Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Oct 5.
Published in final edited form as: J Am Chem Soc. 2021 Jun 30;143(27):10429–10440. doi: 10.1021/jacs.1c05339

Py-macrodipa: A Janus Chelator Capable of Binding Medicinally Relevant Rare-Earth Radiometals of Disparate Sizes

Aohan Hu , Eduardo Aluicio-Sarduy , Victoria Brown , Samantha N MacMillan , Kaelyn V Becker ‡,§, Todd E Barnhart , Valery Radchenko ∥,, Caterina F Ramogida ⊥,, Jonathan W Engle ‡,§, Justin J Wilson
PMCID: PMC8491276  NIHMSID: NIHMS1742395  PMID: 34190542

Abstract

Nuclear medicine leverages different types of radiometals for disease diagnosis and treatment, but these applications usually require them to be stably chelated. Given the often-disparate chemical properties of these radionuclides, it is challenging to find a single chelator that binds all of them effectively. Towards addressing this problem, we recently reported a macrocyclic chelator macrodipa with an unprecedented “dual-size-selectivity” pattern for lanthanide (Ln3+) ions, characterized by its high affinity for both the large and the small Ln3+ (J. Am. Chem. Soc. 2020, 142, 13500). Here, we describe a second-generation “macrodipa-type” ligand, py-macrodipa. Its coordination chemistry with Ln3+ was thoroughly investigated experimentally and computationally. These studies reveal that the Ln3+−py-macrodipa complexes exhibit enhanced thermodynamic and kinetic stabilities compared to Ln3+−macrodipa, while retaining the unusual dual-size selectivity. Nuclear medicine applications of py-macrodipa for chelating radiometals with disparate chemical properties were assessed using the therapeutic 135La3+ and diagnostic 44Sc3+, radiometals representing the two size extremes within the Ln3+ series. Radiolabeling and stability studies demonstrate that the rapidly formed complexes of these radionuclides with py-macrodipa are highly stable in human serum. Thus, in contrast to gold standard chelators like DOTA and macropa, py-macrodipa can be harnessed for the simultaneous, efficient binding of radiometals with disparate ionic radii like La3+ and Sc3+, signifying a substantial achievement in nuclear medicine. This concept could enable the facile incorporation of a breadth of medicinally relevant radiometals into chemically identical radiopharmaceutical agents. The fundamental coordination chemistry learned from py-macrodipa provides valuable insight for future chelator developments.

Graphical Abstract

graphic file with name nihms-1742395-f0001.jpg

INTRODUCTION

Nuclear medicine is a rapidly growing and critically important branch of radiology that uses radionuclides to diagnose and treat diseases.14 Radionuclides that are suitable for different applications within nuclear medicine are chosen based on their radioactive emission properties and physical half-lives. Based on these criteria, nearly the entire periodic table is represented by elements with potential relevance in nuclear medicine.59 As such, the chemical properties of these radionuclides can vary widely, requiring different strategies for making them into useful radiopharmaceutical agents. For metallic radionuclides, their incorporation into useful therapeutic and diagnostic moieties requires the use of a chelator. These chelators need to be designed and optimized to rapidly and efficiently bind to the desired radiometal and form stable coordination compounds that remain intact in vivo.10 The chelators that have been used in clinically approved metal-based radiopharmaceutical agents are all derivatives of two structures, 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (DOTA, Chart 1)11 and diethylenetriaminepentaacetic acid (DTPA, Chart 1). Although both chelators are highly effective for many smaller ions, like Lu3+ and In3+, they are insufficient for use with larger ions. As such, new chelators are needed for medicinally valuable radionuclides with different chemical properties.

Chart 1.

Chart 1.

Structures of Ligands Discussed in This Work.

A key challenge in this area has been to design ligands that are effective for very large metal ions. Due to their more diffuse charge density and steric requirements, chelating large metal ions are typically inefficient with the conventional chelators DOTA and DTPA, thus driving a need for more suitable analogues. In this context, the ligand macropa or bp18c6 (Scheme 1), which possesses an unusual selectivity for larger lanthanide (Ln3+) ions over smaller ones (Table 1),12 has arisen as a promising candidate. This chelator has shown particular promise for its ability to chelate and stably retain large radiometal ions including 132/135La3+ and 225Ac3+.1315 A potential shortcoming of macropa for its broader use in nuclear medicine, however, is its poor affinity and stability in binding smaller ions, a feature that renders it incompatible with medicinally relevant radiometals of this type. Consequently, chelators that are capable of stably holding radionuclides of disparate sizes would be of significant value for use in nuclear medicine.

Scheme 1.

Scheme 1.

Depiction of the Conformational Toggle Present in “Macrodipa-Type” Ligands when Chelating Ln3+ Ions with Different Sizes. Reproduced from J. Am. Chem. Soc. 2020, 142, 13500. Copyright 2020 American Chemical Society.

Table 1.

Stability Constants of the Ln3+ Complexes Formed with py-macrodipa, macrodipa, and macropa.

Ln3+ py-macrodipaa macrodipab macropac
log KLnL log KLnL log KLnL
La3+ 14.31(6) 12.19 14.99
Ce3+ 14.65(8) 12.50 15.11
Pr3+ 14.81(6) 12.41 14.70
Nd3+ 14.51(7) 12.25 14.36
Sm3+ 13.66(4) 11.52 13.80
Eu3+ 13.29(6) 10.93 13.01
Gd3+ 12.63(5) 10.23 13.02
Tb3+ 11.95(3) 9.68 11.79
Dy3+ 11.47(4) 9.36 11.72
Ho3+ 10.69(1) 9.36 10.59
Er3+ 10.60(3) 9.71 10.10
Tm3+ 10.92(2) 10.13 9.59
Yb3+ 11.31(4) 10.48 8.89
Lu3+ 11.54(2) 10.64 8.25
Sc3+ 15.83(2) 14.37(7)
a

0.1 M KCl, this work. The values in the parentheses are one standard deviation of the last significant figure.

b

0.1 M KCl. For La3+−Lu3+, ref 16; for Sc3+, this work.

c

0.1 M KCl, ref 12.

We recently reported a macrocyclic chelator macrodipa (Chart 1) that shows an unprecedented selectivity pattern for the Ln3+ ions by exhibiting good affinities for both the largest and smallest members of this series.16 This dual-size-selectivity pattern arises from a significant conformational toggle of this chelator that enables it to accommodate Ln3+ ions of different sizes, as depicted in Scheme 1. For large Ln3+, a 10-coordinate, nearly C2-symmetric complex is formed (Conformation A), whereas for small Ln3+ an 8-coordinate asymmetric complex arises (Conformation B). We believe that this concept of conformational switching upon coordinating metal ions with different sizes potentially prompts a new class of chelators for nuclear medicine that can accommodate a wide range of radionuclides with disparate chemical properties.

Despite this promising and potentially valuable selectivity pattern of macrodipa, its coordination compounds with Ln3+ are kinetically labile (vide infra), a property limiting its use in nuclear medicine. To overcome this shortcoming, we considered different design strategies to enhance the overall thermodynamic and kinetic stability of macrodipa, while maintaining its unusual selectivity pattern. To achieve this goal, we targeted the ligand py-macrodipa (Scheme 1), an “upgraded macrodipa” with one of the ethereal oxygen donors replaced by a pyridyl nitrogen donor. We anticipated that the pyridyl nitrogen donor would enhance both the thermodynamic and kinetic stability of macrodipa. The pyridyl nitrogen donor is a stronger donor for coordinating Ln3+ compared to the ethereal oxygen. For instance, in comparing the ligands Oxyaapa and pypa, which differ only by the substitution of an ethereal oxygen donor for a pyridyl donor (Chart 1), the latter forms thermodynamically more stable complexes with Ln3+.17,18 Furthermore, the planar pyridyl unit is likely to rigidify the metal complex and reinforce its kinetic stability. The incorporation of rigidifying groups into ligand backbones is an established means of enhancing metal complex kinetic inertness,19 as exemplified by comparing the highly efficacious chelator CHX-A”-DTPA, which contains a rigid trans-cyclohexyl ring, to its more flexible analogue DTPA.10

Herein, we report the synthesis and characterization of the “upgraded macrodipa” ligand, py-macrodipa. Its coordination chemistry with all 14 non-radioactive Ln3+ ions and Sc3+, a class of metals ions that have similar chemical properties but different ionic radii,2022 was investigated, enabling us to understand how py-macrodipa interacts with ions of different sizes.23 The candidacy of py-macrodipa for nuclear medicine applications was also evaluated in this work, demonstrated via radiolabeling studies with 135La and 44Sc radionuclides, radiometals with significantly different ionic radii that attain the large and small size extremes within the Ln3+ series. Through these experiments, we show that py-macrodipa is capable of chelating large ions like La3+ and small ions like Sc3+ with both satisfactory thermodynamic and kinetics stabilities, marking a significant improvement over macrodipa, our first-generation “macrodipa-type” ligand. This discovery indicates that py-macrodipa has great potential for use in chelating radiometals with significant size differences, enabling it to be employed with size-mismatched theragnostic pairs.

RESULTS AND DISCUSSION

Ligand Syntheses.

The synthesis of py-macrodipa (Scheme 2) involves the assembly of the previously reported macrocyclic core through a reductive animation24 and the subsequent installation of the two picolinate pendent arms, with an overall yield of 23% over three steps. The identity and purity of the intermediates and final product were ascertained by NMR spectroscopy, mass spectrometry, and analytical HPLC (Figures S1S9). The macrodipa ligand was obtained via published procedures.16,25,26

Scheme 2.

Scheme 2.

Synthetic Route to py-macrodipa.

Ln3+ Complex Stability Constants.

Metal complex stability constants provide a quantitative measure of the thermodynamic affinity of ligands for metal ions. The magnitudes of these stability constants for various ligands are valuable for assessing their use in different metal chelation applications.27,28 In order to study the influence of the newly-introduced pyridyl group in the macrocycle of py-macrodipa, we determined the protonation constants (Ki, Table S1) of py-macrodipa, as well as the stability constants (KLnL, Table 1) of its Ln3+ complexes, via either potentiometric titrations or UV−Vis spectrophotometric titrations.2931 These values for macrodipa, which we previously reported16 except for the newly measured stability constant for Sc3+, are also compared in Tables 1 and S1. These quantities are defined in eqs 12, where the concentration terms represent those at chemical equilibrium, and L signifies the uncomplexed ligand in its fully deprotonated state.

Ki= [HiL] / [H+][Hi-1L] (1)
KLnL= [LnL] / [Ln3+][L] (2)

Figure 1 plots log KLnL values against the Ln3+ 6-coordinate ionic radius22 for py-macrodipa and macrodipa. Similar to macrodipa, py-macrodipa shows a dual-size selectivity across the Ln3+ series, as reflected by its greater affinity for the large and small ions over the intermediately sized ions. This result demonstrates that the incorporation of the pyridine within the macrocycle does not alter the unique size-selectivity profile and further suggests that py-macrodipa is similarly undergoing a conformational toggle upon traversing the Ln3+ series. Remarkably, py-macrodipa shows an overall enhancement in its thermodynamic affinity for the entirety of the Ln3+ series compared to macrodipa. For the large Ln3+ ions, py-macrodipa exhibits an approximate 2-log-unit increase in KLnL over macrodipa, whereas a 1-log-unit enhancement in stability is observed for the small Ln3+.

Figure 1.

Figure 1.

Stability constants of Ln3+ complexes formed with py-macrodipa and macrodipa plotted versus ionic radii.

The presence of the pyridine donor in py-macrodipa thus improves the overall Ln3+-binding affinity, while maintaining the novel dual-size-selectivity pattern. Although the KLnL values of py-macrodipa are several orders of magnitude smaller than those of the conventional chelating agent DOTA,32 for the large Ln3+ ions they compare favorably to those of macropa (Figure 1), which has been successfully applied in nuclear medicine applications for radiometals of this type.13,14 A key advantage of py-macrodipa over macropa, however, is that it retains high affinity for the small Ln3+ ions for which macropa is an ineffective chelator.

Complex Structures—NMR Spectroscopy.

To investigate the inferred switch between Conformations A and B in py-macrodipa across the Ln3+ series, we used NMR spectroscopy to characterize the solution structures of its complexes with the diamagnetic La3+, Lu3+, Sc3+, and Y3+ ions, with ionic radii spanning from 103.2 pm to 74.5 pm. The 90-pm ionic radius22 of Y3+ aligns well with the minimum affinity among all Ln3+ complexes (Figure 1). Thus, an investigation of the coordination chemistry of Y3+−py-macrodipa provides insight on the conformational change that is likely occurring as this minimum of stability is traversed.

The 1H and 13C{1H} NMR spectra of these four complexes were acquired in D2O at pD = 7 (Figures 2 and S28S35). A gradual change of the complex conformation is observed as the Ln3+ gets smaller. For La3+−py-macrodipa, the complex is present as the symmetric Conformation A, whereas the complexes of smaller Lu3+ and Sc3+ exist solely as the asymmetric Conformation B. The complex of the intermediately sized Y3+ ion shows a mixture of both Conformations A and B in a molar ratio of 3.6:1. These conformation are identified and assigned based on our previous detailed multinuclear NMR studies on the La3+, Y3+, and Lu3+ complexes of macrodipa.16 Considering the relatively large difference in ionic radius between La3+ and Y3+, we obtained the 1H NMR spectrum of the py-macrodipa complex formed with paramagnetic Eu3+ (ionic radius = 94.7 pm).22 which clearly represents a symmetric Conformation A. The clearly resolved paramagnetic 1H NMR spectrum of this complex (Figure S36) shows that it adopts the symmetric Conformation A. In this study, we also acquired the 1H and 13C{1H} spectra of the Sc3+−macrodipa complex (Figures S37S38), which reveal it to attain Conformation B, as expected based on the small ionic radius of Sc3+. Collectively, these NMR spectroscopic results show that the inclusion of a pyridyl moiety within py-macrodipa does not alter the conformational toggle responsible for accommodating both large and small Ln3+ ions that is observed for macrodipa. Consistent with our expectations and prior studies with macrodipa, large Ln3+ ions form symmetric complexes that attain Conformation A, whereas the asymmetric Conformation B is preferred for small Ln3+ ions.

Figure 2.

Figure 2.

1H NMR spectra of La3+−, Y3+−, Lu3+−, and Sc3+−py-macrodipa complexes (600 MHz, D2O, pD = 7, 25 °C).

Complex Structures—X-ray Crystallography.

To gain further insight on the conformational toggle of py-macrodipa across the Ln3+ series, we characterized the La3+, Lu3+, and Sc3+ complexes by X-ray crystallography. Crystal structures of these three complexes are shown in Figure 3. These crystal structures are consistent with the NMR spectroscopic data, which show the presence of different conformations for Ln3+ ions of different sizes. As previously observed for macrodipa, the La3+ center in the py-macrodipa complex attains a 10-coordinate geometry, forming a compound with the Conformation A geometry and a slightly distorted C2 symmetry. For this complex, the central La3+ ion is fully encapsulated in the macrocycle, interacting with all of these donors. The two pendent picolinate arms coordinate La3+ from above and below the plane of the macrocycle. The complexes of the smaller ions, Lu3+ and Sc3+, attain the 8-coordinate Conformation B, which is of low symmetry. In Conformation B, in addition to the two picolinate donor arms, only three of the macrocyclic donors, N1–N3, appreciably interact with the metal centers. Furthermore, the Lu3+ and Sc3+ complexes are highly comparable except for presence of an inner-sphere water molecule in the Sc3+ complex that is absent in the Lu3+ complex. With this inner-sphere water molecule, the Sc3+–py-macrodipa complex is consistent with what was previously observed for the Lu3+–macrodipa complex,16 with both structures showing hydrogen-bonding interactions between the inner-sphere water and the macrocycle ethers. By contrast, the coordination sphere of the Lu3+ center in the py-macrodipa complex is completed by one of the ethereal oxygens on the macrocycle in place of the inner-sphere water. Although we consider both Sc3+ and Lu3+ complexes to attain Conformation B, we distinguish them further by referring to the anhydrous complex as “Conformation B0” and the complex with the inner-sphere water molecule to be “Conformation B1”.

Figure 3.

Figure 3.

Crystal structures of (a) [La(py-macrodipa)]+, (b) [Lu(py-macrodipa)]+, and (c) [Sc(py-macrodipa)(OH2)]+ complexes. Thermal ellipsoids are drawn at the 50% probability level. Solvent, counterions, and nonacidic hydrogen atoms are omitted for clarity. Only one of the two [La(py-macrodipa)]+ or [Sc(py-macrodipa)(OH2)]+ complexes in the asymmetric unit is shown.

The presence of the inner-sphere water molecule may be a consequence of the conditions under which these crystals were obtained. Crystals of Lu3+–py-macrodipa were grown from the vapor diffusion of Et2O into a MeOH solution containing the complex, reflecting a situation with a limited H2O content. By contrast, crystals of Sc3+–py-macrodipa were obtained directly from an aqueous solution, a condition that would most likely favor H2O coordination. Moreover, the previously reported Lu3+−macrodipa structure, which also attains Conformation B1, was afforded via the analysis of crystals grown from an aqueous solution as well.16 A recent survey on interatomic distances (r) found within the crystal structures in the Inorganic Crystal Structure Database (ICSD) of Ln3+ complexes with direct oxygen atom coordination revealed an average value of r(Lu−O) = 2.342 ± 0.103 Å for 8-coordinate Lu3+ complexes.33 The distance between the Lu3+ center and ethereal oxygen donor O1 within [Lu(py-macrodipa)]+, r(Lu−O) = 2.509 Å, is substantially longer than the majority of Lu3+−O interactions in the ICSD, suggesting this interaction to be weak. Furthermore, for [Lu(macrodipa)(OH2)]+, the distance between Lu3+ and the bound OH2 is r(Lu−O) = 2.281 Å, indicating a typical coordination interaction between these two species. Based on these findings, we anticipate that in aqueous solution, which is of the most significant relevance to the potential application of this chelator for nuclear medicine, the hydrated Conformation B1 is the dominant species.

DFT Calculations.

To further understand the conformational toggle of py-macrodipa across the Ln3+ series, as well as the origins of the enhanced thermodynamic stability of Ln3+−py-macrodipa complexes compared to those of Ln3+−macrodipa, we carried out DFT calculations with Gaussian 09.34 For comparative purposes, the same theoretical methods that we had previously implemented for Ln3+−macrodipa (Ln = La−Lu)16 were applied to study Ln3+−py-macrodipa. Specifically, the dispersion-corrected hybrid functional ωB97XD was employed.35,36 A large-core relativistic effective core potential (LCRECP) and the associated basis set was assigned to the Ln3+ ions (Ln = La−Lu),37 whereas the 6–31G(d,p) basis set was applied to all other lighter atoms.38,39 Solvation effects in water were accounted for using the SMD solvation model.40,41 Furthermore, we also probed Sc3+−macrodipa in this work to compare it directly to its py-macrodipa analogue. A detailed set of equations describing our computational thought flow is given in the Supporting Information.

To investigate the thermodynamic preferences for Conformations A and B for all Ln3+−py-macrodipa complexes, these structures were optimized, and their standard free energies (G°) were calculated. Due to our observations of both the non-hydrated and hydrated Conformations B0 and B1 by X-ray crystallography, both structures were calculated. Balanced chemical reactions for the conformational switch between Conformations A and B (B0 or B1) in aqueous solution can be written as either eq 3 or 4, where the standard free energy change of these two reactions are annotated as ΔG°(A,B0) and ΔG°(A,B1), respectively.

[Ln(py-macrodipa)]+(Conformation A,aq)[Ln(py-macrodipa)]+(Conformation B0,aq) (3)
[Ln(py-macrodipa)]+(Conformation A,aq) + H2(l)[Ln(py-macrodipa)]+(Conformation B1,aq) (4)

Figure 4a summarizes computed ΔG°(A,B0) and ΔG°(A,B1) values across all Ln3+. As expected from the NMR and X-ray crystallographic data, these ΔG° values are positive for early Ln3+ ions, but negative for late Ln3+ ions, signifying a change in thermodynamic preference from Conformation A to B as the series is traversed. Additionally, the ΔG°(A,B1) values are systematically more negative than those for ΔG°(A,B0) by ~30 kJ·mol−1, suggesting that Conformation B1 is considerably more stable than B0 in aqueous solution. Thus, regarding Conformation B, our subsequent calculations focused primarily on Conformation B1.

Figure 4.

Figure 4.

(a) DFT-computed standard free energy differences between Conformations A and B for Ln3+−py-macrodipa complexes, ΔG°(A,B0) and ΔG°(A,B1), as defined in eqs 34. DFT-computed (b) relative binding energies (ΔΔGB°, eq 6) and (c) relative strain energies (ΔΔGS°, eq 7), comparing the Ln3+−py-macrodipa and Ln3+−macrodipa complexes.

Another feature of the Ln3+−py-macrodipa complexes that warrants computational investigation is their enhanced thermodynamic stabilities over the Ln3+−macrodipa complexes. To understand this property, we analyzed the free energy of the formation of these metal-ligand coordination complexes by breaking this process up into two distinct steps. In the first step, the free ligand adjusts its structure to a conformation that is appropriate for metal binding (eq 5), the free energy change of which is defined as the strain energy (ΔGS°). Once the ligand has adopted the proper conformation, Ln3+ binding occurs (eq 6), resulting in a free energy change defined as the binding energy (ΔGB°).

L2(free,aq)L2(Conformation A or B1,aq) (5)
L2(Conformation A or B1,aq) + Ln3+[LnL]+(Conformation A or B1,aq) (6)

These free energy terms, ΔGS° and ΔGB°, were investigated for both ligand systems in both possible conformations (A and B1). These values are plotted in a relative manner (ΔΔGS° and ΔΔGB°, defined in eqs 78) for a straightforward comparison between the two ligand systems, as shown in Figures 4b and 4c.

ΔΔGS°(A or B1) = ΔGS°(py-macrodipa, A or B1)  ΔGS°(macrodipa, A or B1) (7)
ΔΔGB°(A or B1) = ΔGB°(py-macrodipa, A or B1)  ΔGB°(macrodipa, A or B1) (8)

The difference in binding energies, ΔΔGB°, demonstrate a clear contribution to the enhanced stability of Ln3+−py-macrodipa complexes. Regardless of the complex conformation, both ΔΔGB°(A) and ΔΔGB°(B1) are negative across the whole Ln3+ series, indicating that Ln3+-binding is more favorable for py-macrodipa. This result can be attributed to the presence of the pyridyl group within the macrocycle of this ligand, which is a stronger donor than the analogous ethereal oxygen donor in macrodipa. The ΔΔGS°(A) values are less than zero for all Ln3+, indicating that the rigid planar pyridine group in the macrocycle of py-macrodipa favors the complex formation by preorganizing this ligand for metal binding in Conformation A. By contrast, ΔΔGS°(B1) fluctuates about zero across the Ln3+ series, suggesting that the rigidity introduced by the pyridyl moiety in py-macrodipa does not significantly influence the complex formation in Conformation B1.

Photophysical Studies

Within the Ln3+ series, Eu3+ and Tb3+ are known for their rich photoluminescent properties.4245 For a better understanding and more complete characterization on Ln3+−py-macrodipa complexes, the photophysical properties its Eu3+ and Tb3+ complexes were investigated. These results are summarized in Table 2.

Table 2.

Photophycial Properties of Eu3+− and Tb3+−py-macrodipa Complexes.a

λmax/nm Quantum Yield/% τ(H2O)/ms τ(D2O)/ms q
Eu3+−py-macrodipa 273 0.21(1) 1.12(1) 1.43(1) −0.1
Tb3+−py-macrodipa 273 0.84(4) 3.14(3) 3.14(2) −0.3
a

pH 7.4, 22 °C. The values in the parentheses are one standard deviation of the last significant figure.

The UV−Vis absorption and emission spectra of these four complexes were acquired at pH 7.4 (Figure S42). Both complexes give rise to an absorption band centered at 273 nm, which is attributed to the ππ* transition of the pyridyl units.46 Upon this pyridyl-based excitation, both complexes show characteristic Ln3+-based f-f emissions, through the well-established antenna effect.47,48 Additionally, we determined their photoluminescence quantum yields by a relative method,49 using [Ru(bpy)3]Cl2 (quantum yield = 4.0% in air-saturated water)50 as the reference compound.

A comparison between luminescence lifetimes (τ) of Eu3+ and Tb3+ complexes acquired in H2O and D2O can be used to determine the number of inner-sphere water molecules (q).51 The q values of Eu3+ and Tb3+ complexes are calculated by empirically derived expressions, eqs 910,52,53 where in both equations Δkobs = 1/τ(H2O) − 1/τ(D2O) in ms−1.

qEu= 1.11 × (Δkobs 0.31) (9)
qTb= 5.0 × (Δkobs 0.06) (10)

As shown in Table 2, the measured τ values in H2O and D2O indicate both complexes do not possess any inner-sphere water molecule (q = 0). This observation is consistent to our other experimental (Figures 1 and S36) and computational (Figure 4a) results that suggest these complexes adopt Conformation A, where an inner-sphere water is absent.

DTPA Transchelation Challenge Assays.

For nuclear medicine applications, the kinetic stabilities of radiometal complexes are of critical importance because release of free metal ions is expected to give rise to toxic side effects.1 To assess the suitability of py-macrodipa for use in nuclear medicine, we evaluated the kinetic stabilities of the Ln3+−py-macrodipa complexes and compared them to the corresponding Ln3+−macrodipa complexes. Specifically, we monitored the transchelation reaction of these Ln3+ complexes by UV−Vis spectroscopy in the presence of 100-fold excess DTPA at pH 7.4 and RT (22 °C), conditions that thermodynamically favor loss of the Ln3+ ion from both py-macrodipa and macrodipa.5456 With the large excess of DTPA used, this reaction follows pseudo-first-order kinetics, enabling us to use readily determined half-lives (t1/2) as a comparative measure of the complex kinetic stability. These t1/2 values are listed in Table 2.

For both ligand systems, the kinetic stability of their Ln3+ complexes follows the trends of their thermodynamic stability, with the early and late Ln3+ complexes displaying greater kinetic stability than those in the middle of the 4f series. Notably, the dissociation of Er3+−py-macrodipa complex follows a biexponential function, affording two pseudo-first order rate constants. We tentatively assign the two kinetic processes as arising from dissociations of the distinct Conformations A and B, which presumably possess different kinetic labilities. The 1H NMR spectrum of Y3+−py-macrodipa (Figure 2) provides support for this hypothesis, as it clearly shows a mixture of both conformers. Based on the comparable ionic radii of Er3+ (91 pm) and Y3+ (90 pm),22 Er3+−py-macrodipa is expected to similarly exist in solution with both conformers, providing an explanation for the two observed kinetic processes.

Overall, the kinetic stabilities of Ln3+−macrodipa complexes are poor, a property that is reflected by the fact that none of their half-lives exceeds 30 min. By contrast, py-macrodipa forms Ln3+ complexes with significantly enhanced kinetic stability. In particular, complexes of both the largest La3+ ion (t1/2 = 6.3 d) and smallest Sc3+ ion (t1/2 = 16.6 h) show remarkable kinetic stabilities, a prerequisite for using this ligand in nuclear medicine. These complex t1/2 values are substantially longer than the physical half-lives of their corresponding medicinally relevant radionuclides like 132La (t1/2 = 4.6 h), 135La (t1/2 = 18.9 h), and 44Sc (t1/2 = 4.0 h).57,58 The excellent kinetic stability of La3+− and Sc3+−py-macrodipa indicates that py-macrodipa is a promising candidate for chelating radioactive La3+ and Sc3+.

Radiolabeling Studies.

Having shown that py-macrodipa forms complexes with La3+ and Sc3+ of both high thermodynamic and kinetic stabilities, we next sought to evaluate its ability to radiolabel medicinally relevant radioisotopes of these ions. The radionuclides 135La3+ and 44Sc3+, which have potential applications for Auger electron therapy and positron emission tomography imaging,59,60 were produced and isolated following our established procedures.13,61,62 For comparison, radiolabeling studies were also conducted using macrodipa and DOTA. For py-macrodipa and macrodipa, these experiments were executed at 25 °C, whereas for DOTA the elevated temperature of 80 °C was employed due to its known slower binding kinetics. These studies used a pH 5.5 buffer, and the reactions were monitored at 15, 30, and 60 min.

Table 4 lists the apparent molar activities (AMA) for the 135La- and 44Sc-labeled ligands obtained under these conditions. AMA, which describes the amount of radionuclide activity incorporated normalized to the amount of chelator, has been previously used to compare the radiolabeling efficacies of different ligands.13,6163 Among these three ligands, py-macrodipa affords significantly higher AMAs for both 44Sc and 135La in comparison to the other ligands. Notably, the AMA does not appreciably increase beyond 15 min, indicating that radiolabeling at this mild temperature (25 °C) for this ligand is complete on this short time scale. The AMAs obtained for macrodipa show that although it can bind 135La3+, it is ineffective for 44Sc3+. Conversely, even at elevated temperatures, the established chelator DOTA was only able to bind the smaller 44Sc3+ radionuclide but failed in sufficiently complexing the large 135La3+ radionuclide.

Table 4.

Apparent Molar Activity (AMA, Ci·μmol−1) of 135La3+ and 44Sc3+ Radiolabeling with Ligands py-macrodipa, macrodipa, and DOTA.

Radiolabeling Conditions 135La3+ 44Sc3+
py-macrodipa pH 5.5
25 °C
15 min 1.488 0.798
30 min 1.363 0.888
60 min 1.709 0.733
macrodipa pH 5.5
25 °C
15 min 0.105 0.006
30 min 0.088 0.004
60 min 0.095 0.004
DOTA pH 5.5
80 °C
15 min 0.005 0.419
30 min 0.020 0.395
60 min 0.030 0.464

Collectively, these data show that py-macrodipa exhibits versatile radiolabeling properties that enable it to effectively bind to both the large 135La3+ and small 44Sc3+ radionuclides. Its overall radiolabeling efficiency, as reflected by significantly large AMAs, is substantially better than those of our first-generation ligand macrodipa, thus validating our design strategy of incorporating the pyridine donor within the macrocycle. In comparison to the widely used “gold-standard” ligand DOTA, py-macrodipa can effectively be radiolabeled by ions of both types. To our best knowledge, py-macrodipa is the first chelator capable of effectively binding both La3+ and Sc3+ radionuclides, which possess substantially different ionic radii.

Human Serum Challenge Assays.

Having demonstrated that py-macrodipa is effectively radiolabeled by 135La3+ and 44Sc3+, we next evaluated the kinetic stability of the resulting complexes of these radiometals in human serum at 37 °C to assess the potential value of this ligand for nuclear medicine. The stabilities of 135La3+− and 44Sc3+−py-macrodipa complexes were monitored by radio-HPLC over 48 h and 8 h, respectively, time periods that account for the different physical half-lives of these radionuclides. Although the signal-to-noise of the chromatograms gradually decreases due to physical decays of the radionuclides, only a single peak in the chromatograms corresponding to 135La3+− and 44Sc3+−py-macrodipa complexes is observed over the duration of experiment (Figures 5a5b), indicating that both complexes exhibit superb kinetic stabilities in human serum. It is worth noting that the signal to noise of these chromatograms is fading over time. This observation is related to the physical decay of the radionuclides, which contributes to a gradually decreasing detected intensity. For comparison, the 135La3+− and 44Sc3+−macrodipa complexes were also tested for their stabilities in human serum (Figures 5c5d). Although 135La3+−macrodipa is sufficiently stable in human serum over 48 h, 44Sc3+−macrodipa complex shows deficient kinetic stability. These results highlight that py-macrodipa is an improvement over macrodipa because of its ability to form kinetically stable complexes with both the large 135La3+ and small 44Sc3+. Thus, based on its promising radiolabeling and radiometal-stabilizing properties, py-macrodipa is a valuable radiopharmaceutical candidate.

Figure 5.

Figure 5.

Human serum challenge (37 °C) on the (a) 135La3+−py-macrodipa, (b) 44Sc3+−py-macrodipa, (c) 135La3+−macrodipa, and (d) 44Sc3+−macrodipa complexes monitored by radio-HPLC.

CONCLUSION

In this work, the ligand py-macrodipa was successfully synthesized and its coordination chemistry with the Ln3+ ions was investigated in detail via both experimental and theoretical methods. Building upon the unusual dual-size selectivity of our first-generation ligand macrodipa, we have shown that the presence of pyridyl moiety within the macrocycle of py-macrodipa substantially improves both the thermodynamic and kinetic stability of its Ln3+ complexes. The enhanced Ln3+-binding properties of py-macrodipa manifest in efficient radiolabeling of both 135La3+ and 44Sc3+ and excellent stability of the resulting complexes in human serum. This ligand is the first chelator capable of binding radiometal ions with disparate sizes such as 135La3+ and 44Sc3+, both with satisfactory efficacy, marking a significant improvement over current state-of-the-art gold standard chelator DOTA, which can only be used for a single type of ion. We thus describe py-macrodipa as a “Janus chelator” because of its ability to effectively bind two types of metal ions by switching its conformation. The success of py-macrodipa demonstrates that coordination chemists can tailor a ligand to effectively bind metal ions with disparate properties, rather than optimizing a ligand for a single ion type. Even though this work focused exclusively on the Ln3+ series, we anticipate that this concept can be expanded to other metal ions on the periodic table. Another genre of “Janus chelator” was recently reported, where the chelator switches between phenolate and pyridine pendant donors to effectively bind the Mn ion in both its chemically hard and borderline +3 and +2 oxidation states.64 Thus, our description of py-macrodipa, as a Janus chelator that bends its backbone differently to accommodate ions of different sizes, provides a complementary approach. The design of py-macrodipa is an important achievement for inspiring future ligand design strategies and represents an important step in realizing an omnipotent “Janus chelator” that can bind all metal ions of disparate properties. Ongoing work is focused on the development and implementation of a bifunctional derivative of py-macrodipa that can be used for targeted nuclear medicine.

Supplementary Material

Supporting Information

Table 3.

Half-lives of Ln3+−py-macrodipa and Ln3+−macrodipa Complexes when Challenged with 100 Equivalents of DTPA.a

Ln3+–py-macrodipa Ln3+–macrodipa
La3+ 6.3 ± 0.3 d 1678 ± 33 s
Nd3+ 3.9 ± 0.2 d 615 ± 14 s
Gd3+ 5524 ± 107 s 54 ± 2 s
Er3+ 578 ± 23 s; 61± 3 s 5 ± 1 s
Lu3+ 853 ± 10 s 65 ± 1 s
Sc3+ 16.6 ± 1.7 h 782 ± 18 s
a

[LnL] = 100 μM, pH 7.4, 22 °C.

Acknowledgement.

This research was supported by the National Institutes of Biomedical Imaging and Bioengineering of the National Institutes of Health under Award Numbers R21EB027282 and R01EB029259. This research made use of the NMR Facility at Cornell University, which is supported, in part, by the U.S. National Science Foundation under award number CHE-1531632. This work was supported in part by award from the Department of Energy Office of Science (DE-SC0020417). TRIUMF receives funding via a contribution agreement with the National Research Council of Canada. We thank Mr. Angus Koller and Prof. Eszter Boros at Stony Brook University for valuable discussions.

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c05339.

Experimental procedures and supplementary data for syntheses, potentiometric titrations, UV−Vis spectrophotometric titrations, NMR spectroscopy studies, X-ray crystallography studies, DFT calculations, photophysical studies, DTPA transchelation challenges, radiolabeling studies, and human serum challenges (PDF)

Crystallographic data for La3+−, Lu3+−, and Sc3+−py-macrodipa complexes (CIF)

Geometry outputs for all DFT-optimized structures (ZIP)

Accession Codes

CCDC 2085495–2085497 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

The authors declare no competing financial interest.

REFERENCES

  • (1).Mankoff D Why Nuclear Imaging and Radiotherapy? In Radiopharmaceutical Chemistry; Lewis JS, Windhorst AD, Zeglis BM, Eds.; Springer Nature Switzerland AG: Cham, Switzerland, 2019; pp 3–10. [Google Scholar]
  • (2).Sgouros G; Bodei L; McDevitt MR; Nedrow JR Radiopharmaceutical Therapy in Cancer: Clinical Advances and Challenges. Nat. Rev. Drug Discovery 2020, 19, 589–608. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (3).Dondi M; Kashyap R; Paez D; Pascual T; Zaknun J; Mut Bastos F; Pynda Y Trends in Nuclear Medicine in Developing Countries. J. Nucl. Med 2011, 52, 16S–23S. [DOI] [PubMed] [Google Scholar]
  • (4).Delbeke D; Segall GM Status of and Trends in Nuclear Medicine in the United States. J. Nucl. Med 2011, 52, 24S–28S. [DOI] [PubMed] [Google Scholar]
  • (5).Blower PJ A Nuclear Chocolate Box: the Periodic Table of Nuclear Medicine. Dalton Trans. 2015, 44, 4819–4844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).Cutler CS; Hennkens HM; Sisay N; Huclier-Markai S; Jurisson SS Radiometals for Combined Imaging and Therapy. Chem. Rev 2013, 113, 858–883. [DOI] [PubMed] [Google Scholar]
  • (7).Boros E; Packard AB Radioactive Transition Metals for Imaging and Therapy. Chem. Rev 2019, 119, 870–901. [DOI] [PubMed] [Google Scholar]
  • (8).Kostelnik TI; Orvig C Radioactive Main Group and Rare Earth Metals for Imaging and Therapy. Chem. Rev 2019, 119, 902–956. [DOI] [PubMed] [Google Scholar]
  • (9).Mikolajczak R; van der Meulen NP; Lapi SE Radiometals for Imaging and Theranostics, Current Production, and Future Perspectives. J. Labelled Compd. Radiopharm 2019, 62, 615–634. [DOI] [PubMed] [Google Scholar]
  • (10).Price EW; Orvig C Matching Chelators to Radiometals for Radiopharmaceuticals. Chem. Soc. Rev 2014, 43, 260–290. [DOI] [PubMed] [Google Scholar]
  • (11).Stasiuk GJ; Long NJ The Ubiquitous DOTA and its Derivatives: the Impact of 1,4,7,10-Tetraazacyclododecane-1,4,7,10-tetraacetic Acid on Biomedical Imaging. Chem. Commun 2013, 49, 2732–2746. [DOI] [PubMed] [Google Scholar]
  • (12).Roca-Sabio A; Mato-Iglesias M; Esteban-Gómez D; Tóth É; de Blas A; Platas-Iglesias C; Rodríguez-Blas T Macrocyclic Receptor Exhibiting Unprecedented Selectivity for Light Lanthanides. J. Am. Chem. Soc 2009, 131, 3331–3341. [DOI] [PubMed] [Google Scholar]
  • (13).Aluicio-Sarduy E; Thiele NA; Martin KE; Vaughn BA; Devaraj J; Olson AP; Barnhart TE; Wilson JJ; Boros E; Engle JW Establishing Radiolanthanum Chemistry for Targeted Nuclear Medicine Applications. Chem. - Eur. J 2020, 26, 1238–1242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Thiele NA; Brown V; Kelly JM; Amor-Coarasa A; Jermilova U; MacMillan SN; Nikolopoulou A; Ponnala S; Ramogida CF; Robertson AKH; Rodríguez-Rodríguez C; Schaffer P; Williams C Jr.; Babich JW; Radchenko V; Wilson JJ An Eighteen-Membered Macrocyclic Ligand for Actinium-225 Targeted Alpha Therapy. Angew. Chem., Int. Ed 2017, 56, 14712–14717. [DOI] [PubMed] [Google Scholar]
  • (15).Abou DS; Thiele NA; Gutsche NT; Villmer A; Zhang H; Woods JJ; Baidoo KE; Escorcia FE; Wilson JJ; Thorek DLJ Towards the Stable Chelation of Radium for Biomedical Applications with an 18-Membered Macrocyclic Ligand. Chem. Sci 2021, 12, 3733–3742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).Hu A; MacMillan SN; Wilson JJ Macrocyclic Ligands with an Unprecedented Size-Selectivity Pattern for the Lanthanide Ions. J. Am. Chem. Soc 2020, 142, 13500–13506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (17).Hu A; Keresztes I; MacMillan SN; Yang Y; Ding E; Zipfel WR; DiStasio RA Jr.; Babich JW; Wilson JJ Oxyaapa: A Picolinate-Based Ligand with Five Oxygen Donors that Strongly Chelates Lanthanides. Inorg. Chem 2020, 59, 5116–5132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Li L; Jaraquemada-Peláez M de G; Kuo H-T; Merkens H; Choudhary N; Gitschtaler K; Jermilova U; Colpo N; Uribe-Munoz C; Radchenko V; Schaffer P; Lin K-S; Bénard F; Orvig C Functionally Versatile and Highly Stable Chelator for 111In and 177Lu: Proof-of-Principle Prostate-Specific Membrane Antigen Targeting. Bioconjugate Chem. 2019, 30, 1539–1553. [DOI] [PubMed] [Google Scholar]
  • (19).Hancock RD; Martell AE Ligand Design for Selective Complexation of Metal Ions in Aqueous Solution. Chem. Rev 1989, 89, 1875–1914. [Google Scholar]
  • (20).Cotton SA Scandium, Yttrium & the Lanthanides: Inorganic & Coordination Chemistry. Encyclopedia of Inorganic Chemistry; John Wiley & Sons, Ltd.: Chichester, UK, 2006. [Google Scholar]
  • (21).Seitz M; Oliver AG; Raymond KN The Lanthanide Contraction Revisited. J. Am. Chem. Soc 2007, 129, 11153–11160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (22).Shannon RD Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr., Sect. A: Found. Adv 1976, 32, 751–767. [Google Scholar]
  • (23). In the following discussion, we will include Sc3+ into the term Ln3+ (i.e., Ln = La–Lu, Sc) for simplicity, unless otherwise notified.
  • (24).Lüning U; Baumstark R; Peters K; von Schnering HG Concave Reagents, 3. Synthesis, Basicity, and Conformation of New Concave Pyridines. Liebigs Ann. Chem 1990, 129–143. [Google Scholar]
  • (25).Griffin JLW; Coveney PV; Whiting A; Davey R Design and Synthesis of Macrocyclic Ligands for Specific Interaction with Crystalline Ettringite and Demonstration of a Viable Mechanism for the Setting of Cement. J. Chem. Soc., Perkin Trans 21999, 1973–1981. [Google Scholar]
  • (26).Lukyanenko NG; Basok SS; Filonova LK Macroheterocycles. Part 44. Facile Synthesis of Azacrown Ethers and Cryptands in a Two-Phase System. J. Chem. Soc., Perkin Trans 11988, 3141–3147. [Google Scholar]
  • (27).Martell AE; Hancock RD Stability Constants and Their Measurements. Metal Complexes in Aqueous Solutions; Plenum Press: New York, 1996; pp 217–244. [Google Scholar]
  • (28).Peters JA; Djanashvili K; Geraldes CFGC; Platas-Iglesias C The Chemical Consequences of the Gradual Decrease of the Ionic Radius along the Ln-Series. Coord. Chem. Rev 2020, 406, 213146. [Google Scholar]
  • (29).Gans P; O’Sullivan B GLEE, a New Computer Program for Glass Electrode Calibration. Talanta 2000, 51, 33–37. [DOI] [PubMed] [Google Scholar]
  • (30).Gans P; Sabatini A; Vacca A Investigation of Equilibria in Solution. Determination of Equilibrium Constants with the HYPERQUAD Suite of Programs. Talanta 1996, 43, 1739–1753. [DOI] [PubMed] [Google Scholar]
  • (31).Gans P; Sabatini A; Vacca A Determination of Equilibrium Constants from Spectrophometric Data Obtained from Solutions of Known pH: the Program pHab. Ann. Chim 1999, 89, 45–49. [Google Scholar]
  • (32).Cacheris WP; Nickle SK; Sherry AD Thermodynamic study of Lanthanide Complexes of 1,4,7-Triazacyclononane-N,N′,N″-triacetic Acid and 1,4,7,10-Tetraazacyclododecane-N,N′,N″,N‴-tetraacetic Acid. Inorg. Chem 1987, 26, 958–960. [Google Scholar]
  • (33).Gagné OC Bond-Length Distributions for Ions Bonded to Oxygen: Results for the Lanthanides and Actinides and Discussion of the f-Block Contraction. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater 2018, 74, 49–62. [Google Scholar]
  • (34).Frisch MJ; Trucks GW; Schlegel HB; Scuseria GE; Robb MA; Cheeseman JR; Scalmani G; Barone V; Mennucci B; Petersson GA; Nakatsuji H; Caricato M; Li X; Hratchian HP; Izmaylov AF; Bloino J; Zheng G; Sonnenberg JL; Hada M; Ehara M; Toyota K; Fukuda R; Hasegawa J; Ishida M; Nakajima T; Honda Y; Kitao O; Nakai H; Vreven T; Montgomery JA Jr.; Peralta JE; Ogliaro F; Bearpark M; Heyd JJ; Brothers E; Kudin KN; Staroverov VN; Kobayashi R; Normand J; Raghavachari K; Rendell A; Burant JC; Iyengar SS; Tomasi J; Cossi M; Rega N; Millam JM; Klene M; Knox JE; Cross JB; Bakken V; Adamo C; Jaramillo J; Gomperts R; Stratmann RE; Yazyev O; Austin AJ; Cammi R; Pomelli C; Ochterski JW; Martin RL; Morokuma K; Zakrzewski VG; Voth GA; Salvador P; Dannenberg JJ; Dapprich S; Daniels AD; Farkas Ö; Foresman JB; Ortiz JV; Cioslowski J; Fox DJ Gaussian 09, Revision D. 01; Gaussian, Inc.: Wallingford, CT, 2009. [Google Scholar]
  • (35).Chai J-D; Head-Gordon M Systematic Optimization of Long-Range Corrected Hybrid Density Functionals. J. Chem. Phys 2008, 128, 084106. [DOI] [PubMed] [Google Scholar]
  • (36).Chai J-D; Head-Gordon M Long-Range Corrected Hybrid Density Functionals with Damped Atom–Atom Dispersion Corrections. Phys. Chem. Chem. Phys 2008, 10, 6615–6620. [DOI] [PubMed] [Google Scholar]
  • (37).Dolg M; Stoll H; Savin A; Preuss H Energy-Adjusted Pseudopotentials for the Rare Earth Elements. Theor. Chim. Acta 1989, 75, 173–194. [Google Scholar]
  • (38).Hehre WJ; Ditchfield R; Pople JA Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys 1972, 56, 2257–2261. [Google Scholar]
  • (39).Hariharan PC; Pople JA The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies. Theor. Chim. Acta 1973, 28, 213–222. [Google Scholar]
  • (40).Marenich AV; Cramer CJ; Truhlar DG Universal Solvation Model Based on Solute Electron Density and on a Continuum Model of the Solvent Defined by the Bulk Dielectric Constant and Atomic Surface Tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [DOI] [PubMed] [Google Scholar]
  • (41).Regueiro-Figueroa M; Esteban-Gómez D; de Blas A; Rodríguez-Blas T; Platas-Iglesias C Understanding Stability Trends along the Lanthanide Series. Chem. - Eur. J 2014, 20, 3974–3981. [DOI] [PubMed] [Google Scholar]
  • (42).Parker D; Dickins RS; Puschmann H; Crossland C; Howard JAK Being Excited by Lanthanide Coordination Complexes: Aqua Species, Chirality, Excited-State Chemistry, and Exchange Dynamics. Chem. Rev 2002, 102, 1977–2010. [DOI] [PubMed] [Google Scholar]
  • (43).Bünzli J-CG Lanthanide Luminescence for Biomedical Analyses and Imaging. Chem. Rev 2010, 110, 2729–2755. [DOI] [PubMed] [Google Scholar]
  • (44).Heffern MC; Matosziuk LM; Meade TJ Lanthanide Probes for Bioresponsive Imaging. Chem. Rev 2014, 114, 4496–4539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (45).Kaczmarek M Lanthanide-Sensitized Luminescence and Chemiluminescence in the Systems Containing Most Often Used Medicines; A Review. J. Lumin 2020, 222, 117174. [Google Scholar]
  • (46).Chatterton N; Bretonnière Y; Pécaut J; Mazzanti M An Efficient Design for the Rigid Assembly of Four Bidentate Chromophores in Water-Stable Highly Luminescent Lanthanide Complexes. Angew. Chem., Int. Ed 2005, 44, 7595–7598. [DOI] [PubMed] [Google Scholar]
  • (47).Moore EG; Samuel APS; Raymond KN From Antenna to Assay: Lessons Learned in Lanthanide Luminescence. Acc. Chem. Res 2009, 42, 542–552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (48).Armelao L; Quici S; Barigelletti F; Accorsi G; Bottaro G; Cavazzini M; Tondello E Design of Luminescent Lanthanide Complexes: From Molecules to Highly Efficient Photo-Emitting Materials. Coord. Chem. Rev 2010, 254, 487–505. [Google Scholar]
  • (49).Brouwer AM Standards for Photoluminescence Quantum Yield Measurements in Solution (IUPAC Technical Report). Pure Appl. Chem 2011, 83, 2213–2228. [Google Scholar]
  • (50).Suzuki K; Kobayashi A; Kaneko S; Takehira K; Yoshihara T; Ishida H; Shiina Y; Oishi S; Tobita S Reevaluation of Absolute Luminescence Quantum Yields of Standard Solutions Using a Spectrometer with an Integrating Sphere and a Back-Thinned CCD Detector. Phys. Chem. Chem. Phys 2009, 11, 9850–9860. [DOI] [PubMed] [Google Scholar]
  • (51).Horrocks W. DeW. Jr.; Sudnick DR Lanthanide Ion Probes of Structure in Biology. Laser-Induced Luminescence Decay Constants Provide a Direct Measure of the Number of Metal-Coordinated Water Molecules. J. Am. Chem. Soc 1979, 101, 334–340. [Google Scholar]
  • (52).Supkowski RM; Horrocks W.DeW. Jr. On the Determination of the Number of Water Molecules, q, Coordinated to Europium(III) Ions in Solution from Luminescence Decay Lifetimes. Inorg. Chim. Acta 2002, 340, 44–48. [Google Scholar]
  • (53).Beeby A; Clarkson IM; Dickins RS; Faulkner S; Parker D; Royle L; de Sousa AS; Williams JAG; Woods M Non-Radiative Deactivation of the Excited States of Europium, Terbium and Ytterbium Complexes by Proximate Energy-Matched OH, NH and CH Oscillators: An Improved Luminescence Method for Establishing Solution Hydration States. J. Chem. Soc., Perkin Trans 21999, 493–503. [Google Scholar]
  • (54).Grimes TS; Nash KL Acid Dissociation Constants and Rare Earth Stability Constants for DTPA. J. Solution Chem 2014, 43, 298–313. [Google Scholar]
  • (55).Moeller T; Thompson LC Observations on the Rare Earths—LXXV: The Stabilities of Diethylenetriaminepentaacetic Acid Chelates. J. Inorg. Nucl. Chem 1962, 24, 499–510. [Google Scholar]
  • (56).Pniok M; Kubíček V; Havlíčková J; Kotek J; Sabatie-Gogová A; Plutnar J; Huclier-Markai S; Hermann P Thermodynamic and Kinetic Study of Scandium(III) Complexes of DTPA and DOTA: A Step Toward Scandium Radiopharmaceuticals. Chem. - Eur. J 2014, 20, 7944–7955. [DOI] [PubMed] [Google Scholar]
  • (57).Abel EP; Clause HK; Fonslet J; Nickles RJ; Severin GW Half-Lives of 132La and 135La. Phys. Rev. C 2018, 97, 034312. [Google Scholar]
  • (58).García-Toraño E; Peyrés V; Roteta M; Sánchez-Cabezudo AI; Romero E; Martínez Ortega A Standardisation and Precise Determination of the Half-Life of 44Sc. Appl. Radiat. Isot 2016, 109, 314–318. [DOI] [PubMed] [Google Scholar]
  • (59).Fonslet J; Lee BQ; Tran TA; Siragusa M; Jensen M; Kibédi T; Stuchbery AE; Severin GW 135La as an Auger-Electron Emitter for Targeted Internal Radiotherapy. Phys. Med. Biol 2017, 63, 015026. [DOI] [PubMed] [Google Scholar]
  • (60).Price TW; Greenman J; Stasiuk GJ Current Advances in Ligand Design for Inorganic Positron Emission Tomography Tracers 68Ga, 64Cu, 89Zr and 44Sc. Dalton Trans. 2016, 45, 15702–15724. [DOI] [PubMed] [Google Scholar]
  • (61).Aluicio-Sarduy E; Hernandez R; Olson AP; Barnhart TE; Cai W; Ellison PA; Engle JW Production and in vivo PET/CT Imaging of the Theranostic Pair 132/135La. Sci. Rep 2019, 9, 10658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (62).Vaughn BA; Ahn SH; Aluicio-Sarduy E; Devaraj J; Olson AP; Engle J; Boros E Chelation with a Twist: a Bifunctional Chelator to Enable Room Temperature Radiolabeling and Targeted PET Imaging with Scandium-44. Chem. Sci 2020, 11, 333–342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (63).Aluicio-Sarduy E; Hernandez R; Valdovinos HF; Kutyreff CJ; Ellison PA; Barnhart TE; Nickles RJ; Engle JW Simplified and Automatable Radiochemical Separation Strategy for the Production of Radiopharmaceutical Quality 86Y Using Single Column Extraction Chromatography. Appl. Radiat. Isot 2018, 142, 28–31. [DOI] [PubMed] [Google Scholar]
  • (64).Gale EM; Jones CM; Ramsay I; Farrar CT; Caravan P A Janus Chelator Enables Biochemically Responsive MRI Contrast with Exceptional Dynamic Range. J. Am. Chem. Soc 2016, 138, 15861–15864. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supporting Information

RESOURCES