Next Article in Journal
Cellular Investigations on Mechanistic Biocompatibility of Green Synthesized Calcium Oxide Nanoparticles with Danio rerio
Previous Article in Journal
An Assessment of InP/ZnS as Potential Anti-Cancer Therapy: Quantum Dot Treatment Increases Apoptosis in HeLa Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tailoring the Radionuclide Encapsulation and Surface Chemistry of La(223Ra)VO4 Nanoparticles for Targeted Alpha Therapy

1
Radioisotope Science and Technology Division, Oak Ridge National Laboratory, Oak Ridge, TN 37830, USA
2
Nuclear Nonproliferation Division, Oak Ridge National Laboratory, Oak Ridge, TN 37830, USA
3
Center for Nanophase Materials Sciences Division, Oak Ridge National Laboratory, Oak Ridge, TN 37830, USA
*
Author to whom correspondence should be addressed.
J. Nanotheranostics 2021, 2(1), 33-50; https://doi.org/10.3390/jnt2010003
Submission received: 24 December 2020 / Revised: 18 January 2021 / Accepted: 19 January 2021 / Published: 23 January 2021

Abstract

:
The development of targeted alpha therapy (TAT) as a viable cancer treatment requires innovative solutions to challenges associated with radionuclide retention to enhance local tumor cytotoxicity and to minimize off-target effects. Nanoparticles (NPs) with high encapsulation and high retention of radionuclides have shown potential in overcoming these issues. This article shows the influence of pH on the structure of lanthanum vanadate (LaVO4) NPs and its impact on the radiochemical yield of 223Ra and subsequent retention of its decay daughters, 211Pb and 211Bi. An acidic pH (4.9) results in a high fraction of La(223Ra)VO4 NPs with tetragonal structure (44.6–66.1%) and a 223Ra radiochemical yield <40%. Adjusting the pH to 11 yields >80% of La(223Ra)VO4 NPs with monoclinic structure and increases the 223Ra radiochemical yield >85%. The leakage of decay daughters from La(223Ra)VO4 NPs (pH 11) was <5% and <0.5% when exposed to deionized water and phosphate-buffered saline, respectively. Altering the surface chemistry of La(223Ra)VO4 NPs with carboxylate and phosphate compounds resulted in a threefold decrease in hydrodynamic diameter and a 223Ra radiochemical yield between 74.7% and 99.6%. These results show the importance of tailoring the synthesis parameters and surface chemistry of LaVO4 NPs to obtain high encapsulation and retention of radionuclides.

Graphical Abstract

1. Introduction

Targeted alpha therapy (TAT) has been shown to be highly effective in treating small solid tumors and micrometastases, due to the cytotoxic potential of α-particles, which are characterized by a high linear energy transfer (80–100 keV/µm) and short-range in tissue (40–90 µm) [1,2,3]. In the clinic, current approaches to delivering α-emitting radionuclides involve radionuclide accumulation at the tumor site based on its chemical affinity and the development of radioimmunoconjugates for specific targeting of tumor cells [1,3,4]. Although the accumulation of radionuclides based on their chemical affinity for specific organs and tissues has been successful in treating cancer, this approach limits the potential of delivering therapeutic radionuclides to specific cancer subtypes [5]. Xofigo® (Ra(223Ra)Cl2), the first TAT drug approved by the US Food and Drug Administration, is a clear example of the limitations of targeting by chemical affinity because it can only be used to treat metastatic castration-resistant prostate cancer (mCRPC) that has spread to the bone [3]. Radioimmunoconjugates offer the possibility to treat different cancer subtypes by using targeting vectors that will bind to antigens overexpressed on the surface of cancer cells [1,4,6]. Multiple α-emitting radioimmunoconjugates are undergoing clinical trials as a treatment for different cancers, including mCRPC (phase 1, NCT03724747), solid tumors expressing mesothelin (phase 1, NCT03507452), and acute myeloid leukemia (phase 2, NCT03867682) [7,8,9]. These radioimmunoconjugates can readily deliver the α-emitting radionuclide to the tumor site; however, they cannot prevent leakage of decay daughters from the tumor site or their potential relocation to healthy tissues and organs, due to their bond-breaking recoil energy (100–200 keV) [4,6]. Although the relocation of decay daughters can lead to unwanted radiotoxicity to healthy organs and a decrease in therapeutic efficacy, radioimmunoconjugates with rapid uptake into cancer cells can overcome these limitations and have shown remarkable therapeutic efficacy [10,11,12].
Organic and inorganic nanoparticles (NPs) have been proposed as delivery platforms for α-emitting radionuclides to minimize the relocation of decay daughters from a tumor site. Organic NPs, including liposomes and polymersomes, prevent the relocation of radionuclides by containing all decay daughters within an aqueous volume enclosed by a hydrophobic bilayer composed of either phospholipids or block copolymers [13,14,15,16]. The retention of decay daughters within organic NPs is size-dependent, and the localization of α-emitting radionuclides at the hydrophobic bilayer can lead to poor radionuclide retention [13,14,15]. Enhanced retention of decay daughters was obtained by precipitating 225Ac-InPO4 nanocrystals within polymersomes due to the high stopping power (i.e., high-Z) environment surrounding the α-emitting radionuclide [17]. The principle of surrounding α-emitting radionuclides with a high stopping power environment has been used for developing inorganic NPs. Different compounds have been studied as delivery platforms to minimize the relocation of decay daughters, including lanthanide phosphate [18,19,20,21], lanthanide vanadate [22,23], iron oxide [24,25], titanium dioxide [26,27], zeolites [28,29], barium sulfate [30,31], and gold [32]. The retention of decay daughters within inorganic NPs is based on the stopping power of each compound, the NP density, and the fraction of recoil energy that can be transferred to the NP by translational, rotational, and vibrational modes [18,20,21,33]. Organic and inorganic NPs radiolabeled with α-emitting radionuclides have been tested in vitro to assess the retention of radionuclides and their cytotoxicity against tumor cells and tumor spheroids [14,27,28,34]. In vivo evaluation of radiolabeled NPs has focused on the biodistribution of both NPs and radionuclides, their therapeutic efficacy against different cancers, and their side effects [21,25,32,35]. Both in vitro and in vivo experiments have shown the potential application of NPs as delivery platforms of α-emitting radionuclides in clinical settings.
This work aims to develop lanthanum vanadate (LaVO4) NPs as delivery platforms for 223Ra (t1/2 = 11.4 d) and to enhance their radionuclide retention and surface chemistry for TAT. LaVO4 NPs were studied as alternatives to other lanthanide-based NPs, GdVO4, and LaPO4, because LaVO4 NPs have a simple synthesis and can be precipitated with monoclinic and tetragonal structures. The characteristic monoclinic structure of LaVO4 NPs has exhibited higher radiation resistance and greater capacity to incorporate multivalent cations than the tetragonal structure of GdVO4 NPs [36,37]. Additionally, LaVO4 NPs can be synthesized by coprecipitation in aqueous media at room temperature in approximately 30 min [38], which can be exploited for radiolabeling with short-lived diagnostic radionuclides. The encapsulation and retention of radionuclides were optimized by adjusting the synthesis parameters of La(223Ra, 140Ba)VO4 NPs, where 140Ba (t1/2 = 12.8 d), a processing byproduct, was used to assess its potential as a surrogate of 223Ra. Retention studies were performed in vitro by dialyzing the NP suspension against deionized water or phosphate-buffered saline (PBS). Altering the surface chemistry of NPs with different compounds can enhance their colloidal stability and help them evade the reticuloendothelial system (RES), while providing functional groups for conjugation with targeting vectors and fluorophores [25,31,39,40]. To tackle this, we investigated the modification of LaVO4 NP surface chemistry with different carboxylate and phosphate compounds.

2. Experimental Section

2.1. Materials and Reagents

The following reagents were purchased from Sigma-Aldrich (St. Louis, MO, USA) and were used without further purification: lanthanum(III) chloride heptahydrate (LaCl3·7H2O; ACS reagent, CAS 10025-84-0), ethylenediaminetetraacetic acid disodium salt dihydrate (EDTA; ACS/USP grade, CAS 6381-92-6), ammonium citrate dibasic (NH4-Cit; ACS reagent, 98%, CAS 3012-65-5), sodium citrate tribasic dihydrate (Na-Cit; ACS reagent, ≥ 99%, CAS 6132-04-3), sodium phosphate crystals (Hex; +80 mesh, 96%, CAS 68915-31-1). Sodium orthovanadate (Na3VO4; 99%, Acros Organics, CAS 13721-39-6) and sodium tripolyphosphate (TPP; for analysis, Acros Organics, CAS 7758-29-4) were purchased from Fisher Scientific (Waltham, MA, USA) and used without further purification. Deionized water (18 Ω·cm) was obtained from a MilliporeSigma™ Milli-Q™ Ultrapure Water System. NP suspensions were dialyzed using 10K molecular weight cutoff Slide-A-Lyzer G2 Dialysis Cassettes (ThermoFisher, catalog number 87730).
Radionuclides 223Ra and 140Ba were obtained from the chemical-processing campaign designed to recover 225Ac from a proton-irradiated 232Th target. In that process, the 232Th target was initially dissolved in concentrated HCl, and the solution was loaded onto an anion exchange resin (BIO-RAD AG® MP-1M 100–200 mesh). Thorium(IV) was eluted from the column using 10 M HCl, and a matrix conversion to 1 M citric acid was performed before the Th eluate was loaded onto a cation exchange resin (BIO-RAD AG® 50W-X8 100–200 mesh). Subsequently, 223Ra and 140Ba were eluted from the column by using 2.5 M HNO3, and the eluate was converted to 0.1 M HNO3 for the synthesis of La(223Ra, 140Ba)VO4 NPs. Carrier-free 223Ra was obtained from a 227Ac generator, where the parent radionuclide was produced via thermal neutron irradiation of a 226Ra target. The irradiated target was processed to chemically separate and purify 227Ac, which is dissolved in 0.1 M HNO3 for further handling. The radioactive 227Ac solution was loaded into a cation exchange resin (BIO-RAD AG® 50W-X4 100–200 mesh), and then 223Ra was eluted with 8 M HNO3. The 223Ra eluate was converted to 0.1 M HNO3 for La(223Ra)VO4 NP synthesis. The decay schemes of 223Ra and 140Ba are shown in Figure S1 in the electronic supplementary information (ESI).

2.2. Synthesis of La(223Ra, 140Ba)VO4, La(223Ra)VO4, and LaVO4 Nanoparticles

A modified version of the procedure reported by Huignard et al. [38] was used to synthesize the La(223Ra, 140Ba)VO4, La(223Ra)VO4, and LaVO4 NPs. La(223Ra, 140Ba)VO4 and La(223Ra)VO4 NPs were obtained by coprecipitation in aqueous media, where the radionuclides were mixed with a LaCl3 solution, and then combined with a Na3VO4 solution at a 1:1 molar ratio following different steps (Figure 1). The proposed modifications were motivated by our previous work with GdVO4 NPs [41]. For characterization purposes, nonradioactive LaVO4 NPs were prepared following the same steps described for La(223Ra, 140Ba)VO4 and La(223Ra)VO4 NPs.
In procedure A, a 0.1 M HNO3 solution containing 29 µCi of 223Ra, 21 µCi of 140Ba, and 18 µCi of 140La was evaporated to dryness in a Pyrex v-vial using an infrared heat lamp and a hot plate. A 0.5 mL LaCl3 (0.1 M) solution was added into the Pyrex v-vial, stirred for 10 min, and transferred into a microcentrifuge tube. The LaCl3/223Ra/140Ba solution was added dropwise to a 0.5 mL Na3VO4 (0.1 M, pH 12.5) solution under constant stirring, and the mixture was stirred for 30 min. Then, the La(223Ra, 140Ba)VO4 NP suspension was transferred into the dialysis cassette to evaluate the encapsulation and retention of radionuclides over time.
In procedure B, a 0.1 M HNO3 solution containing 15 µCi of 223Ra, 10 µCi of 140Ba, and 7 µCi of 140La was evaporated to dryness in a Pyrex v-vial using an infrared heat lamp and a hot plate, and the precipitate was mixed with a 0.5 mL LaCl3 (0.1 M) solution for 10 min. A 0.5 mL Na3VO4 (0.1 M, pH 12.5) solution was added dropwise under constant stirring into the Pyrex v-vial containing the LaCl3/223Ra/140Ba solution, and the mixture was stirred for 30 min before the NP suspension was transferred into the dialysis cassette.
In procedure C, a 0.1 M HNO3 solution containing 20 µCi of 223Ra, 10 µCi of 140Ba, and 7 µCi of 140La was evaporated to dryness in a Pyrex v-vial using an infrared heat lamp and a hot plate, and the precipitate was mixed with a 0.5 mL LaCl3 (0.1 M) solution for 10 min. After dropwise addition of a 0.5 mL Na3VO4 (0.1 M, pH 12.5) solution into the LaCl3/223Ra/140Ba solution, the pH of the NP suspension was adjusted to 11 using KOH (1 M). The NP suspension was stirred for 30 min and then transferred into the dialysis cassette.

2.3. Surface Modification of LaVO4 Nanoparticles

LaVO4 NPs, synthesized following procedure C, were modified with different compounds containing carboxylate (NH4-Cit, Na-Cit, EDTA) or phosphate (TPP, Hex) groups to enhance their stability in aqueous media and to allow their functionalization with targeting vectors or fluorophores. Initially, each compound was distributed within a microcentrifuge tube to yield different LaVO4:compound molar ratios (e.g., 6:1, 3:1, 1:1, and 1:2), based on a LaVO4 NP concentration of 10.9 mg/mL. The as-prepared LaVO4 NP suspension was transferred into the microcentrifuge tube containing the respective compound and was mixed vigorously using a vortex mixer for 30 min. The modified LaVO4 NPs were washed three times with deionized water (14,500 rpm, 20 min) to remove unreacted species.

2.4. Characterization of LaVO4 Nanoparticles

The LaVO4 NP precipitate was suspended in aqueous media or evaporated to dryness in a muffle furnace at 60 °C for 6 to 8 h. The NP suspensions were diluted 50 times in deionized water or PBS (1X, pH 7.2). Dynamic light scattering and phase analysis light scattering were performed on a ZetaPALS zeta potential analyzer (Brookhaven Instruments Corporation, Holtsville, NY, USA) to characterize the particle size distribution and the surface charge. The dried precipitate was ground using a mortar and pestle to obtain a powder sample that was used for X-ray diffraction, Fourier-transform infrared (FTIR), and morphological characterization. Powder X-ray diffraction was performed on an AXRD® Benchtop diffractometer (PROTO Manufacturing Inc., Taylor, MI, USA) operated at 30 kV and 20 mA with a 1500 W fine-focus Cu ceramic X-ray tube (λ = 1.5406 Å). Rietveld refinement was used to determine the crystal structure of LaVO4 NPs. Surface characterization was carried out using attenuated total reflectance FTIR spectroscopy on a Nicolet™ iS50 FTIR spectrometer with a diamond crystal (Thermo Scientific™, Waltham, MA, USA). A Hitachi 300 kV cold-field emission gun transmission electron microscope was used to characterize the morphology of NP powders that had been dispersed in isopropanol and drop-casted onto 200 mesh Lacey carbon copper grids.

2.5. Radionuclide Encapsulation within La(223Ra,140Ba)VO4 and La(223Ra)VO4 Nanoparticles

La(223Ra,140Ba)VO4 NPs were transferred into a dialysis cassette after synthesis and were dialyzed against deionized water overnight to remove unreacted species and radionuclides. A 5 mL dialysate aliquot was removed before the dialysate was replaced with deionized water or PBS (1X, pH 7.4) to assess the radionuclide encapsulation over time. Dialysis of La(223Ra)VO4 NPs, synthesized following procedure C, against PBS was performed to expose the NP suspensions to biologically relevant conditions. A 5 mL dialysate aliquot was taken periodically, every 2 to 3 days, and was assayed by γ-ray spectroscopy to determine the activity of 223Ra, 211Pb, 211Bi, 140Ba, and 140La within the dialysate. A high-purity germanium detector (Ortec, Oak Ridge, TN, USA), having a ~100 cm3 crystal active volume and a beryllium window, coupled to a PC-based multichannel analyzer (Mirion Technologies, San Ramon, CA, USA), was used for γ-ray spectroscopy. Energy and efficiency calibrations were determined by γ-ray sources traceable to the National Institute of Standards and Technology. The γ-ray energies and intensities used to calculate the activity of each radionuclide are summarized in Table S1 in the ESI. Equation (1) was used to calculate the fraction of activity found in the dialysate, referred to as leakage:
L e a k a g e ( % ) = A d a × V c A D C ( t ) × 100
where Ada is the activity measured in the dialysate aliquot, Vc is a volume correction to account for the total activity in the dialysate (i.e., Vc = Vdialysate/Vdialysate aliquot), and ADC(t) is the expected activity within the dialysis cassette corrected by radioactive decay. The 223Ra, 140Ba, and 140La radiochemical yields were calculated based on the initial activity and the activity lost during synthesis and dialysis as summarized in Equation (2):
R Y ( % ) = A 0 ( A V S + A D + A T P + A D m a x ) A 0 × 100
where A0 is the initial activity, AVS is the activity lost in the vial and stirrer, AD is the activity lost in the first dialysis, ATP is the activity lost in the transfer pipette and microcentrifuge tube, and ADmax is the maximum activity lost over time during dialysis. All activities were corrected for radioactive decay. To assess the influence of surface modification on the 223Ra radiochemical yield, La(223Ra)VO4 NP suspensions were mixed at a 1:1 (NH4–Cit, Na–Cit, Hex, or TPP) and 3:1 (EDTA) LaVO4:compound molar ratios. Modified La(223Ra)VO4 NPs were washed three times with deionized water (14,500 rpm, 20 min), and each supernatant was carefully removed with a transfer pipette and stored in a microcentrifuge tube for characterization using γ-ray spectroscopy. La(223Ra)VO4 NP suspensions were dispersed in deionized water and characterized using γ-ray spectroscopy to determine the activity of 223Ra after surface modification and cleaning.

3. Results and Discussion

3.1. Characterization of LaVO4 NPs

The procedures used to synthesize LaVO4 NPs were variations on a procedure developed for YVO4:Eu NPs [38], where the order in which LaCl3 and Na3VO4 solutions are mixed was varied (Figure 1). The modified procedures rely on the precipitation of La3+ and VO43− ions in aqueous media to form LaVO4 NPs [38]. Procedure A is similar to the one developed by Huignard et al. for YVO4:Eu NPs, with the difference that the pH of the NP suspension was not adjusted to 11 after dropwise addition of LaCl3 into Na3VO4. Since no additional reagents were added during synthesis, the pH of the LaVO4 NP suspensions was 4.9. A similar pH was obtained for LaVO4 NP suspensions prepared following procedure B, where the Na3VO4 solution was added dropwise into the LaCl3 solution. In procedure C, both LaCl3 and Na3VO4 solutions were mixed in a similar fashion as carried out in procedure B, with the difference that KOH was used to adjust the pH of the LaVO4 NP suspension to 11. It is expected that the order in which both LaCl3 and Na3VO4 are mixed can influence the encapsulation of surrogate cations and radionuclides within LaVO4 NPs [41]. A low pH can lead to the precipitation of vanadate species that are different from VO43−, whereas a pH >13 will lead to the precipitation of hydroxide species, such as La(OH)3 [38,42,43]. Overall, this synthesis procedure is attractive for radiochemical settings, given its simplicity and the fact that different dopant cations can be incorporated into LaVO4 NPs to provide luminescent and/or magnetic properties.
The diffraction patterns of as-prepared LaVO4 NPs, synthesized by the different procedures, match the reference patterns of both tetragonal (powder diffraction file: 00-032-0504, space group I41/amd) and monoclinic (powder diffraction file: 00-050-0367, space group P21/n) crystal structures (Figure 2). Both tetragonal and monoclinic structures have been reported for LaVO4 NPs [44,45]. Since the monoclinic structure is thermodynamically stable, due to its high coordination number (CN; CN = 9) and the large ionic radius of La3+ (1.216 Å for CN = 9 [46]), it is expected that LaVO4 NPs will precipitate primarily with a monoclinic structure [47,48]. Although the tetragonal structure is preferred for smaller lanthanides (i.e., Ln = Ce to Lu), the precipitation of LaVO4 having a tetragonal structure can be obtained by tailoring the synthesis method, the pH of the reaction, and the dopant concentration [49,50,51,52]. LaVO4 NPs, synthesized by procedures A and B, exhibited Bragg peaks with similar intensities, whereas additional Bragg peaks associated with the monoclinic structure were observed for LaVO4 NPs prepared by procedure C (Figure 2). The fraction of each crystal structure and its crystallite size (Table 1) were determined by Rietveld refinement and the Scherrer equation, respectively. The crystallite size was calculated from the (200) plane (i.e., 2 Theta = 23°) in the tetragonal structure and the (012) plane (i.e., 2 Theta = 30°) in the monoclinic structure. Based on Rietveld refinement, a large fraction of LaVO4 NPs prepared using procedure A were characterized by a tetragonal structure (66.1%). The precipitation of LaVO4 NPs with a tetragonal structure may have been influenced by the low pH (4.9) at which the reaction was kept for 30 min. LaVO4 NPs prepared using procedure B were characterized by a similar, yet a distinct fraction of both tetragonal and monoclinic structures (Table 1), suggesting that the order in which the reagents were mixed influences the crystal structure of LaVO4 NPs. In procedure C, Rietveld refinement revealed that a large fraction of LaVO4 NPs was characterized by a monoclinic structure (Table 1), suggesting that adjusting the pH to 11 promoted the precipitation of particles with a thermodynamically stable structure. The adjustment of the suspension’s pH can lead to changes in the interfacial tension, and hence, differences in nucleation and growth kinetics. The changes in nucleation and growth kinetics can contribute to the precipitation of particles with different crystal structure fractions [49].
Transmission electron micrographs showed that the precipitation of LaVO4 NPs had different morphologies, depending on the synthesis procedure followed (Figure 3). The aggregation of LaVO4 NPs, which was a common characteristic between the different procedures, could be promoted by the sample preparation, the lack of stabilizing species on the particle surface, or both. LaVO4 NPs prepared by procedure A were represented by particles having an anisotropic and elongated morphology, analogous to a tetragonal shape (Figure 3a,d). In Figure 3d, the lattice fringes of LaVO4 NPs correspond to an interplanar spacing of 3.2 Å and 3.6 Å associated with the (120) and (200) planes of the monoclinic and tetragonal structures, respectively. Particles having a more defined tetragonal shape were observed in LaVO4 NPs, synthesized following procedure B (Figure 3b,e). The interplanar spacing between lattice fringes corresponds to both monoclinic (d120 = 3.2 Å) and tetragonal (d200 = 3.6 Å) structures (Figure 3e). LaVO4 NPs prepared using procedure C were characterized by a small fraction of large particles having a tetragonal shape and a large fraction of aggregated anisotropic particles (Figure 3c,f). These results are consistent with the characteristic morphology obtained for GdVO4 NPs when the pH was adjusted to >11, where both large trapezoidal particles and aggregates of small crystallites were observed [41]. Based on an interplanar spacing of 3.5 Å and 3.3 Å, the lattice fringes correspond to the (120) and (200) planes of the monoclinic structure (Figure 3f).
The surface chemistry of LaVO4 NPs, synthesized by the different procedures, was characterized by similar vibration bands (Figure 4). The La–O bond is represented by a weak vibration band at 433 cm−1, whereas the strong vibration at 765 cm−1 corresponds to the V–O bond [47,48]. The bands centered at 1397 cm−1 and 1455 cm−1 are related to the symmetrical and asymmetrical vibrations of carboxylate groups, respectively [45,53]. The presence of carboxylate groups can be associated with the formation and precipitation of carbonate species during synthesis, promoted by a high reactivity between the surface hydration layers of LaVO4 NPs and the CO2 from the air [45,53]. The bands located at 1634 cm−1 and 3300 cm−1 correspond to the bending and stretching O–H vibrations from physically absorbed water molecules on the surface of LaVO4 NPs [43,54].
As shown by the transmission electron micrographs, LaVO4 NPs, synthesized by different procedures exhibited significant aggregation (Figure 3). These results were supported by the average hydrodynamic diameter obtained from the intensity distribution after the characterization of LaVO4 NP suspensions using dynamic light scattering. LaVO4 NPs prepared by different procedures had a hydrodynamic diameter >900 nm when diluted in deionized water or PBS (Table 2). Particle aggregation can be associated with hydrogen bonding between surface hydroxyl groups [40]. There was a significant difference between the zeta potential of LaVO4 NPs diluted in deionized water or PBS, depending on the synthesis procedure followed (Table 2). When LaVO4 NPs were diluted in deionized water, LaVO4 NPs, synthesized following procedures A and B, had a positive zeta potential (~17 mV), whereas LaVO4 NPs prepared by procedure C were characterized by a negative zeta potential (−26.5 mV). This difference in zeta potential could be attributed to a higher fraction of carbonate species on the surface of LaVO4 NPs prepared by procedure C, as shown from the FTIR spectrum (Figure 4). Dilution of LaVO4 NPs in PBS promoted a significant change in the magnitude of the zeta potential that depends on the synthesis procedure followed (Table 2). The change in zeta potential could be related to the adsorption or interaction of phosphate species with the LaVO4 NP surface, which can be concluded based on the vibration bands of phosphate groups observed on the FTIR spectrum (Figure S2 in the ESI). It is expected that phosphate species from PBS will have a strong interaction with OH groups and La cations on the particle surface.

3.2. Surface-Modified LaVO4 NPs

The nature of the functional groups of LaVO4 NPs synthesized by the different procedures resulted in poor colloidal stability and limits their functionalization with targeting vectors. To address that limitation, LaVO4 NPs prepared following procedure C were modified through surface adsorption and/or complexation with different compounds containing carboxylate and phosphate groups. Additional vibrational bands were observed after LaVO4 NPs were mixed with each compound at room temperature for 30 min (Figure 5 and Figures S3–S7 in the ESI). The characteristic vibrational bands of the LaVO4 NPs modified with the different compounds are summarized in Table 3. The FTIR spectra of the LaVO4 NPs modified with NH4-Cit showed that at least a 3:1 LaVO4:NH4-Cit molar ratio is required to obtain a significant difference in the transmittance of the vibrational bands and the appearance of the asymmetric carboxylate bidentate bond stretching band at 1557 cm−1 (Figure S3 in the ESI). Additionally, the total mass of LaVO4 NPs was reduced by 12% after surface functionalization with a 1:1 LaVO4:NH4-Cit, suggesting a partial dissolution of the particles.
Alternatively, when using Na-Cit to modify LaVO4 NPs, the asymmetric carboxylate bidentate bond stretching band at 1557 cm−1 was only observed at a 1:1 LaVO4:Na-Cit molar ratio (Figure S4 in the ESI). The FTIR spectra from LaVO4 NPs modified with Na-Cit showed the presence of carbonate species at different LaVO4:Na-Cit molar ratios, based on the symmetric and asymmetric vibrations of carboxylate groups [45,53]. A change of <4% in the mass of the LaVO4 NPs after surface modification with 1:1 LaVO4:Na-Cit molar ratio, suggests that Na-Cit can interact with La cations on the particle surface without causing a significant loss of mass. Modifying LaVO4 NPs with EDTA resulted in the appearance of the asymmetric carboxylate bidentate bond stretching band and its shift to 1579 cm−1 at both 6:1 and 3:1 LaVO4:EDTA molar ratios (Figure S5 in the ESI). The shift of the carboxylate bidentate bond stretching band of LaVO4 + EDTA with respect to that of LaVO4 + NH4-Cit and LaVO4 + Na-Cit is likely associated with the coordination of the carboxylate groups around the La cation [55]. At a 3:1 LaVO4:EDTA molar ratio, LaVO4 + EDTA NPs were partially dissolved based on a 44% decrease in LaVO4 NP mass during surface modification. It is assumed that higher concentrations of EDTA (i.e., 1:1 and 1:2 LaVO4:EDTA molar ratios) completely dissolved LaVO4 NPs because of a cation exchange reaction between LaVO4 NPs and EDTA, leading to the formation of La-EDTA complexes [56].
Modification of LaVO4 NPs with phosphate-containing compounds (TPP and Hex), resulted in the appearance of characteristic phosphate bending and stretching vibration bands at all molar ratios (Figures S6 and S7 in the ESI). Bending vibration bands from PO4 groups are observed at 536 cm−1 and 611 cm−1, whereas a strong stretching vibration band from PO4 groups appears at 1013 cm−1 [57,58]. The fact that the characteristic vibration bands from PO4 groups are displayed after modification with a molar ratio of 6:1 LaVO4:TPP and LaVO4:Hex suggests a strong interaction between the phosphate groups and the surface La cations. Phosphate compounds can exhibit enhanced selectivity toward La cations relative to carboxylic compounds, due to their tetrahedral phosphonic group and the polarizability of the P–O bond [59]. Due to the strong interaction between La cations and phosphate compounds [60], LaVO4 NPs were partially dissolved after surface modification with TPP and Hex (1:1 LaVO4:compound molar ratio), as shown by a decrease in mass of 15% and 55%, respectively. Using Hex as a stabilizing compound may provide La cations with a hydrophobic environment to prevent their transmetalation when testing them in vivo [61].
Surface modification with the different compounds enhanced the colloidal stability of LaVO4 NPs in deionized water based on their particle size distribution and zeta potential. The different compounds yielded a hydrodynamic diameter <300 nm (intensity distribution) and a negative zeta potential (Table 4). LaVO4 + NH4-Cit NPs had the smallest hydrodynamic diameter (144.7 ± 4.8 nm) and lowest zeta potential (−17.0 ± 1.8 mV) among the different compounds tested in deionized water. Although the zeta potential of LaVO4 NPs prepared by procedure C before and after surface modification was not significantly different, a threefold decrease in the hydrodynamic diameter was obtained after surface modification. Surface-modified LaVO4 NPs exhibited significant aggregation when diluted in PBS, resulting in a hydrodynamic diameter >1000 nm. The aggregation of LaVO4 NPs in PBS can be attributed to a compression of the electric double layer induced by the high ionic strength of PBS (1X, 162.7 mM [62]), and thus, a decrease in the magnitude of the zeta potential (Table 4) [40]. The desorption of carboxylate compounds through an exchange with phosphate species in PBS could also cause the aggregation of surface-modified LaVO4 NPs based on the presence of characteristic PO4 vibration bands on the FTIR spectra (Figure S8 in the ESI) [40]. These results suggest a weak interaction between carboxylate compounds with surface La cations, which could be detrimental for biological applications where proteins, such as albumin, can bind to La cations via a ligand exchange reaction [63]. The binding of proteins can affect the biodistribution and pharmacokinetics of NPs and lead to NPs being taken up by the RES [64].

3.3. Encapsulation of 223Ra and Decay Daughters in La(223Ra)VO4 NPs

The encapsulation of 223Ra and retention of its decay daughters (211Pb and 211Bi) were assessed in vitro by dialyzing the La(223Ra)VO4 NP suspensions against deionized water [19,22,23]. The leakage of radionuclides was calculated using Equation (1) and evaluated over time (Figure 6). La(223Ra, 140Ba)VO4 NPs prepared following procedure A exhibited the highest leakage of 223Ra, 140Ba, and their decay daughters (Figure 6a). The leakage of 223Ra reached a maximum of 25.3 ± 3.9% and then decreased to 3.9 ± 1.3% after 10 days in dialysis (Figure 6a). A similar trend was observed for 140Ba, where its leakage was ~15% for the first two days and subsequently decreased to ~2.5% (Figure 6a). The high leakage observed for La(223Ra, 140Ba)VO4 NPs prepared using procedure A can be attributed to the rapid precipitation of La, 223Ra, and 140Ba as hydroxides, due to the high pH of the Na3VO4 solution (pH = 12.5) [38]. Particularly, the precipitation of hydroxide species has been associated with a slow interaction with vanadate ions that can lead to a low yield of LaVO4 NPs [38], which is shown by the low radiochemical yield obtained for both 223Ra (19.2 ± 4.1%) and 140Ba (16.5 ± 2.0%) (Figure 7). It is also expected that the formation of different radionuclide species due to pH changes will promote the uptake of radionuclides by surface sorption [65]. The sorption of radionuclides onto the particle surface could explain the high leakage observed for both 223Ra and 140Ba (Figure 6a). Differences in the radiochemical yield of divalent radionuclides (223Ra, 140Ba) with respect to trivalent 140La can be associated with their mechanism of incorporation within the crystal structure [66]. For both 223Ra and 140Ba, it will be harder to incorporate into tetragonal LaVO4 NPs and occupy La3+ sites, due to the smaller La–O length with respect to monoclinic LaVO4 NPs (2.5089 Å vs. 2.6029 Å) [52].
Among all the NPs produced by the different procedures, La(223Ra, 140Ba)VO4 NPs prepared using procedure B exhibited the lowest leakage of radionuclides (Figure 6b). Although the radiochemical yield of 223Ra and 140Ba increased with respect to those obtained by procedure A, NPs prepared following procedure B had a relatively low radiochemical yield for the different radionuclides (Figure 7). It is hypothesized that precipitation of vanadate species different from LaVO4, due to the low pH of the NP suspension [38], contributed to the low radiochemical yields. The precipitation of other kinds of lanthanum vanadate species was prevented in procedure C by the addition of KOH to adjust the pH of the NP suspension to 11. The radionuclide leakage increased slightly with respect to La(223Ra, 140Ba)VO4 NPs prepared using procedure B (Figure 6c), where the leakage of 223Ra and 140Ba reached a maximum of 4.6 ± 1.5% and 2.0 ± 1.8%, respectively. It is assumed that the dialysate activity of 223Ra and 140Ba may originate from radionuclides that were not coprecipitated within LaVO4 NPs, but rather adsorbed onto the particle surface [65]. The leakage of decay daughters 211Pb and 211Bi was maintained at less than 1.5% over time (Figure 6c), while the 223Ra, 140Ba, and 140La radiochemical yield increased to 87.4 ± 6.9%, 89.0 ± 8.3%, and 94.2 ± 4.3%, respectively. Overall, La(223Ra, 140Ba)VO4 NPs prepared using procedure C exhibited an enhanced radiochemical yield for both radium and barium radionuclides compared with that obtained with BaSO4 NPs (20% for >100 nm and 31–40% for <10 nm) [30,31].
The 223Ra radiochemical yield for La(223Ra, 140Ba)VO4 NPs was similar to that obtained with La(223Ra)PO4 core NPs (91%) and higher than that for core + 2 shells (80%) [19]. Although the leakage of radionuclides was <5% for both BaSO4 and La(223Ra, 140Ba)VO4 NPs, La(223Ra)PO4 core + 2 shells NPs exhibited a greater retention of decay daughters (<0.1% leakage) [19,30,31]. The enhanced retention of decay daughters can be associated with the core-shell structure of those NPs, where the nonradioactive shell acts as a barrier to prevent the relocation of radionuclides [18,19,20,22,23].
The stability and radionuclide retention of La(223Ra)VO4 NPs, synthesized by procedure, C were assessed by dialyzing the NP suspension against PBS (Figure 8). The 223Ra radiochemical yield calculated after dialysis against PBS for 15 days was 99.7 ± 0.2% (n = 3), which is comparable with that reported for intrinsically radiolabeled hydroxyapatite (Hap; 97.0 ± 0.5%) and titanium dioxide (TiO2; 99.1 ± 0.3%) NPs [26]. Although the leakage of decay daughters was not reported, the leakage or released activity of 223Ra from Hap NPs was significantly higher with respect to that of La(223Ra)VO4 NPs [26]. The decrease in radionuclide leakage after dialysis of La(223Ra)VO4 NPs against PBS may be related to the interaction of free radionuclides with phosphate species in solution, changes in NP colloidal stability, or both (Figure 8). It is hypothesized that the sorption and surface complexation of 223Ra and phosphate compounds in solution with LaVO4 NPs decreases the radionuclide activity in the dialysate [24,65]. The interaction of phosphate species with LaVO4 NPs after dilution in PBS was shown in their FTIR spectra (Figure S2 in the ESI). To address the influence of surface modification in the 223Ra radiochemical yield, La(223Ra)VO4 NPs were modified with carboxylate and phosphate compounds, centrifuged, and dispersed in deionized water multiple times. Surface modification with Na-Cit did not significantly alter the 223Ra radiochemical yield (99.6 ± 2.3%) with respect to La(223Ra)VO4 NPs (Figure 9). The 223Ra radiochemical yield of La(223Ra)VO4 NPs modified with NH4-Cit, TPP, and EDTA decreased ~15%, whereas >25% of 223Ra initial activity was lost when using Hex as a stabilizing compound (Figure 9). The decrease in 223Ra radiochemical yield after surface modification is related to the interaction of those stabilizing compounds with La cations on the particle surface and their ability to form La complexes. As shown with LaVO4 NPs, there is a mass loss associated with each stabilizing compound used for surface functionalization, which could lead to a decrease in 223Ra radiochemical yield. Also, the selectivity of stabilizing compounds to specific cations can lead to a lower radiochemical yield.
La(223Ra)VO4 and La(223Ra, 140Ba)VO4 NPs have exhibited promising capabilities as delivery platforms for 223Ra, as well as alkaline and lanthanide radionuclides. The synthesis conditions of LaVO4 NPs can be tailored to enhance the radiochemical yield and the retention of radionuclides within the NPs. It is expected that the fraction of 223Ra incorporated can be increased significantly (up to ~10 at. %) without affecting the crystal structure of LaVO4 NPs [52]. Although the nature of functional groups (i.e., hydroxyl groups) limits the colloidal stability and functionalization of LaVO4 NPs, the particle surface can be modified with bifunctional compounds. The use of bifunctional compounds has the potential to increase the stability of LaVO4 NP suspensions, prevent their accumulation in the RES, and allow their functionalization with targeting vectors. The selection of bifunctional compounds must be followed by an evaluation of the physicochemical characteristics of the modified NPs and an assessment of their encapsulation and retention of radionuclides.

4. Conclusions

LaVO4 NPs are promising delivery platforms for α-emitting radionuclides based on the high encapsulation of 223Ra and the high retention of its decay daughters, 211Pb and 211Bi. Synthesis parameters, particularly the order of mixing the stock solutions and the final pH of the NP suspension, influenced the morphology, crystal structure, and radionuclide encapsulation and retention of La(223Ra, 140Ba)VO4 NPs. Maintaining the La(223Ra, 140Ba)VO4 NP suspension at an acidic pH (4.9) yielded a high fraction of NPs with tetragonal structure and resultant radiochemical yields of 223Ra and 140Ba <40%. Increasing the pH of the NP suspension to 11 resulted in the precipitation of a high fraction of La(223Ra, 140Ba)VO4 NPs with a monoclinic structure and a twofold increase in 223Ra and 140Ba radiochemical yields (>85%). Surface modification of LaVO4 NPs was indicated by the characteristic vibration bands associated with carboxylate and phosphate groups and resulted in a threefold decrease in hydrodynamic diameter and 223Ra radiochemical yields between 74.7% and 99.6%. Altering the surface of La(223Ra, 140Ba)VO4 NPs enhances their colloidal stability and allows their functionalization with targeting vectors or fluorophores. The selection of bifunctional compounds must consider their effects on the encapsulation and retention of radionuclides over time. These results show the potential of La(223Ra)VO4 NPs for TAT, where current efforts are focused on optimizing their surface chemistry for functionalization and evaluating their cytotoxicity using in vitro models.

Supplementary Materials

The following are available online at https://www.mdpi.com/2624-845X/2/1/3/s1, Figure S1: Decay schemes of 223Ra and 140Ba, Table S1: Summary of γ-ray energies and intensities used to calculate the activity of each radionuclide, Figure S2: FTIR spectra of LaVO4 NPs, synthesized following procedure C (a) before and (b) after dispersion in PBS, Figure S3: FTIR spectra of LaVO4 NPs modified with NH4-Cit at different LaVO4:NH4-Cit molar ratios, Figure S4: FTIR spectra of LaVO4 NPs modified with Na-Cit at different LaVO4:Na-Cit molar ratios, Figure S5: FTIR spectra of LaVO4 NPs modified with EDTA at different LaVO4:EDTA molar ratios, Figure S6: FTIR spectra of LaVO4 NPs modified with TPP at different LaVO4:TPP molar ratios, Figure S7: FTIR spectra of LaVO4 NPs modified with Hex at different LaVO4:Hex molar ratios, Figure S8: FTIR spectra of LaVO4 + NH4-Cit and LaVO4 + EDTA NPs (a, c) before and (b, d) after dispersion in PBS (1X, pH = 7.4).

Author Contributions

Conceptualization, M.T.-G., S.M., S.M.D.; methodology, M.T.-G., S.M.D.; radioactive materials preparation, A.P., R.C.; characterization, M.T.-G., A.M., D.A.C.; writing—original draft preparation M.T.-G., S.M., S.M.D.; writing—review and editing, M.T.-G., A.M., R.C., S.M.D. All authors have read and agreed to the published version of this manuscript.

Funding

This research was sponsored by the Laboratory Directed Research and Development Program of Oak Ridge National Laboratory, managed by UT-Battelle, LLC, for the US Department of Energy (DOE).

Note

This manuscript has been authored by UT-Battelle, LLC, under contract DE-AC05-00OR22725 with the US Department of Energy (DOE). The US government retains and the publisher, by accepting the article for publication, acknowledges that the US government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this manuscript, or allow others to do so, for US government purposes. DOE will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan).

Data Availability Statement

Data is contained within the article and supplementary material.

Acknowledgments

Nanomaterials characterization was conducted at the Center for Nanophase Materials Sciences, which is a DOE Office of Science User Facility. This research was supported by the US DOE Isotope Program, managed by the Office of Science for Nuclear Physics, and by the ORNL Laboratory Directed Research and Development program. The authors would like to thank the staff of the Bethel Valley Radiochemical Processing group at ORNL for their contributions in isotope production and purification.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tafreshi, N.K.; Doligalski, M.L.; Tichacek, C.J.; Pandya, D.N.; Budzevich, M.M.; El-Haddad, G.; Khushalani, N.I.; Moros, E.G.; McLaughlin, M.L.; Wadas, T.J.; et al. Development of targeted alpha particle therapy for solid tumors. Molecules 2019, 24, 4314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Navalkissoor, S.; Grossman, A. Targeted Alpha Particle Therapy for Neuroendocrine Tumours: The Next Generation of Peptide Receptor Radionuclide Therapy. Neuroendocrinology 2019, 108, 256–264. [Google Scholar] [CrossRef] [PubMed]
  3. Marques, I.A.; Neves, A.R.; Abrantes, A.M.; Pires, A.S.; Tavares-da-Silva, E.; Figueiredo, A.; Botelho, M.F. Targeted alpha therapy using Radium-223: From physics to biological effects. Cancer Treat. Rev. 2018, 68, 47–54. [Google Scholar] [CrossRef] [PubMed]
  4. Song, H.; Hobbs, R.F.; Vakravelu, R.; Huso, D.L.; Esaias, C.; Apostolidis, C.; Morgenstern, A.; Sgouros, G. Radioimmunotherapy of breast cancer metastases with α-particle emitter 225Ac: Comparing efficacy with 213Bi and 90Y. Cancer Res. 2009, 69, 8941–8948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Juzeniene, A.; Bernoulli, J.; Suominen, M.; Halleen, J.; Larsen, R.H. Antitumor activity of novel bone-seeking, α-emitting 224Ra-solution in a breast cancer skeletal metastases model. Anticancer Res. 2018, 38, 1947–1955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. 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] [PubMed]
  7. Venetoclax and Lintuzumab-Ac225 in AML Patients—Full Text View—ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT03867682?term=actinium+pharmaceuticals&draw=2&rank=4 (accessed on 2 September 2020).
  8. Study to Evaluate the Safety, Tolerability,Pharmacokinetics, and Anti-Tumor Activity of a Thorium-227 Labeled Antibody-Chelator Conjugate, in Patients With Metastatic Castration Resistant Prostate Cancer—Full Text View—ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT03724747?term=NCT03724747&draw=2&rank=1 (accessed on 2 September 2020).
  9. First-in-Human Study of BAY2287411 Injection, a Thorium-227 Labeled Antibody-Chelator Conjugate, in Patients With Tumors Known to Express Mesothelin—Full Text View—ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT03507452?term=NCT03507452&draw=2&rank=1 (accessed on 2 September 2020).
  10. Sathekge, M.; Bruchertseifer, F.; Knoesen, O.; REyneke, F.; Lawal, I.; Lengana, T.; Davis, C.; Mahapane, J.; Corbett, C.; Vorster, M.; et al. 225Ac-PSMA-617 in chemotherapy-naive patients with advanced prostate cancer: A pilot study. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 129–138. [Google Scholar] [CrossRef] [Green Version]
  11. Kratochwil, C.; Bruchertseifer, F.; Giesel, F.L.; Weis, M.; Verburg, F.A.; Mottaghy, F.; Kopka, K.; Apostolidis, C.; Haberkorn, U.; Morgenstern, A. 225Ac-PSMA-617 for PSMA-Targeted α-Radiation Therapy of Metastatic Castration-Resistant Prostate Cancer. J. Nucl. Med. 2016, 57, 1941–1944. [Google Scholar] [CrossRef] [Green Version]
  12. van der Doelen, M.J.; Mehra, N.; Smits, M.; van Oort, I.M.; Jansenn, M.J.R.; Haberkorn, U.; Kratochwil, C.; Gerritsen, W. Clinical experience with PSMA-Actinium-225 (Ac-225) radioligand therapy (RLT) in end-stage metastatic castration-resistant prostate cancer (mCRPC) patients. J. Clin. Oncol. 2018, 36, 344. [Google Scholar] [CrossRef]
  13. Sofou, S.; Thomas, J.L.; Lin, H.Y.; McDevitt, M.R.; Scheinberg, D.A.; Sgouros, G. Engineered liposomes for potential α-particle therapy of metastatic cancer. J. Nucl. Med. 2004, 45, 253–260. [Google Scholar]
  14. Sofou, S.; Kappel, B.J.; Jaggi, J.S.; McDevitt, M.R.; Scheinberg, D.A.; Sgouros, G. Enhanced retention of the α-particle-emitting daughters of actinium-225 by liposome carriers. Bioconjug. Chem. 2007, 18, 2061–2067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. 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. Radiat. Isot. 2014, 85, 45–53. [Google Scholar] [CrossRef] [PubMed]
  16. Thijssen, L.; Schaart, D.R.; De Vries, D.; Morgenstern, A.; Bruchertseifer, F.; Denkova, A.G. Polymersomes as nano-carriers to retain harmful recoil nuclides in alpha radionuclide therapy: A feasibility study. Radiochim. Acta 2012, 100, 473–481. [Google Scholar] [CrossRef]
  17. 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. Isot. 2017, 128, 183–189. [Google Scholar] [CrossRef] [PubMed]
  18. 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]
  19. 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] [Green Version]
  20. Toro-González, M.; Dame, A.N.; Foster, C.M.; Millet, L.J.; Woodward, J.D.; Rojas, J.V.; Mirzadeh, S.; Davern, S.M. Quantitative encapsulation and retention of 227Th and decay daughters in core-shell lanthanum phosphate nanoparticles. Nanoscale 2020, 12, 9744–9755. [Google Scholar] [CrossRef]
  21. 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. Bioconjugate Chem. 2011, 22, 766–776. [Google Scholar] [CrossRef]
  22. Toro-González, M.; 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]
  23. Toro-González, M.; 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. B 2018, 6, 7985–7997. [Google Scholar] [CrossRef]
  24. Mokhodoeva, O.; Vlk, M.; Málková, E.; Kukleva, E.; Mičolová, P.; Štamberg, K.; Šlouf, M.; Dzhenloda, R.; Kozempel, J. Study of 223Ra uptake mechanism by Fe3O4 nanoparticles: Towards new prospective theranostic SPIONs. J. Nanopart. Res. 2016, 18, 1–12. [Google Scholar] [CrossRef]
  25. Cędrowska, E.; Pruszyński, M.; Gawęda, W.; Zuk, M.; Krysiński, P.; Bruchertseifer, F.; Morgenstern, A.; Karageorgou, M.A.; Bouziotis, P.; Bilewicz, A. Trastuzumab conjugated superparamagnetic iron oxide nanoparticles labeled with 225Ac as a perspective tool for combined α-radioimmunotherapy and magnetic hyperthermia of HER2-positive breast cancer. Molecules 2020, 25, 1025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Suchánková, P.; Kukleva, E.; Nykl, E.; Nykl, P.; Sakmár, M.; Vlk, M.; Kozempel, J. Hydroxyapatite and Titanium Dioxide Nanoparticles: Radiolabelling and In Vitro Stability of Prospective Theranostic Nanocarriers for 223Ra and 99mTc. Nanomaterials 2020, 10, 1632. [Google Scholar] [CrossRef] [PubMed]
  27. 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, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Piotrowska, A.; Męczyńska-Wielgosz, S.; Majkowska-Pilip, A.; Koźmiński, P.; Wójciuk, G.; Cędrowska, E.; Bruchertseifer, F.; Morgenstern, A.; Kruszewski, M.; Bilewicz, A. Nanozeolite bioconjugates labeled with 223Ra for targeted alpha therapy. Nucl. Med. Biol. 2017, 47, 10–18. [Google Scholar] [CrossRef] [PubMed]
  29. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Reissig, F.; Zarschler, K.; Hübner, R.; Pietzsch, H.; Kopka, K.; Mamat, C. Sub-10 nm Radiolabeled Barium Sulfate Nanoparticles as Carriers for Theranostic Applications and Targeted Alpha Therapy. ChemistryOpen 2020, 9, 797–805. [Google Scholar] [CrossRef]
  31. 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]
  32. 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]
  33. Mirzadeh, S.; Kumar, K.; Gansow, O.A. The Chemical Fate of 212Bi-DOTA Formed by β Decay of 212Pb(DOTA)2–. Radiochim. Acta 1993, 60, 1–10. [Google Scholar] [CrossRef]
  34. 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]
  35. 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, 1–13. [Google Scholar] [CrossRef] [Green Version]
  36. Neumeier, S.; Arinicheva, Y.; Ji, Y.; Heuser, J.M.; Kowalski, P.M.; Kegler, P.; Schlenz, H.; Bosbach, D.; Deissmann, G. New insights into phosphate based materials for the immobilisation of actinides. Radiochim. Acta 2017, 105, 961–984. [Google Scholar] [CrossRef]
  37. Rafiuddin, M.R.; Grosvenor, A.P. Probing the effect of radiation damage on the structure of rare-earth phosphates. J. Alloys Compd. 2015, 653, 279–289. [Google Scholar] [CrossRef]
  38. Huignard, A.; Gacoin, T.; Boilot, J.P. Synthesis and luminescence properties of colloidal YVO4:Eu phosphors. Chem. Mater. 2000, 12, 1090–1094. [Google Scholar] [CrossRef]
  39. Guerrini, L.; Alvarez-Puebla, R.A.; Pazos-Perez, N. Surface modifications of nanoparticles for stability in biological fluids. Materials 2018, 11, 1154. [Google Scholar] [CrossRef] [Green Version]
  40. Moore, T.L.; Rodriguez-Lorenzo, L.; Hirsch, V.; Balog, S.; Urban, D.; Jud, C.; Rothen-Rutishauser, B.; Lattuada, M.; Petri-Fink, A. Nanoparticle colloidal stability in cell culture media and impact on cellular interactions. Chem. Soc. Rev. 2015, 44, 6287–6305. [Google Scholar] [CrossRef] [Green Version]
  41. Toro-González, M.; Dame, A.N.; Mirzadeh, S.; Rojas, J.V. Encapsulation and retention of 225Ac, 223Ra, 227Th, and decay daughters in zircon-type gadolinium vanadate nanoparticles. Radiochim. Acta 2020, 108, 967–977. [Google Scholar] [CrossRef]
  42. Ropp, R.C.; Carroll, B. Precipitation of rare earth vanadates from aqueous solution. J. Inorg. Nucl. Chem. 1977, 39, 1303–1307. [Google Scholar] [CrossRef]
  43. Ansari, A.A.; Alam, M.; Labis, J.P.; Alrokayan, S.A.; Shafi, G.; Hasan, T.N.; Syed, N.A.; Alshatwi, A.A. Luminescent mesoporous LaVO4:Eu3+ core-shell nanoparticles: Synthesis, characterization, biocompatibility and their cytotoxicity. J. Mater. Chem. 2011, 21, 19310–19316. [Google Scholar] [CrossRef]
  44. Wang, H.; Wang, L. One-pot syntheses and cell imaging applications of poly(amino acid) coated LaVO4:Eu3+ luminescent nanocrystals. Inorg. Chem. 2013, 52, 2439–2445. [Google Scholar] [CrossRef] [PubMed]
  45. Tamilmani, V.; Kumari, A.; Rai, V.K.; Unni Nair, B.; Sreeram, K.J. Bright Green Frequency Upconversion in Catechin Based Yb3+/Er3+ Codoped LaVO4 Nanorods upon 980 nm Excitation. J. Phys. Chem. C 2017, 121, 4505–4516. [Google Scholar] [CrossRef]
  46. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  47. Zhang, F.; Li, G.; Zhang, W.; Yan, Y.L. Phase-Dependent Enhancement of the Green-Emitting Upconversion Fluorescence in LaVO4:Yb3+,Er3+. Inorg. Chem. 2015, 54, 7325–7334. [Google Scholar] [CrossRef] [PubMed]
  48. Salavati-Niasari, M.; Saleh, L.; Mohandes, F.; Ghaemi, A. Sonochemical preparation of pure t-LaVO4 nanoparticles with the aid of tris(acetylacetonato)lanthanum hydrate as a novel precursor. Ultrason. Sonochem. 2014, 21, 653–662. [Google Scholar] [CrossRef]
  49. Fan, W.; Bu, Y.; Song, X.; Sun, S.; Zhao, X. Selective synthesis and luminescent properties of monazite- and zircon-type LaVO4:Ln (Ln = Eu, Sm, and Dy) nanocrystals. Cryst. Growth Des. 2007, 7, 2361–2366. [Google Scholar] [CrossRef]
  50. Ansari, A.A.; Labis, J.P.; Alrokayan, S.A.H. Synthesis of water-soluble luminescent LaVO4:Ln3+ porous nanoparticles. J. Nanopart. Res. 2012, 14, 1–10. [Google Scholar] [CrossRef]
  51. Ropp, R.C.; Carroll, B. Dimorphic lanthanum orthovanadate. J. Inorg. Nucl. Chem. 1973, 35, 1153–1157. [Google Scholar] [CrossRef]
  52. Zhong, J.; Zhao, W. Novel dumbbell-like LaVO4:Eu3+ nanocrystals and effect of Ba2+ codoping on luminescence properties of LaVO4:Eu3+ nanocrystals. J. Sol-Gel Sci. Technol. 2014, 73, 133–140. [Google Scholar] [CrossRef]
  53. Yang, L.; Li, G.; Hu, W.; Zhao, M.; Sun, L.; Zheng, J.; Yan, T.; Li, L. Control Over the Crystallinity and Defect Chemistry of YVO4 Nanocrystals for Optimum Photocatalytic Property. Eur. J. Inorg. Chem. 2011, 2011, 2211–2220. [Google Scholar] [CrossRef]
  54. Yu, C.; Yu, M.; Li, C.; Zhang, C.; Yang, P.; Lin, J. Spindle-like lanthanide orthovanadate nanoparticles: Facile synthesis by ultrasonic irradiation, characterization, and luminescent properties. Cryst. Growth Des. 2009, 9, 783–791. [Google Scholar] [CrossRef]
  55. Da Silva, M.F.P.; Matos, J.R.; Isolani, P.C. Synthesis, characterization and thermal analysis of 1:1 and 2:3 lanthanide(III) citrates. J. Therm. Anal. Calorim. 2008, 94, 305–311. [Google Scholar] [CrossRef]
  56. Nash, K.L.; Brigham, D.; Shehee, T.C.; Martin, A. The kinetics of lanthanide complexation by EDTA and DTPA in lactate media. Dalton Trans. 2012, 41, 14547–14556. [Google Scholar] [CrossRef]
  57. Runowski, M.; Dąbrowska, K.; Grzyb, T.; Miernikiewicz, P.; Lis, S. Core/shell-type nanorods of Tb3+-doped LaPO4, modified with amine groups, revealing reduced cytotoxicity. J. Nanopart. Res. 2013, 15. [Google Scholar] [CrossRef] [Green Version]
  58. Parchur, A.K.; Prasad, A.I.; Rai, S.B.; Tewari, R.; Sahu, R.K.; Okram, G.S.; Singh, R.A.; Ningthoujam, R.S. Observation of intermediate bands in Eu3+ doped YPO4 host: Li+ ion effect and blue to pink light emitter. AIP Adv. 2012, 2, 032119. [Google Scholar] [CrossRef]
  59. Princz, E.; Szilágyi, I.; Mogyorósi, K.; Labádi, I. Lanthanide complexes of ethylenediaminotetramethylene-phosphonic acid. J. Therm. Anal. Calorim. 2002, 69, 427–439. [Google Scholar] [CrossRef]
  60. Buissette, V.; Moreau, M.; Gacoin, T.; Boilot, J.P.; Chane-Ching, J.Y.; Le Mercier, T. Colloidal synthesis of luminescent rhabdophane LaPO4:Ln3+·xH2O (Ln = Ce, Tb, Eu; x ≈ 0.7) nanocrystals. Chem. Mater. 2004, 16, 3767–3773. [Google Scholar] [CrossRef]
  61. Chen, H.; Chen, J.; Wang, L.; Zhou, C.; Ling, B.; Fu, J. A sensitive method for determination of trace amounts of chromate (III) with terbium (III) sodium hexametaphosphate chelate as fluorescent probe. Luminescence 2011, 26, 434–438. [Google Scholar] [CrossRef]
  62. Lloret, N.; Frederiksen, R.S.; Møller, T.C.; Rieben, N.I.; Upadhyay, S.; De Vico, L.; Jensen, J.H.; Nygård, J.; Martinez, K.L. Effects of buffer composition and dilution on nanowire field-effect biosensors. Nanotechnology 2013, 24, 035501. [Google Scholar] [CrossRef]
  63. Schomäcker, K.; Mocker, D.; Münze, R.; Beyer, G.J. Stabilities of lanthanide-protein complexes. Int. J. Radiat. Appl. Instrum. Part A Appl. Radiat. Isot. 1988, 39, 261–264. [Google Scholar] [CrossRef]
  64. Nie, S. Understanding and overcoming major barriers in cancer nanomedicine. Nanomedicine 2010, 5, 523–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Suchánková, P.; Kukleva, E.; Štamberg, 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]
  66. Tian, L.; Mho, S.-I. Enhanced photoluminescence of YVO 4:Eu3+ by codoping the Sr2+, Ba2+ or Pb2+ ion. J. Lumin. 2007, 122, 99–103. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the three synthesis procedures used to prepare La(223Ra, 140Ba)VO4, La(223Ra)VO4, and LaVO4 NPs.
Figure 1. Schematic representation of the three synthesis procedures used to prepare La(223Ra, 140Ba)VO4, La(223Ra)VO4, and LaVO4 NPs.
Jnt 02 00003 g001
Figure 2. The crystal structure of LaVO4 NPs was influenced by the synthesis procedure and the suspension’s pH. Diffraction patterns of LaVO4 NPs, synthesized following different procedures (A, B, and C), are compared with reference powder diffraction files for a LaVO4 compound having tetragonal (00-032-0504) and monoclinic (00-050-0367) crystal structures.
Figure 2. The crystal structure of LaVO4 NPs was influenced by the synthesis procedure and the suspension’s pH. Diffraction patterns of LaVO4 NPs, synthesized following different procedures (A, B, and C), are compared with reference powder diffraction files for a LaVO4 compound having tetragonal (00-032-0504) and monoclinic (00-050-0367) crystal structures.
Jnt 02 00003 g002
Figure 3. The synthesis procedure influenced the morphology of the LaVO4 NPs. Transmission electron micrographs of LaVO4 NPs, synthesized using procedure A (a,d), B (b,e), and C (c,f).
Figure 3. The synthesis procedure influenced the morphology of the LaVO4 NPs. Transmission electron micrographs of LaVO4 NPs, synthesized using procedure A (a,d), B (b,e), and C (c,f).
Jnt 02 00003 g003
Figure 4. FTIR spectra of the LaVO4 NPs, synthesized by procedures A, B, and C, show no significant difference in their surface chemistry.
Figure 4. FTIR spectra of the LaVO4 NPs, synthesized by procedures A, B, and C, show no significant difference in their surface chemistry.
Jnt 02 00003 g004
Figure 5. Surface modification of LaVO4 NPs with stabilizing compounds increases the presence of surface functional groups, as shown by the detection of additional vibrational bands. FTIR spectra of LaVO4 NPs, synthesized by procedure C (upper left), and after surface modification with NH4-Cit, Na-Cit, Hex, or TPP (1:1 LaVO4:compound molar ratio) and EDTA (3:1 LaVO4:EDTA molar ratio).
Figure 5. Surface modification of LaVO4 NPs with stabilizing compounds increases the presence of surface functional groups, as shown by the detection of additional vibrational bands. FTIR spectra of LaVO4 NPs, synthesized by procedure C (upper left), and after surface modification with NH4-Cit, Na-Cit, Hex, or TPP (1:1 LaVO4:compound molar ratio) and EDTA (3:1 LaVO4:EDTA molar ratio).
Jnt 02 00003 g005
Figure 6. Radionuclide leakage of La(223Ra, 140Ba)VO4 NPs, synthesized by procedure (a) A, (b) B, and (c) C (n = 3), over time. La(223Ra, 140Ba)VO4 NPs were dialyzed against deionized water, and dialysate aliquots were taken periodically and assayed using γ-ray spectroscopy.
Figure 6. Radionuclide leakage of La(223Ra, 140Ba)VO4 NPs, synthesized by procedure (a) A, (b) B, and (c) C (n = 3), over time. La(223Ra, 140Ba)VO4 NPs were dialyzed against deionized water, and dialysate aliquots were taken periodically and assayed using γ-ray spectroscopy.
Jnt 02 00003 g006
Figure 7. The radiochemical yield of La(223Ra, 140Ba)VO4 NPs, synthesized by different procedures (n = 3). Enhanced radiochemical yield was observed for La(223Ra, 140Ba)VO4 NPs, synthesized using procedure C.
Figure 7. The radiochemical yield of La(223Ra, 140Ba)VO4 NPs, synthesized by different procedures (n = 3). Enhanced radiochemical yield was observed for La(223Ra, 140Ba)VO4 NPs, synthesized using procedure C.
Jnt 02 00003 g007
Figure 8. The retention of 223Ra and its decay daughters (211Pb, 211Bi) within La(223Ra)VO4 NPs is enhanced when the NPs are dialyzed against PBS compared to dialysis against deionized water. Radionuclide leakage of La(223Ra)VO4 NPs (n = 3), synthesized by procedure C, was assessed after dialysis against PBS (1X, pH 7.2) to expose the particles to biologically relevant conditions.
Figure 8. The retention of 223Ra and its decay daughters (211Pb, 211Bi) within La(223Ra)VO4 NPs is enhanced when the NPs are dialyzed against PBS compared to dialysis against deionized water. Radionuclide leakage of La(223Ra)VO4 NPs (n = 3), synthesized by procedure C, was assessed after dialysis against PBS (1X, pH 7.2) to expose the particles to biologically relevant conditions.
Jnt 02 00003 g008
Figure 9. The 223Ra radiochemical yield decreased after surface modification with different stabilizing compounds. The effect of surface modification of La(223Ra)VO4 NPs, synthesized using procedure C, with different stabilizing compounds was assessed based on their 223Ra radiochemical yield.
Figure 9. The 223Ra radiochemical yield decreased after surface modification with different stabilizing compounds. The effect of surface modification of La(223Ra)VO4 NPs, synthesized using procedure C, with different stabilizing compounds was assessed based on their 223Ra radiochemical yield.
Jnt 02 00003 g009
Table 1. The tetragonal and monoclinic structure fraction and crystallite size of LaVO4 NPs, synthesized by different procedures. The fraction of crystal structure was influenced by the synthesis procedure and the suspension’s pH.
Table 1. The tetragonal and monoclinic structure fraction and crystallite size of LaVO4 NPs, synthesized by different procedures. The fraction of crystal structure was influenced by the synthesis procedure and the suspension’s pH.
ProcedureTetragonalMonoclinic
Fraction (%)Crystallite Size (nm)Fraction (%)Crystallite Size (nm)
A66.134.633.936.9
B46.445.553.632.3
C17.661.982.438.3
Table 2. Intensity diameter and zeta potential of diluted LaVO4 NPs in deionized water and PBS.
Table 2. Intensity diameter and zeta potential of diluted LaVO4 NPs in deionized water and PBS.
ProcedureDilution in Deionized WaterDilution in PBS
Diameter (nm)Zeta Potential (mV)Diameter (nm)Zeta Potential (mV)
A1186 ± 12917.4 ± 1.81036 ± 10128.1 ± 7.9
B1085 ± 9717.6 ± 3.2973 ± 12−19.2 ± 4.6
C995 ± 93−26.5 ± 8.4*−3.9 ± 3.4
* Particle size distribution was not measured, due to the poor stability of LaVO4 NP suspension.
Table 3. Characteristic vibrations and group assignment of LaVO4 NPs, synthesized by different procedures and modified with different stabilizing compounds.
Table 3. Characteristic vibrations and group assignment of LaVO4 NPs, synthesized by different procedures and modified with different stabilizing compounds.
Wavenumber Range (cm−1)Group Assignment
425.3–441.8La–O bond
533.4–542.8PO4 bending vibration
611.0–612.0PO4 bending vibration
743.4–825.8V–O vibration
998.2–1032.0PO4 stretching vibration
1389.5–1407.7Symmetric vibration carboxylate groups (carbonate species)
1434.2–1485.7Asymmetric vibration carboxylate groups (carbonate species)
1553.2–1579.8Asymmetric carboxylate bidentate bond stretching
1632.5–1638.6O–H bending
3208.2–3382.2O–H stretching
Table 4. The colloidal stability of LaVO4 NPs, in deionized water, is enhanced after surface modification. Hydrodynamic diameter (intensity distribution) and zeta potential of LaVO4 NPs, synthesized by procedure C after surface modification with NH4-Cit, Na-Cit, Hex, or TPP (1:1 LaVO4:compound molar ratio) and EDTA (3:1 LaVO4:EDTA molar ratio).
Table 4. The colloidal stability of LaVO4 NPs, in deionized water, is enhanced after surface modification. Hydrodynamic diameter (intensity distribution) and zeta potential of LaVO4 NPs, synthesized by procedure C after surface modification with NH4-Cit, Na-Cit, Hex, or TPP (1:1 LaVO4:compound molar ratio) and EDTA (3:1 LaVO4:EDTA molar ratio).
Stabilizing CompoundDilution in Deionized WaterDilution in PBS
Diameter (nm)Zeta Potential (mV)Diameter (nm)Zeta Potential (mV)
None995 ± 93−26.5 ± 8.4*−3.9 ± 3.4
NH4-Cit144.7 ± 4.8−17.0 ± 1.81003 ± 65−12.4 ± 3.5
Na-Cit207.1 ± 27.5−27.1 ± 3.61110 ± 144−21.4 ± 4.0
EDTA267.8 ± 11.8−32.9 ± 2.41126 ± 182−8.4 ± 2.1
TPP209.1 ± 5.2−30.5 ± 2.11112 ± 157−15.8 ± 2.1
Hex231.4 ± 14.4−33.6 ± 1.61100 ± 40−21.4 ± 4.0
* Particle size distribution was not measured, due to the poor stability of LaVO4 NP suspension.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Toro-González, M.; Peacock, A.; Miskowiec, A.; Cullen, D.A.; Copping, R.; Mirzadeh, S.; Davern, S.M. Tailoring the Radionuclide Encapsulation and Surface Chemistry of La(223Ra)VO4 Nanoparticles for Targeted Alpha Therapy. J. Nanotheranostics 2021, 2, 33-50. https://doi.org/10.3390/jnt2010003

AMA Style

Toro-González M, Peacock A, Miskowiec A, Cullen DA, Copping R, Mirzadeh S, Davern SM. Tailoring the Radionuclide Encapsulation and Surface Chemistry of La(223Ra)VO4 Nanoparticles for Targeted Alpha Therapy. Journal of Nanotheranostics. 2021; 2(1):33-50. https://doi.org/10.3390/jnt2010003

Chicago/Turabian Style

Toro-González, Miguel, Allison Peacock, Andrew Miskowiec, David A. Cullen, Roy Copping, Saed Mirzadeh, and Sandra M. Davern. 2021. "Tailoring the Radionuclide Encapsulation and Surface Chemistry of La(223Ra)VO4 Nanoparticles for Targeted Alpha Therapy" Journal of Nanotheranostics 2, no. 1: 33-50. https://doi.org/10.3390/jnt2010003

Article Metrics

Back to TopTop