Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Precision Atomistic Structures of Actinium-/Radium-/Barium-Doped Lanthanide Nanoconstructs for Radiotherapeutic Applications
My Activity
  • Open Access
Article

Precision Atomistic Structures of Actinium-/Radium-/Barium-Doped Lanthanide Nanoconstructs for Radiotherapeutic Applications
Click to copy article linkArticle link copied!

  • Monojoy Goswami
    Monojoy Goswami
    Chemical Sciences Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States
  • Miguel Toro-González
    Miguel Toro-González
    Radioisotope Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States
  • Jisue Moon
    Jisue Moon
    Radioisotope Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States
    More by Jisue Moon
  • Sandra Davern*
    Sandra Davern
    Radioisotope Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States
    *Email: [email protected]
Open PDFSupporting Information (1)

ACS Nano

Cite this: ACS Nano 2024, 18, 26, 16577–16588
Click to copy citationCitation copied!
https://doi.org/10.1021/acsnano.3c13213
Published June 17, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Lanthanide vanadate (LnVO4) nanoconstructs have generated considerable interest in radiotherapeutic applications as a medium for nanoscale-targeted drug delivery. For cancer treatment, LnVO4 nanoconstructs have shown promise in encapsulating and retaining radionuclides that emit alpha-particles. In this work, we examined the structure formation of LnVO4 nanoconstructs doped with actinium (Ac) and radium (Ra), both experimentally and using large-scale atomistic molecular dynamics simulations. LnVO4 nanoconstructs were synthesized via a precipitation method in aqueous media. The reaction conditions and elemental compositions were varied to control the structure, fluorescence properties, and size distribution of the LnVO4 nanoconstructs. LnVO4 nanoconstructs were characterized by X-ray diffraction, Raman spectroscopy, and fluorescence spectroscopy. Molecular dynamics simulations were performed to obtain a fundamental understanding of the structure–property relationship between radionuclides and LnVO4 nanoconstructs at the atomic length scale. Molecular dynamics simulations with well-established force field (FF) parameters show that Ra atoms tend to distribute across the nanoconstructs’ surface in a broader coordination shell, while the Ac atoms are arranged inside a smaller coordination shell within the nanocluster. The Ba atoms prefer to self-assemble around the surface. These theoretical/simulation predictions of the atomistic structures and an understanding of the relationship between radionuclides and LnVO4 nanoconstructs at the atomic scale are important because they provide design principles for the future development of nanoconstructs for targeted radionuclide delivery.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2024 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Targeted radionuclide therapy (TRT) has emerged as a promising treatment modality for different types of cancer. (1) Cancer cells die because of a combination of direct, bystander, and systemic responses to radioactive emissions. (2,3) Among these emissions, alpha particles are highly cytotoxic because they induce complex DNA damage, resulting in double-strand breaks. The DNA damage is unrepairable and thus can overcome both radiation and chemotherapy resistance. (3) These characteristics are ideal for cancer treatment when alpha-emitting radionuclides are delivered to cancerous tissue. Alpha-emitting radionuclides are delivered by chemical affinity or by targeting highly expressed tumor antigens with radioimmunoconjugates. (4−8) These approaches are susceptible to the relocation of radionuclides to healthy tissue due to transmetalation and the bond-breaking recoil energy of decay daughters. (2,9−11) To overcome these challenges, multiple approaches have been proposed: a local administration of alpha-emitting radionuclides, rapid uptake of the radioimmunoconjugate by the cell, use of pretargeted radioimmunotherapy, and encapsulating of alpha-emitting radionuclides within nanoconstructs. (10,12) Nanoconstructs offer multiple advantages ranging from the encapsulation of multiple radionuclides and enhanced retention of decay daughters, to the modification of their surface with various functionalities. (13−18) These characteristics make nanoconstructs a promising radionuclide delivery vehicle for TRT.
Among nanoconstructs, lanthanide phosphate and lanthanide vanadate have shown a high encapsulation of [225Ac]Ac3+, [223Ra]Ra2+, and [227Th]Th4+. (13,14,19−24) The retention of their decay daughters─partial or quantitative─was influenced by the nanoconstruct composition and structure. (20,21,23) The composition influences the electron density and, thus, the stopping power of the nanoconstruct as a whole. (25) The structure provides a physical barrier to retain the decay daughters within the nanoconstructs. (20,21,23) As decay daughters can escape the nanoconstruct either by direct ejection during recoil or by leaching owing to lattice damage, (19) the radionuclide’s location within the nanoconstruct influences its retention. (26−29) Experimental validation of the radionuclide’s location is challenging because of length and time scales, limited access to characterization techniques, and finite availability of alpha-emitting radionuclides. While we report several experimental data sets here, wet-lab experiments with radionuclides become extremely difficult due to the very nature of the material properties of alpha-emitting radionuclides. Thus, it becomes imperative to perform in silico experiments of radionuclide nanoconstructs that can predict their structure–property relationship to guide and optimize the performance of wet-lab experiments. Large-scale atomistic molecular dynamics (MD) simulations of nanoconstructs, along with experimental observations, can provide a detailed understanding of these structures at the atomic length scale. MD simulations also contribute to a better understanding of the mechanisms behind enhanced decay daughter retention: dissipation of the recoil energy through atomic displacements and translation within the whole nanoconstruct; (19) and implantation into an adjacent nanoconstruct enhanced by agglomeration. (27) Understanding these mechanisms and their relationships is critical for designing radioactive nanoconstructs in clinical settings.
This work elucidates a fundamental understanding of lanthanide vanadate (LnVO4) nanoconstructs on atomic length and time scales. Experimentally, LnVO4 nanoconstructs were synthesized by precipitation in an aqueous media. Two different LnVO4 nanoconstructs were prepared incorporating either [225Ac]Ac3+ in one or [223Ra]Ra2+ in another, respectively. LnVO4 nanoconstructs doped with surrogate cations were characterized with X-ray diffraction (XRD), Raman spectroscopy, fluorescence spectroscopy, dynamic light scattering (DLS), and transmission electron microscopy (TEM). Radioactive LnVO4 nanoconstructs were characterized by γ-ray and fluorescence spectroscopy. Nanoconstruct characterization was performed to understand the influence of the surrogate/radionuclide concentration on the structure and encapsulation efficiency. The large-scale MD simulations provided an understanding of the structure–property relationships between radionuclides and LnVO4 nanoconstructs using experimentally verified force field parameters. These relationships can help design nanoconstructs for TRT, thereby driving precision nanomedicine and radiotherapy administration to eradicate cancer and minimize side effects.

Results and Discussion

Click to copy section linkSection link copied!

Crystalline La(1–x)EuxVO4 nanoconstructs were synthesized by precipitation of LnCl3 and Na3VO4 in the presence of 6-AHA (Figure 1a). Nanoconstructs with a molar concentration of Eu below 10% were characterized by a combination of monoclinic and tetragonal structures (Figure 1a and Table S2). Higher Eu molar concentrations (>10%) resulted in the precipitation of nanoconstructs with a tetragonal structure (Figure 1a and Table S2). The changes in the crystal structure of LnVO4 nanoconstructs have been associated with the synthesis method, the reaction pH, and the concentration of lanthanides. The synthesis method affects the nucleation and growth of nanoconstructs and, thus, the formation of different crystal structures. (30−32) The reaction pH determines the form of the vanadium ion and thus the precipitation of compounds with different Ln:V stoichiometry. (33−35) Lastly, the lanthanide ionic radii influence the crystal structure, where small lanthanides (Ln = Ce to Lu) precipitate preferentially with a tetragonal structure. (31,36) Previous research has reported that rapid precipitation of La3+ and Eu3+ with [VO4]3– can result in the formation of nanoconstructs with different crystal structures (i.e., monoclinic LaVO4 and tetragonal EuVO4). (31) For the synthesized La(1–x)EuxVO4 (x < 10%), the decrease in the unit cell parameters (Table S2 and Figure S2a) suggests that Eu3+ cations have entered the structure of LaVO4 rather than precipitated as a secondary phase. (31) The phase transition dependence of La(1–x)EuxVO4 nanoconstructs with Eu3+ concentration was also confirmed by Raman spectroscopy (Figure 1b). The Raman spectra are characterized by two regions corresponding to the tetrahedral [VO4]3– group (700–900 cm–1) and the Ln–O (<500 cm–1) vibrations (Figure 1b). (30,31,37) Increasing the Eu3+ molar concentration within La(1–x)EuxVO4 nanoconstructs changes the characteristic Raman shift for both [VO4]3– and Ln–O vibrations (Figure 1b). These changes in the Raman shift match the phase transition observed with XRD and with other lanthanide vanadate crystals. The size distribution of La(1–x)EuxVO4 nanoconstructs in water was studied by using DLS (Figure 1c). A significant decrease in the hydrodynamic size was observed from LaVO4 (Zave = 295 nm) to EuVO4 (Zave = 37 nm) nanoconstructs. This drastic change in hydrodynamic size could be caused by the change in crystal structure (monoclinic to tetragonal), and thus, the position of lanthanide cations within the nanoconstructs is altered. This position will affect the interaction of lanthanide cations with the carboxylate group of 6-AHA and, thus, influence the growth and stability of the nanoconstructs. There is no clear trend in the hydrodynamic size for intermediate compositions of La and Eu cations within La(1–x)EuxVO4 nanoconstructs (Figure 1c).

Figure 1

Figure 1. (a, d) Diffraction patterns, (b, e) Raman spectra, and (e, f) size distribution of (a–c) La(1–x)EuxVO4 and (d–f) (La0.9Eu0.1)(1–y)BayVO4 of nanoconstructs.

Lanthanide vanadate nanoconstructs were doped with Ba2+ as a surrogate for [223Ra]Ra2+. (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs were characterized by a tetragonal crystal structure when incorporating up to 25% Ba2+ (Figure 1d). Successful doping of Ba2+ into the La0.9Eu0.1VO4 structure was evidenced by the shift to lower angles of the characteristic peak around 24° (Figure S2b). (38) At a molar concentration of 50% Ba2+, the diffraction pattern changed significantly with respect to the lower concentrations of Ba2+. After performing Rietveld refinement, it was determined that increasing the concentration of Ba2+ within (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs results in a polycrystalline sample. (39) The phases within (La0.9Eu0.1)0.5Ba0.5VO4 nanoconstructs correspond to LaVO4 (space group P121/n1, 72.7%) and Ba2V3O9 (space group P12_1/m1, 27.3%). The fitting of these phases to the diffraction pattern resulted in a χ2 = 7.03, suggesting the model could be reasonable based on the quality of the diffraction pattern (i.e., broad peaks owing to nanoconstruct size). Raman spectra for (La0.9Eu0.1)(1–y)BayVO4 (y = 0, 5, 10, and 25%) matched the characteristic vibrations observed for La(1–x)EuxVO4 nanoconstructs (Figure 1e). In agreement with the XRD patterns, a 50% molar concentration of Ba caused a significant change in the vibrational bands (Figure 1e). No characteristic Ln–O or [VO4]3– vibrations were observed for (La0.9Eu0.1)0.5Ba0.5VO4 nanoconstructs because of the lower level of crystallinity of the sample. The hydrodynamic size of the (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs increased with higher concentrations of Ba (Figure 1f). The increase in size is associated with the precipitation and/or aggregation of nanoconstructs because of a lower concentration of surface lanthanide cations available for complexation with 6-AHA.
The excitation and emission of Eu3+ within La(1–x)EuxVO4 and (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs were studied to assess the influence of dopant ions. Initially, the Eu3+ concentration on La(1–x)EuxVO4 nanoconstructs was varied from 0 to 100%. The excitation spectra were characterized by the charge transfer band (CTB) of the [VO4]3– group (270–350 nm) and the Eu f–f transitions (>350 nm) when using an emission wavelength of 618 nm (Figure 2a). (31,38) The mechanism of luminescence involves: (1) absorption of ultraviolet photons by the [VO4]3– groups, (2) nonradiative transfer to Eu3+ ions, and (3) Eu3+ decay to the ground state by radiative transition. (40) The intensity, shape, and position of the CTB changed with the Eu3+ concentration, whereas only the intensity of the Eu f–f transitions was altered (Figure 2a). A higher CTB intensity, without significant change in location and shape, was obtained when Eu3+ was increased from 2 to 10%. A blue shift of the [VO4]3– CTB was obtained for Eu3+ concentrations greater than 25%. This shift of the [VO4]3– CTB matches the change in crystal structure and the local symmetry of the [VO4]3– group. (38) This change in symmetry causes a shorter distance between Eu3+ and [VO4]3– tetrahedrons and, thus, an increase in overlap between wave functions. (30) The characteristic 5D07FJ (J = 1, 2, 3, 4) Eu3+ transitions were observed in the emission spectra of La(1–x)EuxVO4 nanoconstructs (Figure 2b). (31,32,38) The intensity of the magnetic dipole transition (5D07F1, λ = 591 nm) and electric dipole transition (5D07F2, λ = 618 nm) increased with the Eu3+ concentration up to 50% (Figure 2b). This intensity increase is associated with a greater concentration of emission centers and changes in the crystal structure of La(1–x)EuxVO4 nanoconstructs. (32,38) A change in the crystal structure causes an overlap between wave functions and, thus, higher emission intensity. (30) The intensities of both magnetic and electric dipole transitions were not significantly different for Eu3+ concentrations greater than 50%. This plateau in emission intensity is due to the presence of concentration quenching. (32) An understanding of the Eu3+ environment within the La(1–x)EuxVO4 nanoconstructs is possible by following the emission intensity ratio between the magnetic (5D07F1, λ = 591 nm) and electric (5D07F2, λ = 618 nm) dipole transitions as can be observed from Figure 2c. The ratio showed an increase in magnitude from 2 to 10% Eu3+ concentration, followed by a decrease with higher Eu3+ concentrations. This trend is due to the change in crystal structure and a potential concentration quenching effect with an increase in the Eu3+ molar concentration.

Figure 2

Figure 2. Excitation spectra, emission spectra, and emission peak ratio of (a–c) La(1–x)EuxVO4 and (d–f) (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs.

The concentration of Ba2+ altered the excitation and emission spectra of the (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs, as shown in Figure 2d,e. An increase and red-shift of the [VO4]3– CTB excitation maxima was observed with greater concentrations of Ba2+ (Figure 2d). These changes are associated with the dependence of the [VO4]3– CTB on the lattice parameters. Particularly, a large metal ion such as Ba2+ will cause significant distortions to the structure. (38,39) At 50% Ba2+ concentration, the [VO4]3– CTB intensity decreases significantly because of the change in the crystal structure and crystallinity of the sample (Figure 2d). The emission spectra of (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs were characterized by the 5D07FJ (J = 1, 2, 3, 4) Eu3+ transitions (Figure 2e). The Eu3+ emission intensity was amplified by increasing the Ba2+ concentration to 25% (Figure 2e). This effect is related to changes in the lattice parameters and the potential introduction of defects and traps. (38,39) The emission intensity decreased significantly when 50% Ba2+ was used (Figure 2e). This decrease is caused by both structural changes (i.e., tetragonal phase has a higher emission than monoclinic) and the potential precipitation of Ba aggregates. (31,38) The intensity ratio between the magnetic and electric dipole transitions did not change significantly with the Ba2+ concentration (Figure 2f).
Doping LaVO4 and La0.9Eu0.1VO4 nanoconstructs with [225Ac]Ac3+ and [223Ra]Ra2+ was performed to assess changes in their encapsulation efficiency and fluorescence spectra. Initial experiments involved the synthesis of LaVO4 and La0.9Eu0.1VO4 nanoconstructs doped with [225Ac]Ac3+ as described in the experimental section. La[225Ac]AcVO4 nanoconstructs had an encapsulation efficiency of 65.4 ± 1.6% (n = 3), whereas 29.1 ± 1.7% (n = 5) was obtained for La0.9Eu0.1[225Ac]AcVO4 nanoconstructs. The encapsulation efficiency of La[225Ac]AcVO4 is similar to that obtained with [144Ce]CeVO4 nanoconstructs with 69.9 ± 7.6% (n = 3). The encapsulation efficiency (65–69%) is expected to be driven by the lower molar concentration of [VO4]3– during synthesis (i.e., 1:0.75 Ln:VO4). The LnCl3, [225Ac]Ac3+, and 6-AHA mixture pH may also play a critical role in the encapsulation efficiency of [225Ac]Ac3+. We observed that adjusting the pH above 7.2 can increase the encapsulation efficiency of [144Ce]Ce3+ above 90%. These results show that the reaction’s pH and, thus, the speciation of [225Ac]Ac3+ and vanadate species determine the encapsulation efficiency within LnVO4 nanoconstructs. The decrease in the [225Ac]Ac3+ encapsulation efficiency within La0.9Eu0.1[225Ac]AcVO4 nanoconstructs may have been influenced by the elemental composition and crystal structure. The difference in activity (i.e., number of [225Ac]Ac3+ moles) may also have played a role in the encapsulation efficiency within nanoconstructs (Table S3). Higher [225Ac]Ac3+ activities (2–50 μCi) were used to synthesize La0.9Eu0.1[225Ac]AcVO4 nanoconstructs because of the characterization with fluorescence spectroscopy (Table S3). Although these results suggest that higher activities result in unstable incorporation of [225Ac]Ac3+ within LnVO4, the reported encapsulation efficiencies are comparable to those obtained with La[225Ac]AcPO4 nanoparticles (47 ± 5%). (19) Woodward et al. used a La/Ac molar ratio of approximately 30,000, (19) which is significantly higher compared to the Ln/Ac molar ratios used in this work (Table S3). Lastly, the lower encapsulation efficiency may have been caused by a decrease in the reaction’s pH due to the higher volume of [225Ac]Ac3+ solution used (i.e., 0.1 M HNO3). As was pointed out previously, maintaining the reaction’s pH above 7 increases the encapsulation efficiency, whereas a slightly acidic pH can decrease the encapsulation efficiency below 50%.
The excitation spectrum of La0.9Eu0.1[225Ac]AcVO4 nanoconstructs was characterized by the CTB of the [VO4]3– group and the Eu f–f transitions when an emission wavelength of 618 nm was used (Figure S3a,b). After the decay of [225Ac]Ac3+ (approximately 3 half-lives), the intensity of the CTB and Eu f–f transitions decreased significantly when using an emission wavelength of 618 nm (Figure S4a,b). These changes could be associated with the distortion and/or damage to the crystal structure caused by the alpha-particle decay. The emission spectrum of La0.9Eu0.1[225Ac]AcVO4 nanoconstructs was characterized by the 5D07FJ (J = 1, 2, 3, 4) Eu3+ transitions (Figure S3c,d). No significant changes to the Eu3+ transitions were observed after the decay of [225Ac]Ac3+ (Figure S4c,d). These results were expected since [225Ac]Ac3+ is spectroscopically silent due to its electronic structure. (41) However, clear changes in the emission spectra were observed among La0.9Eu0.1VO4, La0.9Eu0.1[225Ac]AcVO4, and decayed La0.9Eu0.1[225Ac]AcVO4 nanoconstructs (Figure 3a,b). First, an emission peak around 440–450 nm corresponding to the 3T11A1 and 3T2lA1 energy transfer of [VO4]3– is more intense in La0.9Eu0.1[225Ac]AcVO4 nanoconstructs compared to that of La0.9Eu0.1VO4 (Figure 3a). (42,43) The emission of [VO4]3– is influenced by lattice defects within the nanoconstructs, (44,45) suggesting that an increase in emission intensity within La0.9Eu0.1[225Ac]AcVO4 nanoconstructs could be attributed to lattice defects. These lattice defects could be caused by [225Ac]Ac3+, its decay daughters, or the radiation damage induced during alpha-particle emission. This radiation damage could be responsible for the increased intensity of this band after the decay of [225Ac]Ac3+ (Figure 3a). Second, La0.9Eu0.1[225Ac]AcVO4 shows a higher 5D07F4 transition (λ = 700 nm) compared to that of La0.9Eu0.1VO4 nanoconstructs (Figure 3b). This increase in the intensity of the 5D07F4 transition has been associated with the structural distortion of the EuO8 coordination polyhedron from a cubic geometry to the D2 symmetry. (46,47) It is hypothesized that [225Ac]Ac3+ and decay daughters may cause distortion of the EuO8 coordination polyhedron due to differences in ionic radii. (46−48) This hypothesis is supported by the slight decrease in the intensity of the 5D07F4 transition after the decay of [225Ac]Ac3+ (Figure 3b). The excitation spectra of La[225Ac]AcVO4 nanoconstructs were characterized by multiple bands depending on the emission wavelength used (Figure S5a). The emission spectrum of La[225Ac]AcVO4 nanoconstructs had an intense band at 450 nm related to the 3T11A1 and 3T2lA1 energy transfer of [VO4]3–. (42,43) Various bands with less intensity were observed in the emission spectrum of La[225Ac]AcVO4 nanoconstructs (Figure S5b). These bands could originate from the defect centers within the nanoconstructs, rather than from the [VO4]3– species. (44) After decay of [225Ac]Ac3+, no significant changes were observed in the excitation spectra (Figure S6a). Although the emission spectra had the characteristic band from [VO4]3– species, the other bands were equivalent to the background (Figure S6b). As was discussed previously, the intensity of the [VO4]3– emission increased after multiple [225Ac]Ac3+ half-lives because of defects caused by radiation damage. (44,45) Overall, these observations show the potential to use La0.9Eu0.1VO4 nanoconstructs and fluorescence spectroscopy to detect and analyze the effects of small concentrations of radionuclides, dopant ions, and potential radiation damage on synthesized nanoconstructs.

Figure 3

Figure 3. (a, b) Emission spectra of La0.9Eu0.1VO4, La0.9Eu0.1[225Ac]AcVO4, and decayed La0.9Eu0.1[225Ac]AcVO4 nanoconstructs. The emission spectra were normalized to the 618 nm peak corresponding to the Eu3+5D07F2 transition. (c) Emission spectra of La[225Ac]AcVO4 and decayed La[225Ac]VO4 nanoconstructs. The emission spectra were normalized to the 450 nm peak corresponding to the [VO4]3–3T11A1 and 3T2lA1 energy transfer. (d) Encapsulation efficiency of [223Ra]Ra2+ within (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs.

The encapsulation efficiency of [223Ra]Ra2+ within LaVO4 and La0.9Eu0.1VO4 nanoconstructs synthesized as described in the experimental section was negligible (1.4 ± 1.0%, n = 3). The authors have previously shown that the encapsulation of [223Ra]Ra2+ within LaVO4 and GdVO4 nanoconstructs, prepared by precipitation without 6-AHA, was determined by the synthesis method and reaction conditions. (14,22) Precipitation of La0.9Eu0.1VO4 nanoconstructs at room temperature without 6-AHA resulted in a slight increase in [223Ra]Ra2+ encapsulation efficiency (3.8 ± 0.1%, n = 3). Based on the effect of pH on [223Ra]Ra2+ encapsulation efficiency, (22) NH4OH was added to adjust the pH to approximately 11 during coprecipitation of LaCl3 and Na3VO4 at room temperature. Increasing the pH during LaVO4 nanoconstruct precipitation resulted in a [223Ra]Ra2+ encapsulation efficiency of 98.8 ± 0.5% (n = 3). By adjusting the reaction’s pH to 11, the [223Ra]Ra2+ encapsulation efficiency was increased to 99.1 ± 0.2% (n = 2) when using LaVO4 nanoconstructs synthesized as described in the experimental methods (i.e., with 6-AHA and heated at 90 °C for 30 min). Contrary to the nonradiological experiments, La[223Ra]RaVO4 nanoconstructs synthesized under these conditions (i.e., basic pH) precipitated in deionized water over time. Precipitation of La[223Ra]RaVO4 nanoconstructs could be driven by the suspension’s pH and hence alteration of the zeta potential. An enhancement of the [223Ra]Ra2+ encapsulation efficiency with a pH increase may be associated with the incorporation mechanism. The sorption and complexation of [223Ra]Ra2+ on the nanoconstruct surface has been shown to be the driving force for TiO2 and hydroxyapatite nanoparticles. (49−54) Complexing reactions of [223Ra]Ra2+ with [CO3]2–, [VO4]3–, and OH could be important under the reaction conditions tested in this work. (52−55) Based on the studies performed with TiO2 and hydroxyapatite, no significant difference was observed between intrinsic and surface incorporation. (51) Although surface incorporation was not studied in this work, it could play a significant role in the encapsulation efficiency of [223Ra]Ra2+ within LnVO4 nanoconstructs. (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs were doped with [223Ra]Ra2+ to understand the effect of nanoconstruct composition on the encapsulation efficiency. The [223Ra]Ra2+ encapsulation efficiency was greater than 90% for the different Ba2+ concentrations (Figure 3d). The mean [223Ra]Ra2+ encapsulation efficiency decreased from 98.2 ± 0.5% (n = 3) to 92.8 ± 7.3% (n = 3) for La0.9Eu0.1VO4 and (La0.9Eu0.1)0.5Ba0.5VO4 nanoconstructs, respectively. A decrease in [223Ra]Ra2+ encapsulation efficiency could be associated with the competition with Ba2+ for sites within the La0.9Eu0.1VO4 nanoconstruct. Previous research has shown that a low [223Ra]Ra2+ encapsulation efficiency is associated with competing impurities and nonspecific sorption. (50) Based on surrogate studies, the high encapsulation efficiencies obtained could be explained by the precipitation of [223Ra]Ra2+/Ba2+ as a secondary phase (Figure 1d). Previous research has studied the in vitro stability of Ln[223Ra]RaVO4 nanoconstructs. The stability of Ln[223Ra]RaVO4 was influenced by the type of radionuclide, the reaction conditions, and the dialysis buffer. (14,22) These results suggest that the complexation of [225Ac]Ac3+ and [223Ra]Ra2+ on the LnVO4 surface is critical for nanoconstruct stability. Because of limitations in handling radioactive samples, modeling and simulation of LnVO4 nanoconstructs doped with [225Ac]Ac3+ and [223Ra]Ra2+ are required to understand their formation and radionuclide-structure interaction.
From the discussion in the method section, it has been shown that the simulation results of pure lanthanide match fairly well with the experimental findings of pristine LnVO4 nanoconstructs (Figure 7), and the liquid water structure within the nanoconstruct (Figure S1). Relying on these experimentally verified lanthanide simulations and previously reported FF parameters, atomistic MD simulations were performed by introducing three different dopant ions: Ac3+, Ba2+, and Ra2+ for the three different simulation systems. Two dopant ion concentrations were used, 3 mol % for Ac3+ and Ra2+ and 25 mol % for Ba2+, to ensure that the simulations were commensurate with experimental conditions. All the simulations for (1) Ac–La0.9Eu0.1VO4, (2) Ba–La0.9Eu0.1VO4, and (3) Ra–La0.9Eu0.1VO4 were equilibrated for 2 ns in aqueous solution at 67 °C, which is close to the experimental conditions, with subsequent cooling of the nanoconstructs to ambient temperature, T = 27 °C. Once the equilibration was achieved, the statistical average of the structural properties was obtained from a 0.5 ns simulation within the last 200 ps trajectory. For the low-temperature systems (T = 27 °C) the Ac/Ra systems' statistical properties were obtained from 1 ns simulation, and the structural properties were obtained from the last 200 ps trajectory.
Figure 4 shows the Ac–Ln and Ra–Ln structural properties as obtained from radial distribution function (RDF) calculations at 3.0 mol % of Ac/Ra. Figure 4a–c shows the RDF at T = 67 °C, and Figure 4d–g shows the simulation snapshot at the end of 500 ps production runs after equilibrating for 2 ns. From the snapshots, the Ac–Ln structures exhibit larger, connected percolated clusters (Figure 4d), while the Ra–Ln clusters are smaller, separated, and spread across the simulation box (Figure 4e). This result suggests that Ra–Ln can potentially form smaller nanoconstructs, which is highly relevant for radiotherapy applications. These structures exhibit excellent stability as after cooling down the structures to T = 27 °C, the RDF does not show any difference as shown in Figure S2. The structural stability is a result of strong Coulombic interactions between charged atoms that allow oppositely charged atoms to strongly adhere to each other due to electrostatics overtaking entropic forces between the charged atoms. The stable structure can also be seen from the simulation snapshots shown in Figures S3 and S4.

Figure 4

Figure 4. Structural details of Ac–Ln and Ra–Ln nanoconstructs at 3 mol % of Ac/Ra doping at 67 °C. Three different RDFs, Ac/Ra–O, La–O, and Eu–O are shown for the (a) Ac–Ln system and (b) Ra–Ln system. The comparison between Ac–O and Ra–O are shown in (c). Colors are represented in the legends. (d, e) Simulation snapshots of Ac–O and Ra–O systems, respectively. Water molecules are not shown for clarity. (f, g) Closeup of the nanoconstructs for Ac–O and Ra–O systems, respectively. Color scheme in the snapshots: Ac/Ra─blue, La─cyan, Eu─yellow, V─green, and O─red. The low-temperature, T = 27 °C, RDFs are compared in Figure S2 showing structural stability of the nanoconstructs.

The positions of the peaks between the electropositive Ra2+/Ac3+ and the electronegative oxygen atoms from VO4 are representative of the relative positions of two oppositely charged ions inside the nanoconstruct. Because these oppositely charged atoms are attracted by strong electrostatic interactions, the RDFs show close proximity of the positions of respective atoms inside these nanoconstructs. In Figures 4 a–c, both the Ra- and Ac-doped nanoconstructs show that Eu–O (red) and La–O (green) have a similar peak and short-range ordering until approximately 8 Å. The RDFs of Ac–O and Ra–O (black) exhibit a closer first peak for Ac–O with respect to that of Ra–O (Figure 4c). The first peak of Ac–O is at 2.35 Å and at 2.65 Å for Ra–O with coordination numbers of 7.1 and 8.5, respectively. These structural positions and coordination numbers, especially for Ra2+ nanoconstructs, are found to be smaller than what is observed in Ra2+ structures in the chelator-stabilized crystal structure of Ra2+. (56) This is because these inorganic nanoconstructs are stabilized as part of a lanthanide inorganic lattice rather than through coordination within a chelator. Furthermore, the first peak height of Ac–O is more than double than that of the Ra–O first peak height, exhibiting stronger Ac atom agglomeration with shorter separation within the nanoconstruct compared to the Ra atoms that exhibit weaker aggregation with larger separation distances. This Ra atom structural behavior inside the cluster represents Ra atom distribution near the surface of the nanoconstructs. The higher coordination number in Ra–O (8.5) compared with that of the Ac–O (7.1) means that the Ra atoms are distributed in a broader coordination shell compared to the Ac atoms within the cluster. The same can be seen from the broader peak widths in Ra–O compared to Ac–O (Figure 4c). These structural results elucidate the fact that the Ra atoms tend to reside on the surface of the nanoconstructs, while the Ac atoms are distributed inside the core of the clusters. (57) From the structural analysis, we infer that the emitted α particle dose would be higher from the Ra-doped lanthanide due to the Ra distribution on the periphery of the nanoconstructs. However, a near-surface, peripheral distribution of the Ra can also be disadvantageous due to the release of decay daughters that would not occur in the core-placed Ac atom in the Ac-doped lanthanide nanoconstruct.
To understand the effect of a high level of radionuclide doping in lanthanide nanoconstructs, we investigated the structures of 25 mol % of Ba/Ra doping within the lanthanide nanoconstruct following the same simulation techniques. The structures at 67 °C from the simulation snapshots of these nanoconstructs are shown in Figure 5a–d. As seen before in 3 mol % systems, the Ra-doped nanoconstructs show smaller cluster sizes compared with Ba-doped nanoconstructs. The Ba–Ln nanoconstructs percolated all over the system, while the Ra-doped nanoconstructs show the distribution of a smaller cluster spanning the whole simulation box.

Figure 5

Figure 5. Structures of 25 mol % Ba/Ra doped lanthanide nanoconstructs. (a, b) Simulation snapshots of Ba- and Ra-doped nanoconstructs. (c, d) Closeup of Ba- and Ra-doped nanoclusters. Color scheme of the snapshots: Ba/Ra─blue, La─yellow, Eu─purple, V─green, and O─red. Water molecules are not shown for clarity. (e, f) RDF of metallic cations with the oxygen of vanadates for Ba- and Ra-doped nanoconstructs. (g) Comparison of RDFs between Ba- and Ra-doped nanoconstructs.

Ra atoms were observed to be preferentially distributed on the surface of the nanoconstructs, while both Ac and Ba clusters were located inside the cores of the nanoconstructs. The van der Waals radii of Ac, Ba, and Ra are 2.00 2.68, and 2.83 Å, respectively. This suggests that the largest Ra atoms may rearrange themselves on the nanoconstruct surface due to the excluded volume effect of the Ra atoms as observed in different structurally stable materials. (58,59) This explains why we consistently observed Ra first peak positions to be the highest compared to Ac and Ba from the oxygen atoms of vanadate, representing a higher agglomeration of Ra atoms compared to Ac and Ba.
The mean-square displacement (MSD) of the Ac and Ra atoms was monitored to obtain a dynamic perspective of the evolution of these nanoconstructs and the stability of the atoms within the nanoconstructs. The MSD was calculated from the last 1 ns production runs, and is defined by, MSD=1Ni=1N|ri(t)ri(0)|2, where ri(t) and ri(0) are the positions of the atoms at time t, and N is the total number of atoms. The MSDs are listed in Figure 6. For all the cases, a slower MSD of Ac/Ra (blue) and Eu (red) ions compared to La (black) was observed. The Ac/Ra and Eu atoms are electrostatically strongly bound to the nanoconstructs that move with the larger cluster, and hence slow motion (lower MSDs). On the contrary, while many of the La ions are strongly bound to the nanoconstruct, a large number of La atoms stay free in the solvent as observed in the simulation and also in experiments. The freely moving La atoms are responsible for the faster dynamics (larger MSD) for La atoms. For two different systems (Figure 6e), Ac–Ln and Ra–Ln, while Ac and Ra both hold tightly together inside the lanthanide nanoconstruct, Ra atoms show faster MSD values (blue) than Ac atoms (black), indicating faster motion in Ra atoms. The faster motion of the Ra atoms is due to their outer shell distribution that provides higher freedom of movement, as described in Figure 4, while the motions of the Ac atoms were adversely affected by their distribution inside the core of the nanoconstruct, which results in slower motion. The temperature dependence of the MSDs (Figure 6f) exhibits no substantial differences in La atom MSDs and the freedom of motion of free La atoms is not affected by lowering the temperature from 67 to 27 °C. However, electrostatically bound Ac/Ra and Eu atoms show slightly slower MSDs at 27 °C (solid line) than at 67 °C due to entropic effects. The evolution of these Ac–Ln and Ra–Ln nanoconstructs at high temperatures (T = 67 °C) and after cooling to 27 °C are shown in Figures S3 and S4 which, along with the dynamical data in Figure 6, reiterates the stability of these nanoconstructs upon cooling from high temperature as discussed in the structural analysis.

Figure 6

Figure 6. Mean-square-displacement (MSD) of La, Eu, and Ac/Ra of Ac-doped Ln (a,c) and Ra-doped Ln (b,d) compared with La and Eu atoms of the lanthanides. Low- and high-temperature MSDs are shown in (a), (b) at T = 27 °C, and (c), (d) at T = 67 °C respectively. A comparison of Eu (red and magenta), Ac (black), and Ra (blue) MSDs for Ac–Ln and Ra–Ln systems is shown in (e). Comparison with respect to temperature for these two systems is shown in (f); the solid lines are for T = 27 °C, and dashed lines are for T = 67 °C. MSD is lower at higher temperatures.

Conclusions

Click to copy section linkSection link copied!

In conclusion, this study has developed fundamental structural information for several different lanthanide nanoconstructs with three different radionuclides, actinium, barium, and radium, using experiments and large-scale MD simulations. Structural properties of lanthanide nanoconstructs were observed with different concentrations of Ac, Ba, and Ra as dopants. The detailed crystalline structure of nanoconstructs with Ra dopants is especially difficult to achieve experimentally because of the limited availability of radionuclides and the associated radiological hazards. This work showed by simulations the structural details and position of the Ac/Ba/Ra atoms within the LnVO4 nanoconstructs. The La0.9Eu0.1VO4 nanoconstructs doped with [225Ac]Ac3+ and [223Ra]Ra2+ were synthesized via precipitation in aqueous media. Encapsulation of 225Ac was determined by the preparation methods, i.e., nanoconstruct composition and crystal structure. The nanoconstruct composition had no significant effect on the encapsulation of [223Ra]Ra2+. The encapsulation of [223Ra]Ra2+ was however dependent on the reaction’s pH. Both LaVO4 and La0.9Eu0.1VO4 nanoconstructs showed additional emission peaks after doping with [225Ac]Ac3+. Molecular dynamics simulations show atomistic details of the structure–property relationship of Ac-, Ba-, and Ra-doped LnVO4 nanoconstructs at different mol % of dopant ions. The structural properties of Ac/Ba show that they tend to agglomerate inside the nanoconstructs, while the Ra atoms tend to arrange near the surface of the nanoconstructs. Similar structural properties were observed in the experiments with different mole percents of Ac and Ra dopant ions within the lanthanide nanoconstructs. Although experiments with radionuclides are difficult to perform in a wet lab setting because of the associated radioactive hazards, large-scale simulations, such as those presented in this work, can provide a priori structural details of radioactive LnVO4 nanoconstructs, which, in turn, can be corroborated with fewer experiments. Therefore, these experiments and simulations provide structural details and indicate the position of radionuclides within the LnVO4 nanoconstructs. This information is necessary to enhance the drug development process by aiding the design principles of future nanoconstructs for TRT. The application of targeted nanomedicines for cancer radiotherapy is currently evolving, and soon, will have the potential to become a critical component for cancer radiotherapy to treat drug-resistant and metastatic disease.

Methods

Click to copy section linkSection link copied!

Experimental Section

Materials

The chemicals lanthanum(III) chloride heptahydrate (LaCl3·7H2O; American Chemical Society reagent, Sigma-Aldrich, Chemical Abstracts Service [CAS] 10025-84-0), europium(III) chloride heptahydrate (EuCl3·6H2O; American Chemical Society reagent, Sigma-Aldrich, CAS 13759-92-7), barium nitrate (Ba(NO3)2; >99%, Acros Organics, CAS 10022-31-8), 6-aminohexanoic acid (6-AHA; >98.5%, Sigma-Aldrich, CAS 60-32-2), sodium orthovanadate (Na3VO4; 99%, Acros Organics, CAS 13721-39-6), and an ammonia solution (NH4OH; 28–30%, Sigma-Aldrich, CAS 1336-21-6) were used without further purification. Deionized water (18 Ω·cm) was obtained from a MilliporeSigma Milli-Q Ultrapure Water System.
The radionuclide [225Ac]Ac3+ (t1/2 = 9.9 days) was obtained from a thorium generator (228Th, 229Th, and 232Th) at the Oak Ridge National Laboratory after various chemical separation processes involving anion and cation exchange chromatography. (60) The radionuclide [223Ra]Ra2+ (t1/2 = 11.4 d) was obtained from 227Ac generated via the thermal neutron irradiation of a 226Ra target following the processing and purification steps described elsewhere. (22) Both [225Ac]Ac3+ and [223Ra]Ra2+ were dissolved in 0.1 M HNO3 for the nanoconstruct synthesis.

Nanoparticle Synthesis

Lanthanide vanadate (La(1–x)EuxVO4) nanoconstructs were synthesized by adapting the procedures reported by Huignard et al. (40,61) Different Eu3+ molar concentrations (e.g., Eu3+ = 0, 2, 5, 10, 25, 50, 75, and 100%) were used to synthesize La(1–x)EuxVO4 nanoconstructs and evaluate their fluorescent properties, size distribution, and crystal structure. Similarly, (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs were synthesized with different Ba2+ molar concentrations (e.g., Ba2+ = 0, 5, 10, 25, and 50%) as a surrogate for [223Ra]Ra2+. The La(1–x)EuxVO4 and (La0.9Eu0.1)(1–y)BayVO4 were synthesized by precipitation of metal salts (i.e., LaCl3, EuCl3, and Ba(NO3)2) with Na3VO4 in the presence of 6-AHA to control their size and to increase their stability in water. For synthesis, metal salts, 6-Aminohexaonic Acid (6-AHA), and Na3VO4 were dissolved in deionized (DI) water at a concentration of 0.1 M, mixed in a 1:0.75:0.75 molar ratio, and then heated at 90 °C for 30 min. Nanoconstruct suspensions were washed at least three times by centrifugation (14,000–15,000 rpm, 15 min) and resuspended in deionized water. The La(1–x)EuxVO4 and (La0.9Eu0.1)(1–y)BayVO4 doped with 225Ac or 223Ra were prepared following the same steps described previously after mixing the metal salts with each radionuclide solution. The precipitation of LnVO4 nanoconstructs without 6-AHA under basic conditions and at room temperature was also explored. (40)

Characterization

The La(1–x)EuxVO4 and (La0.9Eu0.1)(1–y)BayVO4 nanoconstruct suspensions were diluted in DI water to assess their particle size distribution and fluorescence properties. Particle size distribution was obtained using DLS in a Malvern Zetasizer Nano ZS. The reported hydrodynamic size (Zave) and polydispersity index (PdI) correspond to the mean of three runs with at least 10 measurements, each as defined in the standard operating procedure. Excitation and emission spectra were collected using a Fluorolog-QM spectrometer (Horiba) by diluting nanoconstruct suspensions in DI water. The excitation was measured with a 100-W pulsed xenon lamp as the source. The morphology of La0.9Eu0.1VO4 nanoconstructs was characterized by TEM using a JEOL NEOARM TEM/STEM. A diluted nanoconstruct suspension was drop-cast onto a carbon-Formvar Cu grid and evaporated to dryness at room temperature. The La(1–x)EuxVO4 and (La0.9Eu0.1)(1–y)BayVO4 nanoconstruct suspensions were precipitated by centrifugation and then evaporated to dryness in a muffle furnace at 60 °C. Nanoconstruct precipitates were ground using a mortar and pestle to obtain a fine powder. These powders were characterized using a PANanalytical X’Pert Pro MPD X-ray diffractometer. Rietveld analysis of the XRD data was performed using GSAS II software. (62,63) Raman spectra were recorded by using a Renishaw in Via Qontor confocal Raman microscope with a 532 nm laser. The sample was analyzed at an optical magnification of 20×. Radionuclide activities were determined by γ-ray spectroscopy using a high-purity germanium (HPGe) detector coupled to a PC-based multichannel analyzer (Canberra Industries, Meriden, CT). The HPGe has a crystal active volume of approximately 100 cm3 and a beryllium window. Energy and efficiency calibrations were determined by γ-ray sources traceable to the National Institute of Standards and Technology. The encapsulation efficiency of 225Ac and 223Ra was determined by measuring the nanoconstruct suspension and supernatant after washing them as described previously.

Molecular Dynamics Simulations

The atomistic MD simulations are performed using the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) molecular dynamics package. (64) The initial system was built using randomly distributed individual lanthanum (La), europium (Eu), and vanadate (VO4) ions in a box of 40 nm3 using the Packmol package. (65) The first set of simulations was developed to obtain the structural properties of pure lanthanides (Ln) without the addition of any radionuclides as a control system to compare directly with experiments. Subsequently, three different systems were studied to understand the atomistic structures of Ac, Ba, and Ra with respect to the La, Eu, and O (of VO4) within the lanthanide nanoclusters. The whole system was placed in an aqueous environment within a box of water with a density of 1 g/cm3. This creates three different systems of dilute solutions of lanthanide salt atoms with three different radioactive dopant ions: Ac, Ba, and Ra.
The interaction potentials or force field (FF) parameters between different atomic species were used as follows. For the lanthanide cations (La, Eu), additive Chemistry at Harvard Molecular Mechanics (CHARMM) FF was used to model van der Waals interactions because it shows excellent agreement with experiments. (66) The interaction of VO4 ions in the solvation state was modeled using AMBER FF following past work that has shown excellent agreement with experiments. (67) For the water molecules, the SPC/E water model was used for all the simulations. Electrostatic interactions between all charged ions and partially charged oxygen and hydrogen of water were used. The randomly generated initial system was generated at 0 °C as shown in Figure 7a. The temperature was subsequently increased to the experimental temperature, 67 °C. At T = 67 °C the system was equilibrated for 2 ns to form nanoclusters, as shown in Figure 7b. Once the system was equilibrated at a high temperature, the system was then cooled to ambient temperature, T = 27 °C. The final configuration of the cooled state is shown in Figure 7c. This result compares well with those of lanthanide clusters observed in TEM experiments (Figure 7d). For a pristine lanthanide nanoconstruct suspension, the structures of different ionic species with respect to their counterpart and the structure of water molecules are investigated using radial distribution functions (RDFs), as shown in Figure S1. The oxygen–oxygen, oxygen–hydrogen, and hydrogen–hydrogen RDFs (68) and lanthanide ions RDFs suggest that the performance of the model system and FF parameters are overall excellent to provide confidence in simulation data. The radionuclides, Ac, Ra, and Ba, interactions with water and lanthanides are modeled using experimentally verified Lennard–Jones (LJ) potential, and the electrostatic interactions of the charged particles are modeled using Coulomb potential. (69) The Ac/Ra/Ba LJ interaction parameters are tabulated in Table S1. For further details of the simulation methods, please see the Supporting Information (SI).

Figure 7

Figure 7. Lanthanide vanadate simulations in an aqueous environment. (a) Initial system setup containing La, Eu, and VO4 ions in g/cm3 water randomly distributed in a box of 40 nm3. (b) 3D structure of the Ln nanoconstruct at experimental temperature, 67 °C. Water molecules are not shown for clarity. (c) Top view (2D) snapshot of the lanthanide nanoconstruct as can be observed from the TEM of experimentally prepared nanoconstructs. The background dots are neutral La atoms that stay mostly in water in a solvated state. (d) TEM nanoconstruct of lanthanide vanadate showing a similar structure to that derived in (c).

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnano.3c13213.

  • Description of MD simulations in detail; force-field parameters (Lennard-Jones) used in the MD simulations; RDF of water in LnVO4 and the comparison of the RDF of La, Eu, and oxygen of VO4; RDF of Ac-/Ra-doped lanthanide at different temperatures and the snapshots; Rietveld analysis parameters and XRD patterns; summary of the molar ratio of the nanoconstructs; excitation and emission spectra of 225Ac-doped nanoconstructs and 30 days' decayed 225Ac-doped nanoconstructs, respectively; and different compositions of lanthanides (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Monojoy Goswami - Chemical Sciences Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United StatesOrcidhttps://orcid.org/0000-0002-4473-4888
    • Miguel Toro-González - Radioisotope Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States
    • Jisue Moon - Radioisotope Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

This research was supported by the US Department of Energy’s (DOE’s) Oak Ridge National Laboratory Directed Research & Development (LDRD) Program. The isotopes used in this research was supplied by the U.S. Department of Energy Isotope Program, managed by the Office of Isotope R&D and Production. The MD simulations used resources of the Oak Ridge Leadership Computing Facility at Oak Ridge National Laboratory, which is supported by the DOE Office of Science of the U.S. DOE under Contract No. DE-AC05-00OR22725. Part of the MD simulations used resources of the National Energy Research Scientific Computing Center (NERSC), a DOE Office of Scientific User Facility supported by the DOE Office of Science under Contract No. DE-AC02- 05CH11231. The authors thank Christopher Orosco from ORNL’s Creative Services Communications Division for designing the Supplementary Cover Image for this work.

Abbreviations

Click to copy section linkSection link copied!

6-AHA

6-aminohexanoic acid

CAS

chemical abstracts service

CTB

charge-transfer band

DI

deionized

DLS

dynamic light scattering

FF

force field

MD

molecular dynamics

RDF

radial distribution function

TEM

transmission electron microscopy

TRT

targeted radionuclide therapy

XRD

X-ray diffraction

References

Click to copy section linkSection link copied!

This article references 69 other publications.

  1. 1
    Kim, Y. S.; Brechbiel, M. W. An Overview of Targeted Alpha Therapy. Tumor Biol. 2012, 33, 573590,  DOI: 10.1007/s13277-011-0286-y
  2. 2
    Pouget, J. P.; Constanzo, J. Revisiting the Radiobiology of Targeted Alpha Therapy. Front. Med. 2021, 8, 692436  DOI: 10.3389/fmed.2021.692436
  3. 3
    Canter, B. S.; Leung, C. N.; Christopher Fritton, J.; Bäck, T.; Rajon, D.; Azzam, E. I.; Howell, R. W. Radium-223-Induced Bystander Effects Cause DNA Damage and Apoptosis in Disseminated Tumor Cells in Bone Marrow. Mol. Cancer Res. 2021, 19, 17391750,  DOI: 10.1158/1541-7786.MCR-21-0005
  4. 4
    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.; Morse, D. L. Development of Targeted Alpha Particle Therapy for Solid Tumors. Molecules 2019, 24, 4314,  DOI: 10.3390/molecules24234314
  5. 5
    Ballangrud, Å. M.; Yang, W. H.; Palm, S.; Enmon, R.; Borchardt, P. E.; Pellegrini, V. A.; McDevitt, M. R.; Scheinberg, D. A.; Sgouros, G. Alpha-Particle Emitting Atomic Generator (Actinium-225)-Labeled Trastuzumab (Herceptin) Targeting of Breast Cancer SpheroidsEfficacy versus HER2/Neu Expression. Clin. Cancer Res. 2004, 10, 44894497,  DOI: 10.1158/1078-0432.CCR-03-0800
  6. 6
    Allen, B. J.; Huang, C.-Y.; Clarke, R. A. Targeted Alpha Anticancer Therapies: Update and Future Prospects. Biol. Targets Ther. 2014, 8, 255267,  DOI: 10.2147/BTT.S29947
  7. 7
    Feuerecker, B.; Biechl, P.; Seidl, C.; Bruchertseifer, F.; Morgenstern, A.; Schwaiger, M.; Eisenreich, W. Diverse Metabolic Response of Cancer Cells Treated with a 213 Bi-Anti-EGFR-Immunoconjugate. Sci. Rep. 2021, 11, 6227,  DOI: 10.1038/s41598-021-84421-4
  8. 8
    Song, H.; Hobbs, R. F.; Vajravelu, R.; Huso, D. L.; Esaias, C.; Apostolidis, C.; Morgenstern, A.; Sgouros, G. Radioimmunotherapy of Breast Cancer Metastases with α-Particle Emitter 225Ac: Comparing Efficacy with 213 Bi and 90Y. Cancer Res. 2009, 69, 89418948,  DOI: 10.1158/0008-5472.CAN-09-1828
  9. 9
    Eychenne, R.; Chérel, M.; Haddad, F.; Guérard, F.; Gestin, J. F. Overview of the Most Promising Radionuclides for Targeted Alpha Therapy: The “Hopeful Eight.. Pharmaceutics 2021, 13, 906,  DOI: 10.3390/pharmaceutics13060906
  10. 10
    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, 321336,  DOI: 10.3390/ph8020321
  11. 11
    Merkx, R. I. J.; Rijpkema, M.; Franssen, G. M.; Kip, A.; Smeets, B.; Morgenstern, A.; Bruchertseifer, F.; Yan, E.; Wheatcroft, M. P.; Oosterwijk, E.; Mulders, P. F. A.; Heskamp, S. Carbonic Anhydrase IX-Targeted α-Radionuclide Therapy with 225Ac Inhibits Tumor Growth in a Renal Cell Carcinoma Model. Pharmaceuticals 2022, 15, 570,  DOI: 10.3390/ph15050570
  12. 12
    Membreno, R.; Cook, B. E.; Zeglis, B. M. Pretargeted Radioimmunotherapy Based on the Inverse Electron Demand Diels-Alder Reaction. J. Vis. Exp. 2019, 2019, 59041  DOI: 10.3791/59041
  13. 13
    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  DOI: 10.1063/1.5096880
  14. 14
    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, 967977,  DOI: 10.1515/ract-2019-3206
  15. 15
    Tao, Y.; Sun, Y.; Shi, K.; Pei, P.; Ge, F.; Yang, K.; Liu, T. Versatile Labeling of Multiple Radionuclides onto a Nanoscale Metal–Organic Framework for Tumor Imaging and Radioisotope Therapy. Biomater. Sci. 2021, 9, 29472954,  DOI: 10.1039/D0BM02225J
  16. 16
    Tanahashi, K.; Mikos, A. G. Effect of Hydrophilicity and Agmatine Modification on Degradation of Poly(Propylene Fumarate-Co-Ethylene Glycol) Hydrogels. J. Biomed. Mater. Res. Part A 2003, 67A, 11481154,  DOI: 10.1002/jbm.a.10147
  17. 17
    Sanità, G.; Carrese, B.; Lamberti, A. Nanoparticle Surface Functionalization: How to Improve Biocompatibility and Cellular Internalization. Front. Mol. Biosci. 2020, 7, 587012  DOI: 10.3389/fmolb.2020.587012
  18. 18
    VanDyke, D.; Kyriacopulos, P.; Yassini, B.; Wright, A.; Burkhart, E.; Jacek, S.; Pratt, M.; Peterson, C. R.; Rai, P. Nanoparticle Based Combination Treatments for Targeting Multiple Hallmarks of Cancer. Int. J. nano Stud. Technol. 2016, 118,  DOI: 10.19070/2167-8685-SI04001
  19. 19
    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, 766776,  DOI: 10.1021/bc100574f
  20. 20
    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, 614620,  DOI: 10.1016/j.nucmedbio.2015.03.007
  21. 21
    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, 595600,  DOI: 10.1524/ract.2013.2066
  22. 22
    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, 3350,  DOI: 10.3390/jnt2010003
  23. 23
    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, 97449755,  DOI: 10.1039/D0NR01172J
  24. 24
    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, 79857997,  DOI: 10.1039/C8TB02173B
  25. 25
    Kanematsu, N.; Inaniwa, T.; Koba, Y. Relationship between Electron Density and Effective Densities of Body Tissues for Stopping, Scattering, and Nuclear Interactions of Proton and Ion Beams. Med. Phys. 2012, 39, 10161020,  DOI: 10.1118/1.3679339
  26. 26
    Greaves, G.; Hinks, J. A.; Busby, P.; Mellors, N. J.; Ilinov, A.; Kuronen, A.; Nordlund, K.; Donnelly, S. E. Enhanced Sputtering Yields from Single-Ion Impacts on Gold Nanorods. Phys. Rev. Lett. 2013, 111, 065504  DOI: 10.1103/PhysRevLett.111.065504
  27. 27
    Holzwarth, U.; Ojea Jimenez, I.; 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,  DOI: 10.1186/s41181-018-0042-3
  28. 28
    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, 473481,  DOI: 10.1524/ract.2012.1935
  29. 29
    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, 183189,  DOI: 10.1016/j.apradiso.2017.07.030
  30. 30
    Gangwar, P.; Pandey, M.; Sivakumar, S.; Pala, R. G. S.; Parthasarathy, G. Increased Loading of Eu3+ Ions in Monazite LaVO4 Nanocrystals via Pressure-Driven Phase Transitions. Cryst. Growth Des. 2013, 13, 23442349,  DOI: 10.1021/cg3018908
  31. 31
    Jia, C. N.; Sun, L. D.; Yan, Z. G.; Pang, Y. C.; Lü, S. Z.; Yan, C. H. Monazite and Zircon Type LaVO4:Eu Nanocrystals – Synthesis, Luminescent Properties, and Spectroscopic Identification of the Eu3+ Sites. Eur. J. Inorg. Chem. 2010, 2010, 26262635,  DOI: 10.1002/ejic.201000038
  32. 32
    Nuñez, N. O.; Zambrano, P.; García-Sevillano, J.; Cantelar, E.; Rivera-Fernández, S.; De La Fuente, J. M.; Ocaña, M. Uniform Poly(Acrylic Acid)-Functionalized Lanthanide-Doped LaVO4 Nanophosphors with High Colloidal Stability and Biocompatibility. Eur. J. Inorg. Chem. 2015, 2015, 45464554,  DOI: 10.1002/ejic.201500265
  33. 33
    Cheng, X.; Guo, D.; Feng, S.; Yang, K.; Wang, Y.; Ren, Y.; Song, Y. Structure and Stability of Monazite- and Zircon-Type LaVO4 under Hydrostatic Pressure. Opt. Mater. (Amst). 2015, 49, 3238,  DOI: 10.1016/j.optmat.2015.08.011
  34. 34
    Oka, Y.; Yao, T.; Yamamoto, N. Hydrothermal Synthesis of Lanthanum Vanadates: Synthesis and Crystal Structures of Zircon-Type LaVO4 and a New Compound LaV3O9. J. Solid State Chem. 2000, 152, 486491,  DOI: 10.1006/jssc.2000.8717
  35. 35
    Ropp, R. C.; Carroll, B. Precipitation of Rare Earth Vanadates from Aqueous Solution. J. Inorg. Nucl. Chem. 1977, 39, 13031307,  DOI: 10.1016/0022-1902(77)80286-8
  36. 36
    Xu, Z.; Li, C.; Hou, Z.; Peng, C.; Lin, J. Morphological Control and Luminescence Properties of Lanthanide Orthovanadate LnVO4 (Ln = La to Lu) Nano-/Microcrystals Viahydrothermal Process. CrystEngComm 2011, 13, 474482,  DOI: 10.1039/C0CE00161A
  37. 37
    Lotfi, S.; El Ouardi, M.; Ait Ahsaine, H.; Madigou, V.; BaQais, A.; Assani, A.; Saadi, M.; Arab, M. Low-Temperature Synthesis, Characterization and Photocatalytic Properties of Lanthanum Vanadate LaVO4. Heliyon 2023, 9, e17255  DOI: 10.1016/j.heliyon.2023.e17255
  38. 38
    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. 2015, 73, 133140,  DOI: 10.1007/s10971-014-3504-4
  39. 39
    Tian, L.; Mho, S. Enhanced Photoluminescence of YVO4:Eu3+ by Codoping the Sr2+, Ba2+ or Pb2+ Ion. J. Lumin. 2007, 122–123, 99103,  DOI: 10.1016/j.jlumin.2006.01.108
  40. 40
    Huignard, A.; Gacoin, T.; Boilot, J. P. Synthesis and Luminescence Properties of Colloidal YVO4:Eu Phosphors. Chem. Mater. 2000, 12, 10901094,  DOI: 10.1021/cm990722t
  41. 41
    Deblonde, G. J. P.; Zavarin, M.; Kersting, A. B. The Coordination Properties and Ionic Radius of Actinium: A 120-Year-Old Enigma. Coord. Chem. Rev. 2021, 446, 214130  DOI: 10.1016/j.ccr.2021.214130
  42. 42
    Su, J.; Mi, X.; Sun, J.; Yang, L.; Hui, C.; Lu, L.; Bai, Z.; Zhang, X. Tunable Luminescence and Energy Transfer Properties in YVO4:Bi3+, Eu3+ Phosphors. J. Mater. Sci. 2017, 52, 782792,  DOI: 10.1007/s10853-016-0375-9
  43. 43
    Rapp, M.; Lozano, Y.; Fernández-Ramos, M.; Isasi, J.; Palafox, M. A. Superparamagnetic and Light-Emitting Bifunctional Nanocomposites of Iron Oxide and Erbium or Thulium Doped Yttrium Orthovanadate. J. Alloys Compd. 2022, 929, 167065  DOI: 10.1016/j.jallcom.2022.167065
  44. 44
    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, 22112220,  DOI: 10.1002/ejic.201001341
  45. 45
    Wang, F.; Yu, L.; Zhu, Y.; Zhu, Z.; Meng, X.; Lv, Y.; Peng, S.; Yang, L. Defect Control and Optical Performance of Yttrium Orthovanadate Nanocrystals via a Facile PH-Sensitive Synthesis. J. Alloys Compd. 2023, 968, 172259  DOI: 10.1016/j.jallcom.2023.172259
  46. 46
    Ling-Hu, P.; Guo, X.; Hu, J.; Deng, C.; Cui, R. Anomalous 5D0→7F4 Transition of Eu3+-Doped BaLaGaO4 Phosphors for WLEDs and Plant Growth Applications. Adv. Opt. Mater. 2024, 12, 2301760  DOI: 10.1002/adom.202301760
  47. 47
    Chi, F.; Wei, X.; Zhou, S.; Chen, Y.; Duan, C.; Yin, M. Enhanced 5D0 → 7F4 Transition and Optical Thermometry of Garnet Type Ca3Ga2Ge3O12:Eu3+ Phosphors. Inorg. Chem. Front. 2018, 5, 12881293,  DOI: 10.1039/C8QI00199E
  48. 48
    Shannon, R. D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr., Sect. A 1976, 32, 751767,  DOI: 10.1107/S0567739476001551
  49. 49
    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, 339346,  DOI: 10.1134/S1066362219030093
  50. 50
    Kozempel, J.; Vlk, M.; Málková, E.; Bajzíková, A.; Bárta, J.; Santos-Oliveira, R.; Malta Rossi, A. Prospective Carriers of 223Ra for Targeted Alpha Particle Therapy. J. Radioanal. Nucl. Chem. 2015, 304, 443447,  DOI: 10.1007/s10967-014-3615-y
  51. 51
    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. Nanomater. 2020, 10, 1632,  DOI: 10.3390/nano10091632
  52. 52
    Suchánková, P.; Kukleva, E.; Štamberg, K.; Nykl, P.; Sakmár, M.; Vlk, M.; Kozempel, J. Determination, Modeling and Evaluation of Kinetics of 223Ra Sorption on Hydroxyapatite and Titanium Dioxide Nanoparticles. Mater. 2020, 13, 1915,  DOI: 10.3390/ma13081915
  53. 53
    Kukleva, E.; Suchánková, P.; Štamberg, K.; Vlk, M.; Šlouf, M.; Kozempel, J. Surface Protolytic Property Characterization of Hydroxyapatite and Titanium Dioxide Nanoparticles. RSC Adv. 2019, 9, 2198921995,  DOI: 10.1039/C9RA03698A
  54. 54
    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, 36593666,  DOI: 10.1039/C9RA08953E
  55. 55
    Muckley, E. S.; Aytug, T.; Mayes, R.; Lupini, A. R.; Carrillo, J. M. Y.; Goswami, M.; Sumpter, B. G.; Ivanov, I. N. Hierarchical TiO2:Cu2O Nanostructures for Gas/Vapor Sensing and CO2 Sequestration. ACS Appl. Mater. Interfaces 2019, 11, 4846648475,  DOI: 10.1021/acsami.9b18824
  56. 56
    White, F. D.; Thiele, N. A.; Simms, M. E.; Cary, S. K. Structure and Bonding of a Radium Coordination Compound in the Solid State. Nat. Chem. 2024, 16, 168172,  DOI: 10.1038/s41557-023-01366-z
  57. 57
    Akamo, D. O.; Kumar, N.; Li, Y.; Pekol, C.; Li, K.; Goswami, M.; Hirschey, J.; LaClair, T. J.; Keffer, D. J.; Rios, O.; Gluesenkamp, K. R. Stabilization of Low-Cost Phase Change Materials for Thermal Energy Storage Applications. iScience 2023, 26, 107175  DOI: 10.1016/j.isci.2023.107175
  58. 58
    Vives, S.; Ramel, D.; Meunier, C. Evolution of the Structure with the Composition and the Defect Arrangement in the Gadolinium and Samarium Doped and Co-Doped Ceria Systems: A Molecular Dynamics Study. Solid State Ionics 2021, 364, 115611  DOI: 10.1016/j.ssi.2021.115611
  59. 59
    Li, P.; Merz, K. M. Metal Ion Modeling Using Classical Mechanics. Chem. Rev. 2017, 117, 15641686,  DOI: 10.1021/acs.chemrev.6b00440
  60. 60
    Boll, R. A.; Malkemus, D.; Mirzadeh, S. Production of Actinium-225 for Alpha Particle Mediated Radioimmunotherapy. Appl. Radiat. Isot. 2005, 62, 667679,  DOI: 10.1016/j.apradiso.2004.12.003
  61. 61
    Huignard, A.; Buissette, V.; Laurent, G.; Gacoin, T.; Boilot, J. P. Synthesis and Characterizations of YVO4: Eu Colloids. Chem. Mater. 2002, 14, 22642269,  DOI: 10.1021/cm011263a
  62. 62
    Young, R. A.. The Rietveld Method; Oxford University Press, 1981.
  63. 63
    Toby, B. H.; Von Dreele, R. B. GSAS-II: The Genesis of a Modern Open-Source All Purpose Crystallography Software Package. J. Appl. Crystallogr. 2013, 46, 544549,  DOI: 10.1107/S0021889813003531
  64. 64
    Thompson, A. P.; Aktulga, H. M.; Berger, R.; Bolintineanu, D. S.; Brown, W. M.; Crozier, P. S.; in’t Veld, P. J.; Kohlmeyer, A.; Moore, S. G.; Nguyen, T. D.; Shan, R.; Stevens, M. J.; Tranchida, J.; Trott, C.; Plimpton, S. J. LAMMPS - a Flexible Simulation Tool for Particle-Based Materials Modeling at the Atomic, Meso, and Continuum Scales. Comput. Phys. Commun. 2022, 271, 108171  DOI: 10.1016/j.cpc.2021.108171
  65. 65
    Martinez, L.; Andrade, R.; Birgin, E. G.; Martínez, J. M. PACKMOL: A Package for Building Initial Configurations for Molecular Dynamics Simulations. J. Comput. Chem. 2009, 30, 21572164,  DOI: 10.1002/jcc.21224
  66. 66
    Qiao, B.; Skanthakumar, S.; Soderholm, L. Comparative CHARMM and AMOEBA Simulations of Lanthanide Hydration Energetics and Experimental Aqueous-Solution Structures. J. Chem. Theory Comput. 2018, 14, 17811790,  DOI: 10.1021/acs.jctc.7b01018
  67. 67
    Priest, C.; Zhou, J.; Jiang, D.; en Solvation of the Vanadate Ion in Seawater Conditions from Molecular Dynamics Simulations. Inorg. Chim. Acta 2017, 458, 3944,  DOI: 10.1016/j.ica.2016.12.027
  68. 68
    Clark, G. N. I.; Cappa, C. D.; Smith, J. D.; Saykally, R. J.; Head-Gordon, T. The Structure of Ambient Water. Mol. Phys. 2010, 108, 14151433,  DOI: 10.1080/00268971003762134
  69. 69
    Goswami, M.; Kumar, N.; Li, Y.; Hirschey, J.; LaClair, T. J.; Akamo, D. O.; Sultan, S.; Rios, O.; Gluesenkamp, K. R.; Graham, S. Understanding Supercooling Mechanism in Sodium Sulfate Decahydrate Phase-Change Material. J. Appl. Phys. 2021, 129, 245109  DOI: 10.1063/5.0049512

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Open PDF

ACS Nano

Cite this: ACS Nano 2024, 18, 26, 16577–16588
Click to copy citationCitation copied!
https://doi.org/10.1021/acsnano.3c13213
Published June 17, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

929

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a, d) Diffraction patterns, (b, e) Raman spectra, and (e, f) size distribution of (a–c) La(1–x)EuxVO4 and (d–f) (La0.9Eu0.1)(1–y)BayVO4 of nanoconstructs.

    Figure 2

    Figure 2. Excitation spectra, emission spectra, and emission peak ratio of (a–c) La(1–x)EuxVO4 and (d–f) (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs.

    Figure 3

    Figure 3. (a, b) Emission spectra of La0.9Eu0.1VO4, La0.9Eu0.1[225Ac]AcVO4, and decayed La0.9Eu0.1[225Ac]AcVO4 nanoconstructs. The emission spectra were normalized to the 618 nm peak corresponding to the Eu3+5D07F2 transition. (c) Emission spectra of La[225Ac]AcVO4 and decayed La[225Ac]VO4 nanoconstructs. The emission spectra were normalized to the 450 nm peak corresponding to the [VO4]3–3T11A1 and 3T2lA1 energy transfer. (d) Encapsulation efficiency of [223Ra]Ra2+ within (La0.9Eu0.1)(1–y)BayVO4 nanoconstructs.

    Figure 4

    Figure 4. Structural details of Ac–Ln and Ra–Ln nanoconstructs at 3 mol % of Ac/Ra doping at 67 °C. Three different RDFs, Ac/Ra–O, La–O, and Eu–O are shown for the (a) Ac–Ln system and (b) Ra–Ln system. The comparison between Ac–O and Ra–O are shown in (c). Colors are represented in the legends. (d, e) Simulation snapshots of Ac–O and Ra–O systems, respectively. Water molecules are not shown for clarity. (f, g) Closeup of the nanoconstructs for Ac–O and Ra–O systems, respectively. Color scheme in the snapshots: Ac/Ra─blue, La─cyan, Eu─yellow, V─green, and O─red. The low-temperature, T = 27 °C, RDFs are compared in Figure S2 showing structural stability of the nanoconstructs.

    Figure 5

    Figure 5. Structures of 25 mol % Ba/Ra doped lanthanide nanoconstructs. (a, b) Simulation snapshots of Ba- and Ra-doped nanoconstructs. (c, d) Closeup of Ba- and Ra-doped nanoclusters. Color scheme of the snapshots: Ba/Ra─blue, La─yellow, Eu─purple, V─green, and O─red. Water molecules are not shown for clarity. (e, f) RDF of metallic cations with the oxygen of vanadates for Ba- and Ra-doped nanoconstructs. (g) Comparison of RDFs between Ba- and Ra-doped nanoconstructs.

    Figure 6

    Figure 6. Mean-square-displacement (MSD) of La, Eu, and Ac/Ra of Ac-doped Ln (a,c) and Ra-doped Ln (b,d) compared with La and Eu atoms of the lanthanides. Low- and high-temperature MSDs are shown in (a), (b) at T = 27 °C, and (c), (d) at T = 67 °C respectively. A comparison of Eu (red and magenta), Ac (black), and Ra (blue) MSDs for Ac–Ln and Ra–Ln systems is shown in (e). Comparison with respect to temperature for these two systems is shown in (f); the solid lines are for T = 27 °C, and dashed lines are for T = 67 °C. MSD is lower at higher temperatures.

    Figure 7

    Figure 7. Lanthanide vanadate simulations in an aqueous environment. (a) Initial system setup containing La, Eu, and VO4 ions in g/cm3 water randomly distributed in a box of 40 nm3. (b) 3D structure of the Ln nanoconstruct at experimental temperature, 67 °C. Water molecules are not shown for clarity. (c) Top view (2D) snapshot of the lanthanide nanoconstruct as can be observed from the TEM of experimentally prepared nanoconstructs. The background dots are neutral La atoms that stay mostly in water in a solvated state. (d) TEM nanoconstruct of lanthanide vanadate showing a similar structure to that derived in (c).

  • References


    This article references 69 other publications.

    1. 1
      Kim, Y. S.; Brechbiel, M. W. An Overview of Targeted Alpha Therapy. Tumor Biol. 2012, 33, 573590,  DOI: 10.1007/s13277-011-0286-y
    2. 2
      Pouget, J. P.; Constanzo, J. Revisiting the Radiobiology of Targeted Alpha Therapy. Front. Med. 2021, 8, 692436  DOI: 10.3389/fmed.2021.692436
    3. 3
      Canter, B. S.; Leung, C. N.; Christopher Fritton, J.; Bäck, T.; Rajon, D.; Azzam, E. I.; Howell, R. W. Radium-223-Induced Bystander Effects Cause DNA Damage and Apoptosis in Disseminated Tumor Cells in Bone Marrow. Mol. Cancer Res. 2021, 19, 17391750,  DOI: 10.1158/1541-7786.MCR-21-0005
    4. 4
      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.; Morse, D. L. Development of Targeted Alpha Particle Therapy for Solid Tumors. Molecules 2019, 24, 4314,  DOI: 10.3390/molecules24234314
    5. 5
      Ballangrud, Å. M.; Yang, W. H.; Palm, S.; Enmon, R.; Borchardt, P. E.; Pellegrini, V. A.; McDevitt, M. R.; Scheinberg, D. A.; Sgouros, G. Alpha-Particle Emitting Atomic Generator (Actinium-225)-Labeled Trastuzumab (Herceptin) Targeting of Breast Cancer SpheroidsEfficacy versus HER2/Neu Expression. Clin. Cancer Res. 2004, 10, 44894497,  DOI: 10.1158/1078-0432.CCR-03-0800
    6. 6
      Allen, B. J.; Huang, C.-Y.; Clarke, R. A. Targeted Alpha Anticancer Therapies: Update and Future Prospects. Biol. Targets Ther. 2014, 8, 255267,  DOI: 10.2147/BTT.S29947
    7. 7
      Feuerecker, B.; Biechl, P.; Seidl, C.; Bruchertseifer, F.; Morgenstern, A.; Schwaiger, M.; Eisenreich, W. Diverse Metabolic Response of Cancer Cells Treated with a 213 Bi-Anti-EGFR-Immunoconjugate. Sci. Rep. 2021, 11, 6227,  DOI: 10.1038/s41598-021-84421-4
    8. 8
      Song, H.; Hobbs, R. F.; Vajravelu, R.; Huso, D. L.; Esaias, C.; Apostolidis, C.; Morgenstern, A.; Sgouros, G. Radioimmunotherapy of Breast Cancer Metastases with α-Particle Emitter 225Ac: Comparing Efficacy with 213 Bi and 90Y. Cancer Res. 2009, 69, 89418948,  DOI: 10.1158/0008-5472.CAN-09-1828
    9. 9
      Eychenne, R.; Chérel, M.; Haddad, F.; Guérard, F.; Gestin, J. F. Overview of the Most Promising Radionuclides for Targeted Alpha Therapy: The “Hopeful Eight.. Pharmaceutics 2021, 13, 906,  DOI: 10.3390/pharmaceutics13060906
    10. 10
      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, 321336,  DOI: 10.3390/ph8020321
    11. 11
      Merkx, R. I. J.; Rijpkema, M.; Franssen, G. M.; Kip, A.; Smeets, B.; Morgenstern, A.; Bruchertseifer, F.; Yan, E.; Wheatcroft, M. P.; Oosterwijk, E.; Mulders, P. F. A.; Heskamp, S. Carbonic Anhydrase IX-Targeted α-Radionuclide Therapy with 225Ac Inhibits Tumor Growth in a Renal Cell Carcinoma Model. Pharmaceuticals 2022, 15, 570,  DOI: 10.3390/ph15050570
    12. 12
      Membreno, R.; Cook, B. E.; Zeglis, B. M. Pretargeted Radioimmunotherapy Based on the Inverse Electron Demand Diels-Alder Reaction. J. Vis. Exp. 2019, 2019, 59041  DOI: 10.3791/59041
    13. 13
      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  DOI: 10.1063/1.5096880
    14. 14
      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, 967977,  DOI: 10.1515/ract-2019-3206
    15. 15
      Tao, Y.; Sun, Y.; Shi, K.; Pei, P.; Ge, F.; Yang, K.; Liu, T. Versatile Labeling of Multiple Radionuclides onto a Nanoscale Metal–Organic Framework for Tumor Imaging and Radioisotope Therapy. Biomater. Sci. 2021, 9, 29472954,  DOI: 10.1039/D0BM02225J
    16. 16
      Tanahashi, K.; Mikos, A. G. Effect of Hydrophilicity and Agmatine Modification on Degradation of Poly(Propylene Fumarate-Co-Ethylene Glycol) Hydrogels. J. Biomed. Mater. Res. Part A 2003, 67A, 11481154,  DOI: 10.1002/jbm.a.10147
    17. 17
      Sanità, G.; Carrese, B.; Lamberti, A. Nanoparticle Surface Functionalization: How to Improve Biocompatibility and Cellular Internalization. Front. Mol. Biosci. 2020, 7, 587012  DOI: 10.3389/fmolb.2020.587012
    18. 18
      VanDyke, D.; Kyriacopulos, P.; Yassini, B.; Wright, A.; Burkhart, E.; Jacek, S.; Pratt, M.; Peterson, C. R.; Rai, P. Nanoparticle Based Combination Treatments for Targeting Multiple Hallmarks of Cancer. Int. J. nano Stud. Technol. 2016, 118,  DOI: 10.19070/2167-8685-SI04001
    19. 19
      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, 766776,  DOI: 10.1021/bc100574f
    20. 20
      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, 614620,  DOI: 10.1016/j.nucmedbio.2015.03.007
    21. 21
      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, 595600,  DOI: 10.1524/ract.2013.2066
    22. 22
      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, 3350,  DOI: 10.3390/jnt2010003
    23. 23
      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, 97449755,  DOI: 10.1039/D0NR01172J
    24. 24
      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, 79857997,  DOI: 10.1039/C8TB02173B
    25. 25
      Kanematsu, N.; Inaniwa, T.; Koba, Y. Relationship between Electron Density and Effective Densities of Body Tissues for Stopping, Scattering, and Nuclear Interactions of Proton and Ion Beams. Med. Phys. 2012, 39, 10161020,  DOI: 10.1118/1.3679339
    26. 26
      Greaves, G.; Hinks, J. A.; Busby, P.; Mellors, N. J.; Ilinov, A.; Kuronen, A.; Nordlund, K.; Donnelly, S. E. Enhanced Sputtering Yields from Single-Ion Impacts on Gold Nanorods. Phys. Rev. Lett. 2013, 111, 065504  DOI: 10.1103/PhysRevLett.111.065504
    27. 27
      Holzwarth, U.; Ojea Jimenez, I.; 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,  DOI: 10.1186/s41181-018-0042-3
    28. 28
      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, 473481,  DOI: 10.1524/ract.2012.1935
    29. 29
      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, 183189,  DOI: 10.1016/j.apradiso.2017.07.030
    30. 30
      Gangwar, P.; Pandey, M.; Sivakumar, S.; Pala, R. G. S.; Parthasarathy, G. Increased Loading of Eu3+ Ions in Monazite LaVO4 Nanocrystals via Pressure-Driven Phase Transitions. Cryst. Growth Des. 2013, 13, 23442349,  DOI: 10.1021/cg3018908
    31. 31
      Jia, C. N.; Sun, L. D.; Yan, Z. G.; Pang, Y. C.; Lü, S. Z.; Yan, C. H. Monazite and Zircon Type LaVO4:Eu Nanocrystals – Synthesis, Luminescent Properties, and Spectroscopic Identification of the Eu3+ Sites. Eur. J. Inorg. Chem. 2010, 2010, 26262635,  DOI: 10.1002/ejic.201000038
    32. 32
      Nuñez, N. O.; Zambrano, P.; García-Sevillano, J.; Cantelar, E.; Rivera-Fernández, S.; De La Fuente, J. M.; Ocaña, M. Uniform Poly(Acrylic Acid)-Functionalized Lanthanide-Doped LaVO4 Nanophosphors with High Colloidal Stability and Biocompatibility. Eur. J. Inorg. Chem. 2015, 2015, 45464554,  DOI: 10.1002/ejic.201500265
    33. 33
      Cheng, X.; Guo, D.; Feng, S.; Yang, K.; Wang, Y.; Ren, Y.; Song, Y. Structure and Stability of Monazite- and Zircon-Type LaVO4 under Hydrostatic Pressure. Opt. Mater. (Amst). 2015, 49, 3238,  DOI: 10.1016/j.optmat.2015.08.011
    34. 34
      Oka, Y.; Yao, T.; Yamamoto, N. Hydrothermal Synthesis of Lanthanum Vanadates: Synthesis and Crystal Structures of Zircon-Type LaVO4 and a New Compound LaV3O9. J. Solid State Chem. 2000, 152, 486491,  DOI: 10.1006/jssc.2000.8717
    35. 35
      Ropp, R. C.; Carroll, B. Precipitation of Rare Earth Vanadates from Aqueous Solution. J. Inorg. Nucl. Chem. 1977, 39, 13031307,  DOI: 10.1016/0022-1902(77)80286-8
    36. 36
      Xu, Z.; Li, C.; Hou, Z.; Peng, C.; Lin, J. Morphological Control and Luminescence Properties of Lanthanide Orthovanadate LnVO4 (Ln = La to Lu) Nano-/Microcrystals Viahydrothermal Process. CrystEngComm 2011, 13, 474482,  DOI: 10.1039/C0CE00161A
    37. 37
      Lotfi, S.; El Ouardi, M.; Ait Ahsaine, H.; Madigou, V.; BaQais, A.; Assani, A.; Saadi, M.; Arab, M. Low-Temperature Synthesis, Characterization and Photocatalytic Properties of Lanthanum Vanadate LaVO4. Heliyon 2023, 9, e17255  DOI: 10.1016/j.heliyon.2023.e17255
    38. 38
      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. 2015, 73, 133140,  DOI: 10.1007/s10971-014-3504-4
    39. 39
      Tian, L.; Mho, S. Enhanced Photoluminescence of YVO4:Eu3+ by Codoping the Sr2+, Ba2+ or Pb2+ Ion. J. Lumin. 2007, 122–123, 99103,  DOI: 10.1016/j.jlumin.2006.01.108
    40. 40
      Huignard, A.; Gacoin, T.; Boilot, J. P. Synthesis and Luminescence Properties of Colloidal YVO4:Eu Phosphors. Chem. Mater. 2000, 12, 10901094,  DOI: 10.1021/cm990722t
    41. 41
      Deblonde, G. J. P.; Zavarin, M.; Kersting, A. B. The Coordination Properties and Ionic Radius of Actinium: A 120-Year-Old Enigma. Coord. Chem. Rev. 2021, 446, 214130  DOI: 10.1016/j.ccr.2021.214130
    42. 42
      Su, J.; Mi, X.; Sun, J.; Yang, L.; Hui, C.; Lu, L.; Bai, Z.; Zhang, X. Tunable Luminescence and Energy Transfer Properties in YVO4:Bi3+, Eu3+ Phosphors. J. Mater. Sci. 2017, 52, 782792,  DOI: 10.1007/s10853-016-0375-9
    43. 43
      Rapp, M.; Lozano, Y.; Fernández-Ramos, M.; Isasi, J.; Palafox, M. A. Superparamagnetic and Light-Emitting Bifunctional Nanocomposites of Iron Oxide and Erbium or Thulium Doped Yttrium Orthovanadate. J. Alloys Compd. 2022, 929, 167065  DOI: 10.1016/j.jallcom.2022.167065
    44. 44
      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, 22112220,  DOI: 10.1002/ejic.201001341
    45. 45
      Wang, F.; Yu, L.; Zhu, Y.; Zhu, Z.; Meng, X.; Lv, Y.; Peng, S.; Yang, L. Defect Control and Optical Performance of Yttrium Orthovanadate Nanocrystals via a Facile PH-Sensitive Synthesis. J. Alloys Compd. 2023, 968, 172259  DOI: 10.1016/j.jallcom.2023.172259
    46. 46
      Ling-Hu, P.; Guo, X.; Hu, J.; Deng, C.; Cui, R. Anomalous 5D0→7F4 Transition of Eu3+-Doped BaLaGaO4 Phosphors for WLEDs and Plant Growth Applications. Adv. Opt. Mater. 2024, 12, 2301760  DOI: 10.1002/adom.202301760
    47. 47
      Chi, F.; Wei, X.; Zhou, S.; Chen, Y.; Duan, C.; Yin, M. Enhanced 5D0 → 7F4 Transition and Optical Thermometry of Garnet Type Ca3Ga2Ge3O12:Eu3+ Phosphors. Inorg. Chem. Front. 2018, 5, 12881293,  DOI: 10.1039/C8QI00199E
    48. 48
      Shannon, R. D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr., Sect. A 1976, 32, 751767,  DOI: 10.1107/S0567739476001551
    49. 49
      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, 339346,  DOI: 10.1134/S1066362219030093
    50. 50
      Kozempel, J.; Vlk, M.; Málková, E.; Bajzíková, A.; Bárta, J.; Santos-Oliveira, R.; Malta Rossi, A. Prospective Carriers of 223Ra for Targeted Alpha Particle Therapy. J. Radioanal. Nucl. Chem. 2015, 304, 443447,  DOI: 10.1007/s10967-014-3615-y
    51. 51
      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. Nanomater. 2020, 10, 1632,  DOI: 10.3390/nano10091632
    52. 52
      Suchánková, P.; Kukleva, E.; Štamberg, K.; Nykl, P.; Sakmár, M.; Vlk, M.; Kozempel, J. Determination, Modeling and Evaluation of Kinetics of 223Ra Sorption on Hydroxyapatite and Titanium Dioxide Nanoparticles. Mater. 2020, 13, 1915,  DOI: 10.3390/ma13081915
    53. 53
      Kukleva, E.; Suchánková, P.; Štamberg, K.; Vlk, M.; Šlouf, M.; Kozempel, J. Surface Protolytic Property Characterization of Hydroxyapatite and Titanium Dioxide Nanoparticles. RSC Adv. 2019, 9, 2198921995,  DOI: 10.1039/C9RA03698A
    54. 54
      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, 36593666,  DOI: 10.1039/C9RA08953E
    55. 55
      Muckley, E. S.; Aytug, T.; Mayes, R.; Lupini, A. R.; Carrillo, J. M. Y.; Goswami, M.; Sumpter, B. G.; Ivanov, I. N. Hierarchical TiO2:Cu2O Nanostructures for Gas/Vapor Sensing and CO2 Sequestration. ACS Appl. Mater. Interfaces 2019, 11, 4846648475,  DOI: 10.1021/acsami.9b18824
    56. 56
      White, F. D.; Thiele, N. A.; Simms, M. E.; Cary, S. K. Structure and Bonding of a Radium Coordination Compound in the Solid State. Nat. Chem. 2024, 16, 168172,  DOI: 10.1038/s41557-023-01366-z
    57. 57
      Akamo, D. O.; Kumar, N.; Li, Y.; Pekol, C.; Li, K.; Goswami, M.; Hirschey, J.; LaClair, T. J.; Keffer, D. J.; Rios, O.; Gluesenkamp, K. R. Stabilization of Low-Cost Phase Change Materials for Thermal Energy Storage Applications. iScience 2023, 26, 107175  DOI: 10.1016/j.isci.2023.107175
    58. 58
      Vives, S.; Ramel, D.; Meunier, C. Evolution of the Structure with the Composition and the Defect Arrangement in the Gadolinium and Samarium Doped and Co-Doped Ceria Systems: A Molecular Dynamics Study. Solid State Ionics 2021, 364, 115611  DOI: 10.1016/j.ssi.2021.115611
    59. 59
      Li, P.; Merz, K. M. Metal Ion Modeling Using Classical Mechanics. Chem. Rev. 2017, 117, 15641686,  DOI: 10.1021/acs.chemrev.6b00440
    60. 60
      Boll, R. A.; Malkemus, D.; Mirzadeh, S. Production of Actinium-225 for Alpha Particle Mediated Radioimmunotherapy. Appl. Radiat. Isot. 2005, 62, 667679,  DOI: 10.1016/j.apradiso.2004.12.003
    61. 61
      Huignard, A.; Buissette, V.; Laurent, G.; Gacoin, T.; Boilot, J. P. Synthesis and Characterizations of YVO4: Eu Colloids. Chem. Mater. 2002, 14, 22642269,  DOI: 10.1021/cm011263a
    62. 62
      Young, R. A.. The Rietveld Method; Oxford University Press, 1981.
    63. 63
      Toby, B. H.; Von Dreele, R. B. GSAS-II: The Genesis of a Modern Open-Source All Purpose Crystallography Software Package. J. Appl. Crystallogr. 2013, 46, 544549,  DOI: 10.1107/S0021889813003531
    64. 64
      Thompson, A. P.; Aktulga, H. M.; Berger, R.; Bolintineanu, D. S.; Brown, W. M.; Crozier, P. S.; in’t Veld, P. J.; Kohlmeyer, A.; Moore, S. G.; Nguyen, T. D.; Shan, R.; Stevens, M. J.; Tranchida, J.; Trott, C.; Plimpton, S. J. LAMMPS - a Flexible Simulation Tool for Particle-Based Materials Modeling at the Atomic, Meso, and Continuum Scales. Comput. Phys. Commun. 2022, 271, 108171  DOI: 10.1016/j.cpc.2021.108171
    65. 65
      Martinez, L.; Andrade, R.; Birgin, E. G.; Martínez, J. M. PACKMOL: A Package for Building Initial Configurations for Molecular Dynamics Simulations. J. Comput. Chem. 2009, 30, 21572164,  DOI: 10.1002/jcc.21224
    66. 66
      Qiao, B.; Skanthakumar, S.; Soderholm, L. Comparative CHARMM and AMOEBA Simulations of Lanthanide Hydration Energetics and Experimental Aqueous-Solution Structures. J. Chem. Theory Comput. 2018, 14, 17811790,  DOI: 10.1021/acs.jctc.7b01018
    67. 67
      Priest, C.; Zhou, J.; Jiang, D.; en Solvation of the Vanadate Ion in Seawater Conditions from Molecular Dynamics Simulations. Inorg. Chim. Acta 2017, 458, 3944,  DOI: 10.1016/j.ica.2016.12.027
    68. 68
      Clark, G. N. I.; Cappa, C. D.; Smith, J. D.; Saykally, R. J.; Head-Gordon, T. The Structure of Ambient Water. Mol. Phys. 2010, 108, 14151433,  DOI: 10.1080/00268971003762134
    69. 69
      Goswami, M.; Kumar, N.; Li, Y.; Hirschey, J.; LaClair, T. J.; Akamo, D. O.; Sultan, S.; Rios, O.; Gluesenkamp, K. R.; Graham, S. Understanding Supercooling Mechanism in Sodium Sulfate Decahydrate Phase-Change Material. J. Appl. Phys. 2021, 129, 245109  DOI: 10.1063/5.0049512
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsnano.3c13213.

    • Description of MD simulations in detail; force-field parameters (Lennard-Jones) used in the MD simulations; RDF of water in LnVO4 and the comparison of the RDF of La, Eu, and oxygen of VO4; RDF of Ac-/Ra-doped lanthanide at different temperatures and the snapshots; Rietveld analysis parameters and XRD patterns; summary of the molar ratio of the nanoconstructs; excitation and emission spectra of 225Ac-doped nanoconstructs and 30 days' decayed 225Ac-doped nanoconstructs, respectively; and different compositions of lanthanides (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.