum phosphate nanoparticles was the next interesting idea to encapsulate 225Ac and retain the recoil atoms [84]. Due to the chemical similarity between Ac3+ and La3+ cations, inorganic nanoparticles with a diameter of 3–5 nm were successfully doped with 225Ac (radiochemical yield of ~47%), forming La(225Ac)PO4. The results showed that ~50% of 221Fr and 213Bi radionuclides released from the nanoparticles within 6 d and radioactivity remained constant for over 30 d. In the next step, nanoparticles were functionalised with α-hydroxy acid and conjugated to the monoclonal antibody, 201B, targeting thrombomodulin receptors highly expressed in the lung endothelium. Biodistribution and micro SPECT/CT imaging performed on female BALB/c mice revealed rapid and specific uptake of La-(225Ac)PO4 NPs-mAb in the lungs after intravenous injection. The same in vivo experiments demonstrated that after 1 h post-injection, ~50% of the recoil daughter nuclides were retained at the target site. In contrast to in vitro studies, where 50% of 213Bi remained constant for many days, the loss of 213Bi decreased significantly down to 10% within 5 d after binding to mouse lung tissue. The likely reason for such low retention of 213Bi in the mouse body is because of the entrapment of radionuclide by endothelial cells lining the lung capillaries. This causes less diffusion through the tissue. As was expected, non-targeted lanthanum phosphate nanoparticles accumulated mostly in the liver and spleen due to the recognition by the reticuloendothelial (RE) system. Even the results seem to be significantly improved in comparison to conventional approaches based on the DOTA chelator, still ~50% of recoil atoms present a risk in future therapy with 225Ac.
A new approach using gold-coated lanthanide phosphate nanoparticles was proposed to reduce the toxicity of escaped daughter nuclides [85]. 225Ac was loaded in the {La0.5Gd0.5}PO4 core. A GdPO4 layer was used to improve the daughter’s nuclide retention (Figure 4). The Au shell facilitated mAb 201b antibody conjugation reduced the possible Gd toxicity for future in vivo applications. Two layers of GdPO4 significantly increased the retention of 221Fr from 50% [84] to 70%. Adding four shells retained 98% of 221Fr decay daughter, which decreased over one week and stabilised at 88%. Moreover, the multilayered particles retained 99.99% of the 225Ac parent radionuclide, over a 3 week period. The composition and type of core also influenced the retention ability. Increased retention of 225Ac is higher in nanoparticles in which part of the core contained La [83]. Biodistribution studies showed high binding affinity of {La0.5Gd0.5}(225Ac)PO4@GdPO4@Au-mAb-201b NPs to the lungs, and were in agreement with SPECT/CT data. Retention of 213Bi in the nanoparticles was ~69% in lung tissue after 1 h post-injection, decreasing to 84% after 24 h. A similar trend observed for the spleen and liver. It is worth noting that only 2.8% and 1.5% of the 213Bi from the injected dose was observed in the kidneys after 1 and 24 h, respectively.
To minimise the effect derived from phagocytic cells as well as to improve lung-specific uptake, clodronate liposomes were injected into mice 24 h before the injection of mAb-201b conjugated nanoparticles [82]. The results showed ~50% greater accumulation of nanoparticles in the lung compared to mice that were not pretreated with clodronate. Additionally, the retention of 213Bi daughters in the lung decreased with time from 70% for 1 h to 91% for 24 h, which was higher than presented previously [83]. In vivo experiments performed on EMT-6 tumour cells demonstrated high cytotoxicity for nanoparticles linked to mAb-201b, leading to significant decrease in lung colonies compared to untargeted nanoparticles. 225Ac multilayered nanoconstructs seem to be the most improved technology, and they are very promising for the future alpha therapy.
TiO2 nanoparticles used as carriers for 225Ac were proposed by Cędrowska et al. [86]. The nanoparticles were labelled with 225Ac with a high yield of 99.8% and functionalised with silane-PEG-SP(5-11) conjugates (average 80 molecules per one nanoparticle) to target NK1 receptors overexpressed in gliomas. 225Ac3+ cations were adsorbed on the surface of TiO2 nanoparticles through an ion-exchange reaction with hydroxyl groups. Retention of 225Ac and 221Fr in PBS or NaCl solutions after 4 days ranged from 95 to 98%. In cerebrospinal fluid (CSF), the retention decreased to ~70%. In glioblastoma cancer cells, the cytotoxicity of 225Ac-TiO2-silane -PEG-SP(5-11) radiobioconjugate was higher in comparison to non-targeted or non-radioactive nanoparticles, showing the potential of nanoconstructs in TAT of brain tumours.
In another work proposed by Salvanou et al., 225Ac radionuclide was conjugated to 2–3 nm gold nanoparticles via a DOTA-derivative chelator (TADOTAGA) for local radiation treatment of cancer (Figure 5) [52]. Radiochemical yield assessed by ITLC was ~86%. In vitro experiments performed on U87 glioblastoma cancer cells showed significant cytotoxicity of 225Ac-Au@TADOTAGA radioconjugate. After 48 h of treatment with 0.5 kBq/mL synthesised nanoparticles, less than 30% of cells were detected as viable. Intravenous injection of 225Ac-Au@TADOTAGA in U87 MG tumour-bearing SCID mice caused predominate accumulation in the kidneys, liver and spleen. Biodistribution studies after intratumoural injection showed high tumour uptake at 2 h post-injection (60.67 ± 3.87% IA/g), which slowly decreased over time. The preliminary therapeutic efficacy studies performed over a period of 22 days revealed tumour growth retardation upon intratumoural injection of 225Ac-Au@TADOTAGA in comparison to mice injected with normal saline. These results are very promising. However, further preclinical evaluations are needed to use this radioconjugate as an injectable radiopharmaceutical for local radiation treatment of tumour.
Gadolinium vanadate nanoparticles have been proposed as a carrier for two α-emitters: 225Ac and 227Th. Gonzalez et al. explored multifunctional gadolinium vanadate core–shell nanoparticles doped with europium ions and 225Ac radionuclide [92]. Gd0.8Eu0.2VO4 core–shell nanoparticles were successfully obtained by precipitating Ln3+ and VO 4 3 ions, using sodium citrate as a complexing agent. 225Ac and decay products retention was assessed between Gd0.8Eu0.2VO4 core and core +2 nonradioactive shells. The radiochemical yield for 225Ac in Gd0.8Eu0.2VO4 core was 58.3%, whereas the additional two Gd0.8Eu0.2VO4 nonradioactive shells on the nanoparticle core increased the yield to 94.8%. 211Fr leakage reached a maximum of ~67.6% after 28 days in dialysis, which was higher compared to nanoparticles with two added shells (45.5%) and lanthanum phosphate nanoconstructs reported previously [83,84,85]. Due to the presence of the citrate groups on the nanoparticle surface, the retention of 213Bi was greater than 211Fr, reaching <15% and ~22% leakage from core and core +2 shells, respectively. Considering the intrinsic properties like luminescence and magnetic functionalities, the radionuclide retention capabilities, the small particle size (below 10 nm) and short time required for the synthesis (<1 h), Gd0.8Eu0.2VO4 core nanoparticles show great potential for medical applications, including α therapy.
The same group proposed gadolinium vanadate nanocrystals (NCs) as carriers of 225Ac and 227Th radionuclides and contrast agents [91]. The synthesis of GdVO4 core and core +2 shell NCs with a tetragonal structure was performed using the same method as previously described [92]. The maximum leakage of 225Ac from the NCs core was ~15% and decreased to 2.4% with the addition of two nonradioactive shells. The leakage of 221Fr from Gd(225Ac)VO4 after 23 days in dialysis reached 69.5%, and was improved to 20% by adding two shells. These retention results for 221Fr in Gd(225Ac)VO4 are comparable to those using 100 nm polymersomes, which retained 37% of this radionuclide [110], whereas the retention of 225Ac and 221Fr radionuclides in Gd(225Ac)VO4/2GdVO4 NCs are close to La0.5Gd0.5PO4 core + 2 shell nanoparticles [83]. Similarly to previous studies, the leakage of 213Bi from Gd(225Ac)VO4 and Gd(225Ac)VO4/2GdVO4 was 22.5% and 19.6%, respectively [92].

3.3. Astatine-211

As mentioned before, the carbon–astatine bond is considerably weaker than the carbon–iodine bond and many 211At-labelled vectors, which exhibit excellent ex vivo stability dehalogenate in vivo [93]. Because of these problems, others solutions for 211At labelling, including nanotechnology, have been reported. Hartman et al. tested the use of 20–50 nm US-tubes as a carrier for 211At in targeted α radionuclide therapy [93]. The idea was to introduce 211At inside the nanotube. They found that a more stable conjugate is formed when 211At is present as the mixed halogen (211AtCl) than in the anionic (211At-) form. The labelling yield of US-tubes with 211AtCl is 91.3%, which is much better than labelling with 211At- (24.7%). The authors found that a small quantity of 211AtCl remained on the exterior of the US-tube sidewall defect sites or tube ends during synthesis. After a simple wash with metabisulfite, the retention of 211AtCl@US-tubes was 60.7%. The overall labelling yield of the 211AtCl@US-tubes was comparable to the other elaborated astatination methods [48,49,50,51,52]. An important feature of functionalised SWNTs as drug carriers is their rapid clearance via renal glomerular filtration, despite their large molecular weights [112]. This clearance is called fibrillar pharmacology [113] and contrasts with large protein pharmacokinetic profiles.
Kućka et al. [94] and Cędrowska et al. [95] used astatine’s affinity towards metallic silver and proposed the application of silver nanoparticles and silver impregnated TiO2 labelled with 211At for cancer therapy. The surface of nanoparticles was coated with a hydrophilic polymer, poly(ethylene oxide). In both papers, the effect of the labelling yield using different reducing and oxidising agents was studied. It was found that in reducing conditions, where astatine exists as At-, labelling was nearly 100%. Under oxidising conditions, the labelling decreased to about 50% [94]. The Ag-labelled nanoparticles were stable even in a large excess of competing chloride ions [94] and in PBS, cysteine and glutation solutions. In human blood serum, 5% of leakage was observed after 1 h, which increased to 7.8% after 14 h.
In the next three studies, Bilewicz et al. applied the discovered high affinity of astatine to the gold surface for astatination of peptide, substance P (5-11) and monoclonal antibody, Trastuzumab [96,97,114]. Based on DFT calculations of At–Au interactions, and taking in account the proposed reaction between At and gold clusters:
At + Aun + H2O → AunAt + ½H2 + OH
the authors found that the source of the gold cluster’s high affinity to At is related to the least negative oxidation–reduction potential of astatine (6 kcal/mol) in the halogen group [114]. In the experimental studies, gold nanoparticles with 5 and 15 nm diameters were modified with substance P(5-11), a peptide fragment which targets the NK1 receptors on the glioma cells [96]. The monoclonal antibody specific for HER2 receptors, Trastuzumab, was also attached [97]. The obtained bioconjugates were quantitatively labelled with 211At by chemisorption on the gold surface. The labelled bioconjugates almost retained 211At in human serum and cerebrospinal fluid at 37 °C for 24 h. Additionally, in vitro biological studies indicated that 211At-Au-PEG-substance P(5-11) radiobioconjugate exhibited a high cytotoxic effect in vitro on glioma cancer cells [96]. The performed studies on HER2-overexpressing human ovarian SKOV-3 cells indicated high internalisation of 211At-AuNP-PEG-Trastuzumab in the cell and localisation of radiobioconjugates in the perinuclear area and its high cytotoxicity [97].

3.4. Lead-212

212Bi (t1/2 = 60.5 min) is a potentially interesting α-emitting radionuclide for internal alpha therapy. Unfortunately, its short half-life often limits the application of 212Bi to situations when the tumour cells are rapidly accessible to the targeting agent. To expand the range of applications, an interesting method uses the parent radionuclide, 212Pb (t1/2 = 10.6 h), which generates in vivo 212Bi [115]. In comparison to 212Bi radiopharmaceuticals, 212Pb have a much broader applicability because the half-life of 212Pb corresponds better with the pharmacokinetics of various biomolecules. 212Pb is transformed to 212Bi trough β- decay. Due to small mass of electrons, the calculated recoil energy of the Bi nucleus is ~0.5 eV and is not sufficient to break a chemical bond, which requires ~10 eV. However, over 30% of the γ-rays emitted during 212Pb decay are internally converted. The resulting conversion cascade and Auger electrons brings 212Bi to highly ionised states, such as Bi5+ and Bi7+. Therefore, the energy required to neutralise the charge is sufficient to break chemical bonds [116]. Previous attempts to prepare a potential in vivo generator with 212Pb complexed by the DOTA chelator failed, because ~36% of Bi escaped as a result of the radioactive decay 212Pb→212Bi [115]. Because the free radiobismuth escapes the complex during the decay, toxicity emerges when unchelated 212Bi accumulates in various organs, mainly in kidneys. Thus, alternative chelators for 212Pb complexation need to be further explored. Problems with the retention of the decay product (212Bi) is much smaller than α-decayed 225Ac or 223Ra because the recoil energy after β- decay is ~2 × 105 lower than after α decay.
Henriksen et al. [99] made the first attempts to incorporate 212Pb into nanostructures. Containers with incorporated 212Pb/212Bi were prepared by ionophore-mediated loading of 212Pb into liposomes. Uptake of n.c.a. 212Pb in liposome was 65% but increased to 90% after addition of a Pb2+ carrier. At least 95% of the 212Pb and 212Bi activity was retained in the liposomes. The performed studies indicate that liposomes give high retention of 212Pb and 212Bi formed from the 212Pb β- decay.
Montafon et al. incorporated 212Pb into indium-DTPA-tagged liposomes [98]. They found that the origin of the encapsulation is related to the dynamics of the surface, which makes the membrane partially permeable. Therefore, no ligand was necessary to allow the transfer of 212Pb from the external solution to the liposome internal part. A significant improvement of labelling was observed when DTPA is present in the internal part of liposome. Forming a strong complex with Pb2+ accelerated the concentration of the 212Pb inside the liposome. In optimised conditions, the labelling yield is ~75% and can be obtained with a mean value of 2–3 lead atoms per liposome. Unfortunately, no stability tests have been performed.
An interesting solution using fullerene for 212Pb/212Bi immobilisation was proposed by Diener et al. [101]. They labelled C60 by the recoiled 212Pb from α decay of its parent, 0.15 s 216Po, generated in situ from the decay of 224Ra (t1/2 = 3.66 d). Unfortunately, the yield of incorporating212Pb into C60 was very low, about 0.1 to 0.6%. So to produce one therapeutic dose of 212Pb@C60 (100 MBq), about 50 GBq of 224Ra is needed [8]. Also C60 cannot retain inside 212Bi formed by β- decay of 212Pb.

3.5. Thorium-227

227Th belongs to the actinium series, and the 227Th decay chain contains emission five α-particles to reach stable 207Pb (Figure 1).227Th emits α particle with energy of 5.9 MeV. However, 227Th and all its daughter nuclides deposited at a target tissue produce 34 MeV of energy, the most among the α-emitters studied in cancer therapy. The daughter of 227Th is 223Ra, which is the first in class α-emitter approved for castration-resistant prostate cancer. The gadolinium vanadate nanoparticles, after tests with 225Ac, were proposed as carriers for 227Th [92]. The radiochemical yield of 227Th encapsulation in core + 2 shell nanoparticles was 81.9%, less than in the case of 225Ac. The leakage of immobilised 227Th within Gd(227Th)VO4 reached 1.6% after 12 days and decreased to less than 1.5% in the case of Gd(227Th)VO4 nanoparticle + two shells GdVO4 layers. The first decay daughter (223Ra) retention was increased after the addition of two nonradioactive GdVO4 shells from 61% to 75%. The labelling yield of GdVO4 was lower compared to DOTA or HOPO radiobioconjugates, but retention of 223Ra was drastically higher [117,118]. The ability of GdVO4 core–shell nanoparticles to retain radionuclides gives them the potential to increase specific activity and the possibility to functionalise to make them suitable for targeted therapy because of proton relaxivity for using magnetic resonance imaging.

4. Conclusions

Radionuclide delivery systems using nanoparticles have great potential in the field of nuclear medicine. Today, immobilisation of 223Ra in inorganic nanoparticles seems to be the only possibility of the use 223Ra in targeted alpha therapy. Similarly, recently proposed targeted nanobrachytherapy using nanoparticles labelled with α-emitters has great potential for the treatment of small tumours and tumour metastases. However, an attempt for the immobilisation in nanostructure radionuclides, such as 225Ac and 212Pb, that can be complexed by “classical” chelators does not seem realistic. Recently, clinical trials for cell internalising radiobioconjugates such as 225Ac-PSMA-617 or 225Ac-DOTA-octreotide, where 225Ac was chelated by DOTA ligand, showed their exceptional therapeutic efficacy, while the toxic effects induced by the release of decay 213Bi were negligible. Therefore, there is no need in this case to use nanotechnology to immobilise 225Ac and its decay products, as we know nanoparticles after injection, before they reach the tumour, usually accumulate in such critical organs as spleen, liver or lungs causing damage of these organs. This problem remains to be solved.

Author Contributions

This review article is provided by following authors contributions: conceptualisation, A.M.-P. and A.B.; writing, A.M.-P., W.G., K.Ż.-M., K.W. and A.B.; visualisation, A.M.-P., W.G., K.Ż.-M., K.W. and A.B.; supervision, A.M.-P. and A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by two grants from National Science Centre, SONATA 2018/31/D/ST4/01488 and OPUS 2016/21/B/ST4/02133. The contribution of PhD students Kinga Żelechowska-Matysiak and Kamil Wawrowicz was realised within Project No POWR.03.02.00-00-I009/17-00 (Operational Project Knowledge Education Development 2014-2020 co-financed by European Social Fund).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Baskar, R.; Lee, K.A.; Yeo, R.; Yeoh, K.-W. Cancer and radiation therapy: Current advances and future directions. Radiother. Oncol. 2014, 112, 140–144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Iagaru, A.; Mittra, E.S.; Ganjoo, K.; Knox, S.J.; Goris, M.L. 131I-Tositumomab (Bexxar®) vs. 90Y-Ibritumomab (Zevalin®) therapy of low-grade refractory/relapsed Non-Hodgkin Lymphoma. Mol. Imag. Biol. 2010, 12, 198–203. [Google Scholar] [CrossRef] [PubMed]
  3. Yu, Z.; Ananias, H.J.K.; Carlucci, G.; Hoving, H.D.; Helfrich, W.; Dierckx, R.A.J.O.; Wang, F.; de Jong, I.J.; Elsinga, P.H. An update of radiolabeled bombesin analogs for gastrin releasing peptide receptor targeting. Curr. Pharm. 2013, 13, 3329–3334. [Google Scholar] [CrossRef]
  4. Hennrich, U.; Kopka, K. Lutathera®: The First FDA- and EMA-approved radiopharmaceutical for peptide receptor radionuclide therapy. Pharmaceuticals 2019, 12, 114. [Google Scholar] [CrossRef] [Green Version]
  5. Zoller, F.; Eisenhut, M.; Haberkorn, U.; Mier, W. Endoradiotherapy in cancer treatment—Basic concepts and future trends. Eur. J. Pharm. 2009, 625, 55–62. [Google Scholar] [CrossRef]
  6. Baidoo, K.E.; Yong, K.; Brechbiel, M.W. Molecular pathways: Targeted alpha-particle radiation therapy. Clin. Cancer Res. 2013, 19, 530–537. [Google Scholar] [CrossRef] [Green Version]
  7. Zalutsky, M.R. Radionuclide therapy. In Radiochemistry and Radiopharmaceutical Chemistry in Life Sciences; Rösch, F., Ed.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2003; pp. 315–348. [Google Scholar]
  8. 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 α-particle therapy of cancer. J. Nucl. Med. 2014, 55, 107–114. [Google Scholar] [CrossRef] [Green Version]
  9. Clara, J.A.; Monge, C.; Yang, Y.Z.; Takebe, N. Targeting signalling pathways and the immune microenvironment of cancer stem cells—A clinical update. Nat. Rev. Clin. Oncol. 2020, 17, 204–232. [Google Scholar] [CrossRef]
  10. Terraneo, N.; Jacob, F.; Dubrovska, A.; Grünberg, J. Novel therapeutic strategies for ovarian cancer stem cells. Front. Oncol. 2020, 10, 319. [Google Scholar] [CrossRef] [Green Version]
  11. Wicha, M.S. Cancer stem cells and metastasis: Lethal seeds—Commentary. Clin. Cancer Res. 2006, 12, 5606–5607. [Google Scholar] [CrossRef] [Green Version]
  12. Al-Ejeh, F.; Smart, C.E.; Morrison, B.J.; Chenevix-Trench, G.; Lopez, J.A.; Lakhani, S.R.; Brown, M.P.; Khanna, K.K. Breast cancer stem cells: Treatment resistance and therapeutic opportunities. Carcinogenesis 2011, 32, 650–658. [Google Scholar] [CrossRef] [Green Version]
  13. Al Hajj, M.; Becker, M.W.; Wichal, M.; Weissman, I.; Clarke, M.F. Therapeutic implications of cancer stem cells. Curr. Opin. Genet. Dev. 2004, 14, 43–47. [Google Scholar] [CrossRef] [PubMed]
  14. Ceder, J.; Elgqvist, J. Targeting prostate cancer stem cells with alpha-particle therapy. Front. Oncol. 2017, 6, 273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Majkowska-Pilip, A.; Bruchertseifer, F.; Weis, M.; Bonelli, M.; Laurenza, M.; Apostolidis, C.; Morgenstern, A. In vitro evaluation of 225Ac-DOTA-substance P for targeted alpha therapy of glioblastoma multiforme. Chem. Biol. Drug Des. 2018, 92, 1344–1356. [Google Scholar] [CrossRef] [PubMed]
  16. Henriksen, G.; Schoultz, B.W.; Michaelsen, T.E.; Bruland, Ø.S.; Larsen, R.H. Sterically stabilized liposomes as a carrier for α-emitting radium and actinium radionuclides. Nucl. Med. Biol. 2004, 31, 441–449. [Google Scholar] [CrossRef] [PubMed]
  17. NRC licensing decision. J. Nucl. Med. 2013, 54, 19–20N.
  18. 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]
  19. McDevitt, M.R.; Sgouros, G.; Sofou, S. Targeted and nontargeted α-particle therapies. Annu. Rev. Biomed. Eng. 2018, 20, 73–93. [Google Scholar] [CrossRef]
  20. Parker, C.; Lewington, V.; Shore, N.; Kratochwil, C.; Levy, M.; Lindén, O.; Noordzij, W.; Park, J.; Saad, F. Targeted Alpha Therapy, an Emerging Class of Cancer Agents—A Review. JAMA Oncol. 2018. [Google Scholar] [CrossRef]
  21. Ferrier, M.G.; Radchenko, V.; Wilbur, D.S. Radiochemical aspects of alpha emitting radionuclides for medical application. Radiochim. Acta 2019, 107, 1065–1085. [Google Scholar] [CrossRef]
  22. Jeon, J. Review of therapeutic applications of radiolabeled functional nanomaterials. Int. J. Mol. Sci. 2019, 20, 2323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Kleynhansa, J.; Groblera, A.F.; Ebenhan, T.; Sathekge, M.M.; Zeevaart, J.-R. Radiopharmaceutical enhancement by drug delivery systems: A review. J. Control. Release 2018, 287, 177–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Farzin, L.; Sheibani, S.; Moassesi, M.E.; Shamsipur, M. An overview of nanoscale radionuclides and radiolabeled nanomaterials commonly used for nuclear molecular imaging and therapeutic functions. J. Biomed. Mater. Res. Part. A 2019, 107A, 251–285. [Google Scholar] [CrossRef] [Green Version]
  25. Kozempel, J.; Mokhodoeva, O.; Vlk, M. Progress in targeted alpha-particle therapy. What we learned about recoils release from in vivo generators. Molecules 2018, 23, 581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. 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] [Green Version]
  27. 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-225-HuM195, in nonhuman primates. J. Nucl. Med. 2004, 45, 129–137. [Google Scholar]
  28. 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]
  29. 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]
  30. Essler, M.; Gärtner, F.C.; Neff, F.; Blechert, B.; Senekowitsch-Schmidtke, R. Bruchertseifer, F.; Morgenstern, A.; Seidl, C. Therapeutic efficacy and toxicity of 225Ac-labelled vs. 223Bi-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]
  31. Miederer, M.; Henriksen, G.; Alke, A.; Mossbrugger, I.; Quintanilla-Martinez, L.; Senekowitsch-Schmidtke, R.; Essler, M. Preclinical evaluation of the alpha-particle generator nuclide 225Ac for somatostatin receptor radiotherapy of neuroendocrine tumors. Clin. Cancer Res. 2008, 14, 3555–3561. [Google Scholar] [CrossRef] [Green Version]
  32. Pruszynski, M.; D’Huyvetter, M.; Bruchertseifer, F.; Morgenstern, A.; Lahoutte, T. Evaluation of an anti-HER2 nanobody labeled with 225Ac for targeted α-particle therapy of cancer. Mol. Pharm. 2018, 15, 1457–1466. [Google Scholar] [CrossRef] [PubMed]
  33. Cordier, D.; Forrer, F.; Bruchertseifer, F.; Morgenstern, A.; Apostolidis, C.; Good, S.; Müller-Brand, J.; Mäcke, 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] [Green Version]
  34. Krolicki, L.; Bruchertseifer, F.; Kunikowska, J.; Koziara, H.; Królicki, B.; Jakuciński, M.; Pawlak, D.; Apostolidis, C.; Mirzadeh, S.; Rola, R.; et al. Prolonged survival in secondary glioblastoma following local injection of targeted alpha therapy with 213Bi-substance P analogue. Eur. J. Nucl. Med. Mol. Imaging 2018, 45, 1636–1644. [Google Scholar] [CrossRef] [Green Version]
  35. Autenrieth, M.E.; Horn, T.; Kurtz, F.; Nguyen, K.; Morgenstern, A.; Bruchertseifer, F.; Schwaiger, M.; Blechert, M.; Seidl, C.; Senekowitsch-Schmidtke, R.; et al. Intravesical radioimmunotherapy of carcinoma in situ of the urinary bladder after BCG failure. Urologe A 2017, 56, 40–43. [Google Scholar] [CrossRef] [PubMed]
  36. Krolicki, L.; Bruchertseifer, F.; Kunikowska, J.; Koziara, H.; Królicki, B.; Jakuciński, M.; Pawlak, D.; Apostolidis, C.; Rola, R.; Merlo, A.; et al. Targeted alpha therapy of glioblastoma multiforme: Clinical experience with 213Bi- and 225Ac-Substance P. In Proceedings of the 10th International Symposium on Targeted Alpha Therapy, Kanazawa, Japan, 30 May–1 June 2017. [Google Scholar]
  37. Holzwarth, U.; Jimenez, I.O.; Calzolai, L. A random walk approach to estimate the confinement of α-particle emitters in nanoparticles for targeted radionuclide therapy. EJNMMI Radiopharm. Chem. 2018, 3, 9. [Google Scholar] [CrossRef] [Green Version]
  38. Chang, C.A. Selectivity of macrocyclic aminocarboxylates for alkaline-earth metal ions and stability of their complexes. J. Chem. Soc. Dalton Trans. 1996, 11, 2347–2350. [Google Scholar] [CrossRef] [Green Version]
  39. Matyskin, A.V.; Hansson, N.L.; Brown, P.L.; Ekberg, C. Barium and radium complexation with ethylenediaminetetraacetic acid in aqueous alkaline sodium chloride media. J. Solut. Chem. 2017, 46, 1951–1969. [Google Scholar] [CrossRef] [Green Version]
  40. Henriksen, G.; Hoff, P.; Larsen, R.H. Evaluation of potential chelating agents for radium. Appl. Radiat. Isotop. 2002, 56, 667–671. [Google Scholar] [CrossRef]
  41. Gott, M.; Yang, P.; Kortz, U.; Stephan, H.; Pietzsch, H.-J.; Mamat, C. A 224Ra-labeled polyoxopalladate as a putative Radiopharmaceutical. Chem. Commun. 2019, 55, 7631–7634. [Google Scholar] [CrossRef]
  42. Zalusky, M.R.; Reardon, D.A.; Pozzi, E.; Vaidyanathan, G.; Bigner, D.D. Targeted alpha-particle radiotherapy with 211At-labeled monoclonal antibodies. Nucl. Med. Biol. 2007, 34, 779–785. [Google Scholar] [CrossRef] [Green Version]
  43. Andersson, H.; Cederkrantz, E.; Bäck, T.; Divgi, C.; Elgqvist, J.; Himmelman, J.; Horvath, G.; Jacobsson, L.; Jensen, H.; Lindegren, S.; et al. Intraperitoneal alpha-particle radioimmunotherapy of ovarian cancer patients: Pharmacokinetics and dosimetry of (211)At-MX35 F(ab′)2—A phase I study. J. Nucl. Med. 2009, 50, 1153–1160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Champion, J.; Seydou, M.; Sabatie-Gogova, A.; Renault, E.; Montavon, G.; Galland, N. Assessment of an effective quasirelativistic methodology designed to study astatine chemistry in aqueous solution. Phys. Chem. Chem. Phys. 2011, 13, 14984–14992. [Google Scholar] [CrossRef]
  45. Berei, K.; Vasaros, L. Organic Chemistry of Astatine; Hungarian Academy of Sciences: Budapest, Hungary, 1981. [Google Scholar]
  46. Zalutsky, M.R.; Reardon, D.A.; Akabani, G.; Coleman, R.E.; Friedman, A.H.; Friedman, H.S.; McLendon, R.E.; Wong, T.Z.; Bigner, D.D. Clinical experience with alpha-particle emitting 211At: Treatment of recurrent brain tumor patients with 211At-labeled chimeric antitenascin monoclonal antibody 81C6. J. Nucl. Med. 2008, 49, 30–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Hadley, S.; Wilbur, D.S.; Gray, M.A.; Atcher, R.W. Astatine-211 labelling of an antimelanoma antibody and its Fab fragment using N-succinimidyl p-[211At]astatobenzoate: Comparisons in vivo with the p-[125I]iodobenzoyl conjugate. Bioconjug. Chem. 1991, 2, 171–179. [Google Scholar] [CrossRef] [PubMed]
  48. Wilbur, D.S.; Chyan, M.K.; Hamlin, D.K.; Kegley, B.B.; Risler, R.; Pathare, P.M.; Quinn, J.; Vessella, R.L.; Foulon, C.; Zalutsky, M.R.; et al. Reagents for astatination of biomolecules: Comparison of the in vivo distribution and stability of some radioiodinated/astatinated benzamidyl and nido-carboranyl compounds. Bioconjug. Chem. 2004, 15, 203–223. [Google Scholar] [CrossRef]
  49. Yordanov, A.Y.; Deal, K.; Garmestani, K.; Kobayashi, H.; Herring, B.; Waldmann, T.A.; Brechiel, M.W. Synthesis and biodistribution study of a new 211At-calixarene complex. J. Label. Compd. Radiopharm. 2000, 4, 1219–1225. [Google Scholar] [CrossRef]
  50. Guérard, F.; Rajerison, H.; Chauvet, A.F.; Barbet, J.; Meyer, G.; Da Silva, I.; Gestin, J.F. Radiolabelling of proteins with stabilised hypervalent astatine-211: Feasability and stability. J. Nucl. Med. 2011, 52, 1486. [Google Scholar]
  51. Pruszyński, M.; Bilewicz, A.; Zalutsky, M.R. Preparation of Rh[16S4-diol]211At and Ir[16S4-diol]211At complexes as potential precursors for astatine radiopharmaceuticals. Part I: Synthesis. Bioconjug. Chem. 2008, 19, 958–965. [Google Scholar] [CrossRef] [Green Version]
  52. 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] [Green Version]
  53. Cai, Z.; Yook, S.; Lu, Y.; Bergstrom, D.; Winnik, M.A.; Pignol, J.-P.; Reilly, R.M. Local radiation treatment of HER2-positive breast cancer using trastuzumab-modified gold nanoparticles labeled with 177Lu. Pharm. Res. 2017, 34, 579–590. [Google Scholar] [CrossRef]
  54. Yook, S.; Cai, Z.; Lu, Y.; Winnik, M.A.; Pignol, J.P.; Reilly, R.M. Intratumorally injected 177Lu-labeled gold nanoparticles: Gold nanoseed brachyteraphy with application for neoadjuvant treatment of locally advancef breat cancer. J. Nucl. Med. 2016, 57, 936–942. [Google Scholar] [CrossRef] [Green Version]
  55. Chattopadhyay, N.; Fonge, H.; Cai, Z.; Scollard, D.; Lechtman, E.; Done, S.J.; Pignol, J.-P.; Reilly, R.M. Role of Antibody-Mediated Tumor Targeting and Route of administration in nanoparticle tumor accumulation in vivo. Mol. Pharmaceutics 2012, 9, 2168–2179. [Google Scholar] [CrossRef] [PubMed]
  56. Cai, Z.; Chattopadhyay, N.; Yang, K.; Kwon, Y.L.; Yook, S.; Pignol, J.-P.; Reilly, R.M. 111In-labeled trastuzumab-modified gold nanoparticles are cytotoxic in vitro to HER2-positive breast cancer cells and arrest tumor growth in vivo in athymic mice after intratumoral injection. Nucl. Med. Biol. 2016, 43, 818–826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Jonasdottir, T.J.; Fisher, D.R.; Borrebæk, J.; Bruland, Ø.S.; Larsen, R.H. First in vivo evaluation of liposome-encapsulated 223Ra as a potential alpha-particle-emitting cancer therapeutic agent. Anticancer Res. 2006, 26, 2841–2848. [Google Scholar]
  58. Jonasdottir, T.J.; Borrebæk, J.; Bruland, Ø.S.; Salberg, G.; Moe, L.; Larsen, R.H. Evaluation of Liposomal Radiochemotherapy for Treatment of Cancer. Vet. Comp. Oncol. 2005, 3, 58–59. [Google Scholar] [CrossRef]
  59. Reissig, F.; Hübner, R.; Steinbach, J.; Pietzsch, H.-J.; Mamat, C. Facile preparation of radium-doped, functionalized nanoparticles as carriers for targeted alpha therapy. Inorg. Chem. Front. 2019, 6, 1341–1349. [Google Scholar] [CrossRef]
  60. Reissig, F.; Zarschler, K.; Hübner, R.; Pietzsch, H.-J.; Kopka, K.; Mamat, C. Sub-10 nm radiolabeled barium sulfate nanoparticles as carriers for theranostic applications and targeted alpha therapy. ChemistryOpen 2020, 9, 1–10. [Google Scholar] [CrossRef]
  61. Severin, A.V.; Vasiliev, A.N.; Gopin, A.V.; Vlasova, I.E.; Chernykh, E.V. Dynamics of sorption–desorption of 223Ra therapeutic α-emitter on granulated hydroxyapatite. Radiochemistry 2019, 61, 339–346. [Google Scholar] [CrossRef]
  62. Vasiliev, A.N.; Severin, A.; Lapshina, E.; Chernykh, E.; Ermolaev, S.; Kalmykov, S. Hydroxyapatite particles as carriers for 223Ra. J. Radioanal. Nucl. Chem. 2017, 311, 1503–1509. [Google Scholar] [CrossRef]
  63. Kozempel, J.; Vlk, M.; Malkova, E.; Bajzıkova, A.; Barta, J.; Santos-Oliveira, R.; Rossi, M. A Prospective carriers of 223Ra for targeted alpha therapy. J. Radioanal. Nucl. Chem. 2015, 304, 443–447. [Google Scholar] [CrossRef]
  64. Suchankova, P.; Kukleva, E.; Stamberg, K.; Nykl, P.; Vlk, M.; Kozempel, J. Study of 223Ra uptake mechanism on hydroxyapatite and titanium dioxide nanoparticles as a function of pH. RSC Adv. 2020, 10, 3659–3666. [Google Scholar] [CrossRef] [Green Version]
  65. Piotrowska, A.; Leszczuk, E.; Bruchertseifer, F.; Morgenstern, A.; Bilewicz, A. Functionalized NaA nanozeolites labeled with 224,225Ra for targeted alpha therapy. J. Nanopart. Res. 2013, 15, 2082. [Google Scholar] [CrossRef] [Green Version]
  66. Piotrowska, A.; Męczynska-Wielgosz, S.; Majkowska-Pilip, A.; Kozmiński, P.; Wójciuk, G.; Cędrowska, E.; Bruchertseifer, F.; Morgenstern, A.; Kruszewski, M.; Bilewicz, A. Nanozeolite bioconjugates labeled with Ra-223 for targeted alpha therapy. Nucl. Med. Biol. 2017, 47, 10–18. [Google Scholar] [CrossRef] [PubMed]
  67. Rojas, J.V.; Woodward, J.D.; Chen, N.; Rondinone, A.J.; Castano, C.H.; Mirzadeh, S. Synthesis and characterization of lanthanum phosphate nanoparticles as carriers for 223Ra and 225Ra for targeted alpha therapy. Nucl. Med. Biol. 2015, 42, 614–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Mokhodoeva, O.; Vlk, M.; Malkova, E.; Kukleva, E.; Micolova, P.; Stamberg, K.; Slouf, M.; Dzhenloda, R.; Kozempel, J. Study of 223Ra uptake mechanism by Fe3O4 nanoparticles: Towards new prospective theranostic SPIONs. J. Nanopart. Res. 2016, 18, 301–313. [Google Scholar] [CrossRef]
  69. Bilewicz, A.; Cędrowska, E.; Gawęda, W.; Bruchertseifer, F.; Morgenstern, A. Barium ferrite magnetic nanoparticles labeled with Ra-223: A new potential magnetic radiobioconjugate for targeted alpha therapy. J. Label. Compd. Radiopharm. 2019, 62, 103. [Google Scholar] [CrossRef] [Green Version]
  70. Kazakov, A.G.; Garashchenko, B.L.; Yakovlev, R.Y.; Vinokurov, S.E.; Kalmykov, S.N.; Myasoedov, B.F. An experimental study of sorption/desorption of selected radionuclides on carbon nanomaterials: A quest for possible applications in future nuclear medicine. Diam. Relat. Mater. 2020, 104, 107752. [Google Scholar] [CrossRef]
  71. Garashchenko, B.L.; Dogadkin, N.N.; Borisova, N.E.; Yakovlev, R.Y. Sorption of 223Ra and 211Pb on modified nanodiamonds for potential application in radiotherapy. J. Radioanal. Nucl. Chem. 2018, 318, 2415–2423. [Google Scholar] [CrossRef]
  72. 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] [PubMed]
  73. Chang, M.Y.; Seideman, J.; Sofou, S. Enhanced loading efficiency and retention of 225Ac in rigid liposomes for potential targeted therapy of micrometastases. Bioconjug. Chem. 2008, 19, 1274–1282. [Google Scholar] [CrossRef] [PubMed]
  74. Zhu, C.; Sempkowski, M.; Holleran, T.; Linz, T.; Bertalan, T.; Josefsson, A.; Bruchertseifer, F.; Morgenstern, A.; Sofou, S. Alpha-particle radiotherapy: For large solid tumors diffusion trumps targeting. Biomaterials 2017, 130, 67–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Sofou, S.; Kappel, B.J.; Jaggi, J.S.; McDevitt, M.R.; Scheinberg, D.A.; Sgouros, G. Enhanced retention of the alpha-particle-emitting daughters of Actinium-225 by liposome carriers. Bioconjug. Chem. 2007, 18, 2061–2067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. De Kruijff, R.M.; van der Meer, A.J.G.M.; Windmeijer, C.A.A.; Kouwenberg, J.J.M.; Morgenstern, A.; Bruchertseifer, F.; Sminia, P.; Denkova, A.G. The therapeutic potential of polymersomes loaded with 225Ac evaluated in 2D and 3D in vitro glioma models. Eur. J. Pharm. Biopharm. 2018, 127, 85–91. [Google Scholar] [CrossRef] [PubMed]
  77. De Kruijff, R.M.; Drost, K.; Thijssen, L.; Morgenstern, A.; Bruchertseifer, F.; Lathouwers, D.; Wolterbeek, H.T.; Denkova, A.G. Improved 225Ac daughter retention in InPO4 containing polymersomes. Appl. Radiat. Isotop. 2017, 128, 183–189. [Google Scholar] [CrossRef] [PubMed]
  78. De Kruijff, R.M.; Raavé, R.; Kip, A.; Molkenboer-Kuenen, J.; Morgenstern, A.; Bruchertseifer, F.; Heskamp, S.; Denkova, A.G. The in vivo fate of 225Ac daughter nuclides using polymersomes as a model carrier. Sci. Rep. 2019, 9, 11671. [Google Scholar] [CrossRef] [Green Version]
  79. 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. [Google Scholar] [CrossRef]
  80. Mwakisege, J.; Mirzadeh, S.; Boll, R.A.; Schweitzer, G. Synthesis and chemical stability of actinium-fullerenes. In Proceedings of the 234th American Chemical Society National Meeting, Boston, MA, USA, 19–23 August 2007. [Google Scholar]
  81. Akiyama, K.; Haba, H.; Sueki, K.; Tsukada, K.; Asai, M.; Toyoshima, A.; Nagame, Y.; Katada, M. Metallofullerene: Toward 225Ac nanogenerator in fullerene. Chem. Lett. 2009, 38, 978–979. [Google Scholar] [CrossRef]
  82. McLaughlin, M.F.; Robertson, D.; Pevsner, P.H.; Wall, J.S.; Mirzadeh, S.; Kennel, S.J. LnPO4 nanoparticles doped with Ac-225 and sequestered daughters for targeted alpha therapy. Cancer Biother. Radiopharm. 2014, 29, 34–41. [Google Scholar] [CrossRef]
  83. McLaughlin, M.F.; Woodward, J.; Boll, R.A.; Rondinone, A.J.; Mirzadeh, S.; Robertson, J.D. Gold-coated lanthanide phosphate nanoparticles for an 225Ac in vivo alpha generator. Radiochim. Acta 2013, 101, 595–600. [Google Scholar] [CrossRef]
  84. Woodward, J.; Kennel, S.J.; Stuckey, A.; Osborne, D.; Wall, J.; Rondinone, A.J.; Standaert, R.F.; Mirzadeh, S. LaPO4 nanoparticles doped with actinium-225 that partially sequester daughter radionuclides. Bioconjug. Chem. 2011, 22, 766–776. [Google Scholar] [CrossRef] [PubMed]
  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] [Green Version]
  86. Cędrowska, E.; Pruszynski, M.; Majkowska-Pilip, A.; Męczyńska-Wielgosz, S.; Bruchertseifer, F.; Morgenstern, A.; Bilewicz, A. Functionalized TiO2 nanoparticles labelled with 225Ac for targeted alpha radionuclide therapy. J. Nanopart. Res. 2018, 20, 83. [Google Scholar] [CrossRef] [Green Version]
  87. 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] [PubMed] [Green Version]
  88. 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] [Green Version]
  89. 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] [PubMed] [Green Version]
  90. 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]
  91. González, M.T.; Dame, A.N.; Mirzadeh, S.; Rojas, J.V. Gadolinium vanadate nanocrystals as carriers of α-emitters (225Ac, 227Th) and contrast agents. J. Appl. Phys. 2019, 125, 214901. [Google Scholar] [CrossRef]
  92. González, M.T.; Copping, R.; Mirzadeh, S.; Rojas, J.V. Multifunctional GdVO4:Eu core–shell nanoparticles containing 225Ac for targeted alpha therapy and molecular imaging. J. Mater. Chem. 2018, 6, 7985. [Google Scholar] [CrossRef]
  93. Hartman, K.B.; Hamlin, D.K.; Wilbur, D.S.; Wilson, L.J. 211AtCl@US-tube nanocapsules: A new concept in radiotherapeutic-agent design. Small 2007, 3, 1496–1499. [Google Scholar] [CrossRef]
  94. Kučka, J.; Hrubý, M.; Konák, C.; Kozempel, J.; Lebeda, O. Astatination of nanoparticles containing silver as possible carriers of 211At. Appl. Radiat. Isot. 2006, 64, 201–206. [Google Scholar] [CrossRef]
  95. Cędrowska, E.; Łyczko, M.; Piotrowska, A.; Bilewicz, A.; Stolarz, A.; Trzcińska, A.; Szkliniarz, K.; Wąs, B. Silver impregnated nanoparticles of titanium dioxide as carriers for 211At. Radiochim. Acta 2016, 104, 267. [Google Scholar] [CrossRef]
  96. Dziawer, Ł.; Kozmiński, P.; Męczynska-Wielgosz, S.; Pruszyński, M.; Łyczko, M.; Wąs, B.; Celichowski, G.; Grobelny, J.; Jastrzębski, J.; Bilewicz, A. Gold nanoparticle bioconjugates labelled with 211At for targeted alpha therapy. RSC Adv. 2017, 7, 41024–41032. [Google Scholar] [CrossRef] [Green Version]
  97. Dziawer, Ł.; Majkowska-Pilip, A.; Gaweł, D.; Godlewska, M.; Pruszyński, M.; Jastrzębski, J.; Wąs, B.; Bilewicz, A. Trastuzumab-Modified Gold Nanoparticles Labeled with 211At as a Prospective Tool for Local Treatment of HER2-Positive Breast Cancer. Nanomaterials 2019, 9, 632. [Google Scholar] [CrossRef] [Green Version]
  98. Le Du, A.; Mougin-Degraef, M.; Botosoa, E.; Rauscher, A.; Faivre Chauvet, A.; Barbet, J.; Montavon, G. In vivo 212Pb/212Bi generator using indium-DTPA-tagged liposomes. Radiochim. Acta 2011, 99, 743–749. [Google Scholar] [CrossRef] [Green Version]
  99. Henriksen, G.; Schoultz, B.W.; Hoff, P.; Larsen, R.H. Potential in vivo generator for alpha-particle therapy with 212Bi: Presentation of a system to minimize escape of daughter nuclide after decay of 212Pb to 212Bi. Radiochim. Acta 2003, 91, 109–113. [Google Scholar] [CrossRef]
  100. Severin, A.V.; Orlova, M.A.; Shalamova, E.S.; Egorov, A.V.; Sirotina, M.A. Nanohydroxyapatite and its textures as potential carriers of promising short-lived lead isotopes. Russ. Chem. Bull. Int. Ed. 2019, 68, 2197–2204. [Google Scholar] [CrossRef]
  101. Diener, M.D.; Alford, J.M.; Kennel, S.J.; Mirzadeh, S. 212Pb@C60 and its water-soluble derivatives: Synthesis, stability, and suitability for radioimmunotherapy. J. Am. Chem. Soc. 2007, 129, 5131–5138. [Google Scholar] [CrossRef]
  102. Colletti, P.M. New treatment option: 223Ra chloride, the first approved unsealed α-emitting radiopharmaceutical. Clin. Nucl. Med. 2013, 38, 724–725. [Google Scholar] [CrossRef]
  103. Torchilin, V. Multifunctional nanocarriers. Adv. Drug Deliv. Rev. 2012, 64, 302–315. [Google Scholar] [CrossRef]
  104. Proffitt, R.T.; Williams, L.E.; Presant, C.A.; Tin, G.W.; Uliana, J.A.; Gamble, R.C.; Baldeschwieler, J.D. Tumor-imaging potential of liposomes loaded with In-111-NTA: Biodistribution in mice. J. Nucl. Med. 1983, 24, 45–51. [Google Scholar]
  105. Presant, C.A.; Proffitt, R.T.; Turner, A.F.; Winsor, D.; Werner, J.L.; Williams, L.E.; Kennedy, P.; Wiseman, C.; Gal, K.; McKenna, R.J.; et al. Successful imaging of human cancer with radiolabeled phospholipid vesicles (liposomes). Cancer 1988, 62, 905–911. [Google Scholar] [CrossRef]
  106. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. 1976, 32, 752. [Google Scholar] [CrossRef]
  107. Chakraborty, S.; Das, T.; Sarma, H.D.; Venkatesh, M.; Banerjee, S. Preparation and preliminary studies on 177Lu-labeled hydroxyapatite particles for possible use in the therapy of liver cancer. Nucl. Med. Biol. 2008, 35, 589–597. [Google Scholar] [CrossRef] [PubMed]
  108. Albernaz, M.D.S.; Ospina, C.A.; Rossi, A.M.; Santos-Oliveira, R. Radiolabelled nanohydroxyapatite with 99mTc: Perspectives to nanoradiopharmaceuticals construction. Artif. Cells Nanomed. Biotechnol. 2014, 42, 88–91. [Google Scholar] [CrossRef]
  109. Rideau, E.; Dimova, R.; Schwille, P.; Wurm, F.R.; Landfester, K. Liposomes and polymersomes: A comparative review towards cell mimicking. Chem. Soc. Rev. 2018, 47, 8572–8610. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, G.; de Kruijff, R.M.; Rol, A.; Thijssen, L.; Mendes, E.; Morgenstern, A.; Bruchertseifer, F.; Stuart, M.C.A.; Wolterbeek, H.T.; Denkova, A.G. Retention studies of recoiling daughter nuclides of 225Ac in polymer vesicles. Appl. Rad. Isot. 2014, 85, 45–53. [Google Scholar] [CrossRef]
  111. Barua, S.; Mitragotri, S. Challenges associated with penetration of nanoparticles across cell and tissue barriers: A review of current status and future prospects. NanoToday 2014, 9, 223–243. [Google Scholar] [CrossRef]
  112. Singh, R.; Pantarotto, D.; Lacerda, L.; Pastorin, G.; Klumpp, C.; Prato, M.; Bianco, A.; Kostarelos, K. Tissue biodistribution and blood clearance rates of intravenously administered carbon nanotube radiotracers. Proc. Natl. Acad. Sci. USA 2006, 103, 3357–3362. [Google Scholar] [CrossRef] [Green Version]
  113. Alidori, S.; Thorek, D.L.J.; Beattie, B.J.; Ulmert, D.; Almeida, B.A.; Monette, S.; Scheinberg, D.A.; McDevitt, M.R. Carbon nanotubes exhibit fibrillar pharmacology in primates. PLoS ONE 2017, 12, e0183902. [Google Scholar] [CrossRef]
  114. Ostrowski, S.; Majkowska-Pilip, A.; Bilewicz, A.; Dobrowolski, J.C. On AunAt clusters as potential astatine carriers. RSC Adv. 2017, 7, 35854–35857. [Google Scholar] [CrossRef] [Green Version]
  115. Yong, K.; Brechbiel, M.W. Towards translation of 212Pb as a clinical therapeutic; getting the lead in! Dalton Trans. 2011, 40, 6068. [Google Scholar] [CrossRef] [PubMed]
  116. Mirzadeh, S.; Kumar, K.; Gansow, O.A. The chemical fate 212Bi-DOTA formed by β-decay of 212Pb(DOTA)2- complex. Radiochim. Acta 1993, 60, 1–10. [Google Scholar] [CrossRef]
  117. Hagemann, U.B.; Wickstroem, K.; Wang, E.; Shea, A.O.; Sponheim, K.; Karlsson, J.; Bjerke, R.M.; Ryan, O.B.; Cuthbertson, A.S. In Vitro and In Vivo Efficacy of a Novel CD33-Targeted Thorium-227 Conjugate for the Treatment of Acute Myeloid Leukemia. Mol. Cancer Ther. 2016, 15, 2422–2431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Dahle, J.; Borrebaek, J.; Melhus, K.B.; Bruland, S.; Salberg, G.; Olsen, D.R.; Larsen, R.H. Initial evaluation of (227)Th-p-benzyl-DOTA-rituximab for low-dose rate alpha-particle radioimmunotherapy. Nucl. Med. Biol. 2006, 33, 271–279. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Decay chains of 225Ac and 227Th-223Ra radionuclides.
Figure 1. Decay chains of 225Ac and 227Th-223Ra radionuclides.
Nanomaterials 10 01366 g001
Figure 2. Structure of the liposome.
Figure 2. Structure of the liposome.
Nanomaterials 10 01366 g002
Figure 3. Structure of the polymersomes labelled with 225Ac.
Figure 3. Structure of the polymersomes labelled with 225Ac.
Nanomaterials 10 01366 g003
Figure 4. Gold-coated lanthanide gadolinium phosphate nanoparticle (La0.5Gd0.5PO4@4 shells GdPO4@Au) [84].
Figure 4. Gold-coated lanthanide gadolinium phosphate nanoparticle (La0.5Gd0.5PO4@4 shells GdPO4@Au) [84].
Nanomaterials 10 01366 g004
Figure 5. Schematic of an 225Ac-Au@TADOTAGA radiobioconjugate.
Figure 5. Schematic of an 225Ac-Au@TADOTAGA radiobioconjugate.
Nanomaterials 10 01366 g005
Table 1. Physical properties of α-emitting radionuclides of interest for medical applications.
Table 1. Physical properties of α-emitting radionuclides of interest for medical applications.
RadionuclideHalf-LifeEnergy of Emitted α Particles (MeV)Main Production RouteAvailability
212Bi60.6 min6.051 (25.1%)
6.090 (9.8%)
8.785 (64%)
224Ra/212Bi generatorRelatively high
213Bi45.6 min5.875 (2.2%)
8.376 (97.8%) 1
225Ac/213Bi generatorModerate
225Ac10.0 d5.732 (8.0%)
5.791 (8.6%)
5.793 (18.1%)
5.830 (50.7%)
229Th/225Ac
226Ra(p,2n)225Ac
232Th(p,spall.) 225Ac
Moderate
211At7.2 h5.870 (41.8%)
7.450 (58.2%)
natBi(α,2n)211AtModerate
223Ra11.4 d5.540 (9.0%)
5.607 (25.2%)
5.716 (51.6%)
5.747 (9.0%)
227Ac/223Ra generatorCommercially available
224Ra3.6 d5.449 (5.0%)
5.685 (94.9%)
228Th/224Ra generatorRelatively high
226Th30.7 min6.234 (22.8%)
6.337 (75.5%)
230U/226Th generatorLow
227Th18.7 d5.709 (8.3%)
5.713 (5.0%)
5.757 (20.4%)
5.978 (23.5%)
6.038 (24.2%)
227Ac/227Th generatorModerate
212Pb10.6 d6.051 (25.1%) 2
6.090 (9.8%) 2
8.785 (64%) 2
224Ra/212Pb generatorCommercially available
1 energy of α particles emitted by daughter-213Po; 2 energy of α particles emitted by daughter-212Bi.
Table 2. Functional nanomaterials labelled with alpha-emitting radionuclides.
Table 2. Functional nanomaterials labelled with alpha-emitting radionuclides.
RadionuclideNanomaterialsAttached VectorTargeting CancerReason of Using NanomaterialsRef.
223RaLiposomefolic acid and F(ab’)2(IgG1)Ovarian carcinoma cell line, OvCar-3Recoil, radionuclide immobilisation[16]
Doxorubicin-containing-liposomes (Caelyx®/Doxil®)folic acid and F(ab’)2(IgG1)-Radionuclide immobilisation[57]
Doxorubicin-containing-liposomes (Caelyx®/Doxil®)--Radionuclide immobilisation[58]
Alendronate functionalized [223Ra] barium sulphate--Radionuclide immobilisation[59]
Very small (<10 nm) alendronate functionalized [223Ra] barium sulphate--Radionuclide immobilisation[60]
Hydroxyapatite--Radionuclide immobilisation[61]
Hydroxyapatite--Radionuclide immobilisation[62]
Hydroxyapatite--Radionuclide immobilisation[63]
Hydroxyapatite--Radionuclide immobilisation[64]
TiO2--Radionuclide immobilisation[64]
Nanozeolite A--Recoil, radionuclide immobilisation[65]
Nanozeolite ASubstance P (5-11)Glioma cells with NK1 receptorRecoil, radionuclide immobilisation, nanobrachyterapy[66]
Lanthanum phosphate--Radionuclide immobilisation[67]
Superparamagnetic iron oxide nanoparticles--Radionuclide immobilisation[68]
Barium ferriteTrastuzumabBreast and ovarian cancer cells with SKOV-3 receptorRecoil, radionuclide immobilisation[69]
Reduced graphite oxide--Radionuclide immobilisation[70]
EDTA functionalized-nanodiamond--Radionuclide immobilisation[71]
225AcPegylated liposomes--Recoil[72]
Pegylated liposomePSMA J591 antibody(PSMA)–expressing cellsRecoil[8]
Pegylated liposomeTrastuzumabSKOV-3 ovarian cellsRecoil[73]
pH-tunable liposomes--Recoil[74]
Pegylated multivesicular LiposomesTrastuzumabSKOV-3 ovarian cellsRecoil[75]
Polymersome--Recoil[76]
InPO4 nanoparticles inside polymersomes--Recoil[77]
Polymerosomes--Recoil[78]
Polymerosomes--Recoil[79]
Fullerenes--Recoil[80]
Fullerenes--Recoil[81]
(225Ac,Ga0.5,La0.5) PO4@4GdPO4@AuMAb 201bEMT-6 lung tumour cellsRecoil[82]
(225Ac,Ga0.5La0.5) PO4@2-4GdPO4@Au--Recoil[83]
La (225Ac)PO4MAb 201bEMT-6 lung tumour cellsRecoil[84]
(225Ac,Ga0.5,La0.5) PO4@4GdPO4@AuMAb 201bEMT-6 lung tumour cellsRecoil[85]
TiO2Substance P (5-11)NK1 glioma receptorRecoil[86]
Carbon nanotubes-DOTALintuzumabRituximab anti-A33CD20+B-cell lymphoma C33+myelocytic leukaemia A33+ colon adenocarcinomaTargeted amplified delivery, fast clearance[87]
Carbon nanotubes-DOTATumour neovascular-targeting antibodyLS174T xeno-graft tumour modelTargeted amplified delivery, fast clearance[88]
Carbon nanotubes--Increasing specific activity, fast clearance[89]
Lipid vehicleTrastuzumabBT-474, MDA-MB-231, MCF7 breast carcinoma cellTargeting cells with low expression of HER 2[90]
DOTA gold--Nanobrachyterapy[52]
Gadolinium vanadate--Recoil,
multimodality
[91]
227ThEu3+ doped gadolinium vanadate--Recoil,
multimodality
[92]
211AtUltrashort nanotubes--Radionuclide immobilisation[93]
Silver nanoparticles--Radionuclide immobilisation[94]
Silver impregnated TiO2--Radionuclide immobilisation[95]
Gold nanoparticlesSubstance P (5-11)Glioma cells with NK1 receptorRadionuclide immobilisation[96]
Gold nanoparticlesTrastuzumabBreast and ovarian cancer cells with SKOV-3 receptorRadionuclide immobilisation[97]
212PbIndium tagged liposomes--Recoil[98]
Sterically stabilized liposomes--Recoil[99]
Hydroxyapatite--Radionuclide immobilisation[100]
Water soluble C60--Recoil, Radionuclide immobilisation[101]

Share and Cite

MDPI and ACS Style

Majkowska-Pilip, A.; Gawęda, W.; Żelechowska-Matysiak, K.; Wawrowicz, K.; Bilewicz, A. Nanoparticles in Targeted Alpha Therapy. Nanomaterials 2020, 10, 1366. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10071366

AMA Style

Majkowska-Pilip A, Gawęda W, Żelechowska-Matysiak K, Wawrowicz K, Bilewicz A. Nanoparticles in Targeted Alpha Therapy. Nanomaterials. 2020; 10(7):1366. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10071366

Chicago/Turabian Style

Majkowska-Pilip, Agnieszka, Weronika Gawęda, Kinga Żelechowska-Matysiak, Kamil Wawrowicz, and Aleksander Bilewicz. 2020. "Nanoparticles in Targeted Alpha Therapy" Nanomaterials 10, no. 7: 1366. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10071366

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