Full Papers doi.org/10.1002/ejic.202100256 Investigating the Bismuth Complexes with Benzoazacrown Tri- and Tetra-Acetates Bayirta V. Egorova,*[a] Ekaterina V. Matazova,[a] Gleb Yu. Aleshin,[a] Anastasia D. Zubenko,[b] Anna V. Pashanova,[b, c] Ekaterina A. Konopkina,[a] Artem A. Mitrofanov,[a] Anastasia A. Smirnova,[a] Alexander L. Trigub,[d] Valentina A. Karnoukhova,[b] Olga A. Fedorova,[b, c] and Stepan N. Kalmykov[a, d] The complexation properties towards Bi3 + of benzoazacrown ligands with five (H3BA3A) and six (H4BATA) heteroatoms and corresponding number of acetic arms have been investigated. Complexation constants were determined via potentiometric titration for BA3ABi complex (logβ = 26.2) and refined for BATABi complex (logβ = 31.7). According to NMR, X-ray and EXAFS results, chelation of cation by H3BA3A appears to be outside the macrocyclic cavity both in the solid state and in the aqueous solution. This out-cage location despite the high stability constant causes serum and in vivo instability of complex of smaller azacrown H3BA3A in contrast to H4BATA. Whereas the larger macrocycle and four pendant arms of H4BATA provide long-range coordination effectively shielding the cation thereby ensuring high inertness in vitro, low accumulation of the radionuclide in the organs, and fast clearance. Introduction oxygen ligating atoms. Hence, the most effective chelators used with this cation in radiotherapeutics are polyaminopolyacetates, such as H4DOTA and CHX-DTPA.[6,7] The latter represents rigidified H5DTPA and is commonly used for labeling of heatsensitive molecules to improve the kinetic inertness of the formed complex compared with H5DTPA. The higher kinetic inertness is accompanied by longer radiolabeling,[8] which is crucial for short-lived Bismuth radioisotopes. Recently, it was shown that azacrown ethers with a cavity of 6 heteroatoms and picolinate groups – macropa and derivative ligands – are promising for use with Ac3 + and rare earth elements (REE).[9,10] Additionally, we determined that the aza-18-crown-6 tetraacetate (H4BATA) with the same macrocycle as macropa is a candidate for the Bi3 + cation.[11] We suggested that its high efficiency is associated with the H4DOTA structure: H4BATA seems to be an opened version of H4DOTA, resulting in decreased restriction inside the macrocycle compared with H4DOTA. Additionally, the presence of the rigid benzyl moiety in the macrocycle provides a way for bifunctionalization distanced from the cation-coordinating part. H3NOTA analogs have also been considered for Bismuth cations.[12,13] Based on this, in this work, we studied the complexation features of the aryl-containing azacrowns H4BATA and H3BA3A (as opened H3NOTA) (Scheme 1) with Bi3 +, including the structural features Alpha particles have a high linear energy transfer that causes more efficient DNA breakage of double-stranded cancer cells compared with β particles.[1] Additionally, the short range of the α-particle leads to low toxicity to the surrounding healthy tissues.[2] Bismuth isotopes 212,213Bi are particularly attractive because of their suitable radiation characteristics and the availability of methods to produce these isotopes, such as using convenient radionuclide generator systems in particular.[3,4] Radiopharmaceuticals based on peptides and antibodies labeled with 213Bi isotopes have already demonstrated their potency in clinical trials.[5] One of the challenges for targeted alpha therapy (TAT) using Bi radioisotopes is the design of chelators that are appropriate for forming complexes of high thermodynamic stability and kinetic inertness to avoid possible transchelation and transmetallation in biological media. Bismuth, as an intermediate cation according to the HSAB theory, can be coordinated via combination of basic nitrogen and [a] Dr. B. V. Egorova, E. V. Matazova, Dr. G. Y. Aleshin, E. A. Konopkina, Dr. A. A. Mitrofanov, A. A. Smirnova, Dr. S. N. Kalmykov Chemistry Department Lomonosov Moscow state university 119991 Leninskie Gory, 1/3, Moscow, Russian Federation E-mail: bayirta.egorova@gmail.com [b] Dr. A. D. Zubenko, A. V. Pashanova, V. A. Karnoukhova, Dr. O. A. Fedorova A. N. Nesmeyanov Institute of Organoelement Compounds of Russian Academy of Sciences 119991 Vavilova, 28, GSP-1, Moscow, Russian Federation [c] A. V. Pashanova, Dr. O. A. Fedorova D. Mendeleev University of Chemical Technology of Russia 125047, Miusskaya sqr. 9, Moscow, Russian Federation [d] Dr. A. L. Trigub, Dr. S. N. Kalmykov National Research Center “Kurchatov Institute” 123098 Akademika Kurchatova sqr., 1, Moscow, Russian Federation Supporting information for this article is available on the WWW under https://doi.org/10.1002/ejic.202100256 Eur. J. Inorg. Chem. 2021, 3344 – 3354 Scheme 1. Ligands H3BA3A and H4BATA. 3344 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 of the complexes in aqueous solution via X-ray absorption spectroscopy, 1H NMR, and single crystal X-ray diffraction. The stability constants of the bismuth(III) complexes were determined, and in vivo experiments were conducted. Additionally, full labeling and serum stability studies of the complexes are also described. Results and Discussion Studies of the Bismuth(III) complexes H3BA3A and H4BATA In general, the thermodynamic stability of bismuth(III) complexes of macrocyclic ligands like H4DOTA[14] cannot be determined by potentiometric titrations because of the rapid formation of the hydroxide species even at very low pH. However, in the present research this method was applicable for the determination of the stability constants. An excess of chloride-ion along with fast binding of Bismuth(III) by considered ligands caused an effective hampering of cation’s hydrolysis upon pH increasing and allowed to collect data for constants calculation (Table 1). The stability constants of bismuth(III) complexes with H4BATA were previously reported in a brief communication[11] and have been further clarified in this work, and refined values are presented in Table 1. Table 1. Stepwise (KBiHnL) complexation constants (in log units) of H4BATA (refined for wider pH range vs[11]), H3BA3A, and H4DOTA at 25.0 °C in 0.1 M KNO3[a] and the corresponding pBi values Reaction[a] BiHL + H + ⇄BiH2L BiL + H + ⇄BiHL Bi3 + + L⇄BiL BiLOH + H + ⇄BiL pBi H3BA3A log KBiHnL 2.2(2) 26.2(2) 9.4(2) 24.5 H4BATA H4DOTA[b] [14,15] 3.0(2) 4.2(2) 31.7(2) 8.9(4) 24.1 30.3 25.9 [a] L denotes the ligand; charges of ligand and complex species were omitted for simplicity. [b] I = 1.0 M NaBr The equilibrium state was reached quickly during the titration experiments. The H3BA3A bismuth(III) complex began to dissociate at high pH, forming a hydroxide chelate species which was easily distinguishable in the species distribution diagram (Figure 1). Further pH increase caused visually observable formation of the insoluble hydroxide species in the range of pH 9–10 resulting in the titration experiments with Bi3 + and H3BA3A being terminated. The value of the BA3ABi stability constant was lower than that for DOTABi , however, it was still sufficiently high to form a stable complex that can be defined by potentiometric titration. In the neutral pH range, the main form of the complex in both cases was LBi (Figure 1). The HLBi complex forms in acidic media, and the formation of LBiOH was observed at pH values > 8. To unify logβ for the bismuth complexes with both ligands, we repeatedly titrated solutions of Bi3 + with H4BATA over a wider pH range starting from pH 1.6 (compared to the data presented in[11]). This experiment allowed us to refine logβ values for protonated forms of complex that affected value for BATABi as well, which turned out to be higher by 1.4 orders than that of logβ(DOTABi ). It should be noted that the high basicity of H4BATA would not allow chelation at the initial low pH value, leading to the presence of the BiCln species up to pH 2. Whereas, H3BA3A with lower protonation constants readily binds the cation at the start of the titration, even in the monoprotonated form (Figure 1). For correct comparison of the complexation ability of the different ligands, the different basicity of the ligands should be considered. Thus, pM values (pM = log[Metal]free), calculated for a certain pH by taking into account the known values of the stability constants are often applied for better estimation of the comparative complex stability. Hence, pBi values for complexes with H4BATA, H3BA3A, and the well-known H4DOTA were calculated at pH 7.4 using stability constant values for an excess of ligand: [Bi3 + ]tot = 1.0 μM and [L]tot = 9.0 μM (Table 1). Without the impact of protonation, the obtained pBi values of both the H3BA3A and H4BATA were of the same order of magnitude, while the value of DOTABi complex was higher. These results Figure 1. Species distribution diagram of bismuth(III) in the presence of H3BA3A (a) and H4BATA (b) in an aqueous solution at [Bi3 + ]total = [L]total = 1.0 mM, [Cl ]total = 23 mM. Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org 3345 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 demonstrate that both the studied ligands have a good ability to maintain bismuth cation binding despite the different levels of logβ. Although pBi had a lower value than for H4DOTA, dissociation of complexes with H3BA3A and H4BATA was still expected to be extremely low, which is promising for complexation of bismuth radioisotopes. Structural characterization of the complexes X-ray Crystal Structure of [BA3A·Bi] The structure of complex [BA3A·Bi] was established via singlecrystal X-ray diffraction analysis (Figure 2, S1). The complex crystallizes as a hydrate with three water molecules in the symmetrically independent part of the unit cell. Bond lengths and angles in the macrocyclic ligand were within normal ranges, as confirmed by a Mogul geometry check.[16] The alicyclic crown fragment of the macrocycle was significantly distorted from the idealized planar conformation to adopt the geometry required for effective coordination with a large bismuth(III) cation; for this, the alicyclic part bent out from the average plane of the rigid part. The cation was located outside the cavity of the macrocycle, and was coordinated by three oxygen atoms of the carboxyl groups and three nitrogen atoms of the macrocycle, forming a distorted pentagonal pyramid with an O(6) atom on the vertex and N(1), N(2), N(3), O(4), and O(8) atoms at the base with bond lengths in the range 2.1757(16)-2.5745(18) Å (Table 2). The cation was located outside the cavity of the macrocycle, and was coordinated by three oxygen atoms of the carboxyl groups and three nitrogen atoms of the macrocycle, forming a distorted pentagonal pyramid with an O(6) atom on the vertex and N(1), N(2), N(3), O(4), and O(8) atoms at the base (Table S2). Besides, two shortened distances to oxygen atoms of the macrocycle O(1) and O(2) (2.9010(15) and 2.9832(15) Å, respectively) could be observed. To ensure formation of bonds between Bi and the mentioned atoms we performed QTAIM[17] analysis implemented into multiwfn package[18] (Fig S2, Table S1). Two carboxyl groups of the macrocycle participated in hydrogen bonding with three water molecules. The set of hydrogen bonds connected two complexes into the centrosymmetric dimeric-like structure, with six water molecules located between the complex moieties (Figure S1). NMR study of the complex formation Figure 2. General view of the [BA3ABi] complex in crystal. Anisotropic displacement parameters are drawn at the 50 % probability level, hydrogen atoms were omitted for clarity. The complexation of the ligands H3BA3A and H4BATA with the bismuth cation in aqueous solutions was studied using 1H NMR spectroscopy. The 1H NMR spectra of the free ligands, the spectra after the addition of bismuth nitrate and adjusting the pD of the solution to a certain value were recorded. The spectra in the presence and absence of the Bi3 + ion were analyzed to assess the interactions between the cation and the ligands H3BA3A and H4BATA. Table 2. Selected bond lengths [Å] and angles [°] of the metal coordination environment in [BA3ABi] complex. Bond lengths [Å] Bi(1)-O(6) Bi(1)-O(8) Bi(1)-O(4) Bi(1)-N(2) Bond angles [°] 2.1757(16) 2.3630(16) 2.3805(15) 2.4898(18) Bi(1)-N(1) Bi(1)-N(3) Bi(1)···O(1) Bi(1)···O(2) 2.5436(18) 2.5745(18) 2.9010(15) 2.9832(15) O(6)-Bi(1)-O(8) O(6)-Bi(1)-O(4) O(8)-Bi(1)-O(4) O(6)-Bi(1)-N(2) O(8)-Bi(1)-N(2) O(4)-Bi(1)-N(2) O(6)-Bi(1)-N(1) O(8)-Bi(1)-N(1) O(4)-Bi(1)-N(1) N(2)-Bi(1)-N(1) O(6)-Bi(1)-N(3) O(8)-Bi(1)-N(3) O(4)-Bi(1)-N(3) N(2)-Bi(1)-N(3) 77.54(6) 76.34(6) 76.67(5) 71.80(6) 130.51(6) 129.53(6) 83.95(6) 142.71(5) 67.57(5) 71.01(6) 85.17(6) 68.02(5) 142.92(6) 71.48(6) N(1)-Bi(1)-N(3) O(6)-Bi(1)-O(1) O(6)-Bi(1)-O(2) O(8)-Bi(1)-O(1) O(8)-Bi(1)-O(2) O(4)-Bi(1)-O(1) O(4)-Bi(1)-O(2) N(1)-Bi(1)-O(1) N(1)-Bi(1)-O(2) N(2)-Bi(1)-O(1) N(2)-Bi(1)-O(2) N(3)-Bi(1)-O(1) N(3)-Bi(1)-O(2) O(1)-Bi(1)-O(2) 142.49(6) 144.10(5) 144.18(5) 137.87(5) 101.12(5) 102.66(5) 138.86(5) 63.55(5) 112.54(5) 82.91(5) 83.52(5) 110.89(5) 62.13(5) 51.76(4) Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org 3346 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 Figure 3 shows the 1H NMR spectra of the deprotonated form of the ligand BA3A3 and the complex BA3ABi. Changes in the 1H NMR spectrum observed after the addition of the Bi3 + ion to the D2O solution of H3BA3A demonstrate geminal coupling of the methylene protons belonging to the macrocyclic fragment as well as pendant side-arms. As a result, the seven proton signals of the free ligand BA3A3 transform to twelve proton signals (Figure 3). This behavior is consistent with the presence of a single complex BA3ABi with a relatively rigid structure in solution; fluxional interconversion at ambient temperatures was not observed (on the NMR time scale). Additionally, no change in the number of peaks in the 13C NMR spectrum of the ligand H3BA3A was found upon Bi3 + coordination, suggesting no chemically distinct isomers were formed (Figure S3). In the BA3ABi complex spectrum, the resonances of the protons of the macrocycle, benzene, and pendant arms downfield change because of the polarization effect of the metal ion (Table S2). It can be seen that the H1, H2, and H4 protons are characterized by smaller shifts compared with the H5, H6, and H7 protons, which may indicate weaker binding of the Bi3 + cation by the oxygen atoms of the crown ring. Additionally, the methylene protons of the side arms exhibit the largest signal shifts. In this regard, it can be assumed that all heteroatoms of the macrocycle as well as all three acetate groups of the ligand participate in complexation with the bismuth cation. Interestingly, a splitting of the signals belonging to the geminal H8 protons of two opposite arms can be observed, while methylene H10 protons of the central arm appear as a singlet. This can be explained by the fact that the central acetate group is in the plane of symmetry, which is located perpendicular to the macrocycle, and therefore, the geminal protons H10 are magnetically equivalent. In contrast, the two opposite acetate groups are located above the macrocyclic cavity, coordinating the metal cation. Figure S6 shows the 1H NMR spectra of the free ligand H4BATA, as well as in the presence of the Bi3 + cation (pD = 6.3). The resulting spectrum of the complex is much more complicated than the initial ligand. Analysis of the aromatic region of the spectrum using 2D COSY spectroscopy revealed two sets of signals corresponding to the protons of the benzene ring (Figure S7). Using 2D NOESY spectroscopy, it was possible to identify the H4 protons of the macrocycle closest to benzene (Figure S8). Further complete correlation of signals was complicated by the proximity of a large number of different signals to each other. Based on the potentiometric titration data, the aqueous solution contains only LBi- particles at pH 6.3 (Figure 1b). In this regard, it can be assumed that the obtained NMR spectrum corresponds to complexes with the same composition, but with different structures, i. e., two conformers of the H4BATA complex with Bi3 + cation in a molar ratio of 4 : 1 are simultaneously present in the solution (based on the integral intensity of the protons of the benzene ring). On the 2D NOESY spectrum, in addition to negative NOE’s correlations, positive off-diagonal peaks can be seen. These peaks are correlations from conformational exchange, so the BATABi complex has two different conformations in slow exchange with each other on the NMR time scale and gives the proton spectrum with a doubling of each resonance. Compared to the spectrum of the free H4BATA ligand, the signals of the benzene ring with different values are shifted to the high-field region. This effect can be caused by the shielding of benzene protons by closely spaced carbonyl groups. Thus, it can be assumed that the two conformers of the BATABi complex formed in solution have different positions of the chelating groups relative to the macrocyclic plane. In addition, all four acetate groups are involved in the coordination of the Bi3 + cation. Analysis of the Figure 3. 1H NMR spectra of the free ligand BA3A3 (CL = 8 mM, pD = 12.4) and its complex with Bi3 + (CL = 10 mM, pD = 5.3) in D2O. Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org 3347 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 integral intensities of signals in the spectrum led to the conclusion that each proton appears as an individual signal, thus both conformers of the complex have C1 symmetry. This assumption is in good agreement with the 13C NMR spectrum of the BATABi complex (Figure S9), in which an increased set of signals can be seen compared to the spectrum of the free ligand, namely: 20 signals in the aromatic region instead of 5 signals, and 28 signals in the aliphatic region instead of 7 signals, respectively. X-ray absorption spectroscopy and DFT calculations of bismuth complexes The formation and coordination environment of the bismuth cation in the presence of H3BA3A and H4BATA were further confirmed by EXAFS studies. The nearly similar atomic scattering factors of C, N and O do not allow the atoms to be distinguished in the local surrounding of the Bi atom. First, the absence of Bi Bi interactions indirectly proved complete complexation of Bi3 + by both ligands at pH 5.0–5.5; otherwise, at millimolar concentrations, bismuth forms polynuclear hydroxyl species at pH > 2 or precipitates.[19] Secondly, distances of 2.2–2.5 Å (Figure 4, Table 3) agree well with the X-ray single crystal data for other complexes of bismuth with polyaminopolycarboxylates[20] including crystal structure of [BA3A·Bi] (Table 2). The closest distances in BA3ABi 2.2–2.4 Å correspond to acetic oxygens as O(4), O(6), O(8) while longer ~ 3 Å can be associated with macrocyclic atoms of C, N, O. Furthermore, the obtained structural data concerning the first coordination sphere for both complexes were similar to each other and the total coordination numbers can formally be evaluated as 8 for BATABi and 9 for BA3ABi. Noteworthy, the coordination sphere that is further away is more precisely fitted for the BATABi complex (Table 3, Figure 4), which means higher structural organization in the outer coordination shells (on the long range). This is why the CNs at larger distances are much higher for BA3ABi than for BATABi . To model coordination of Bi3 + in complexes we performed DFT calculations of optimized geometry for both BA3ABi and BATABi (Figure 5, Table 4). Firstly, arrangement of coordinating atoms in BA3ABi agrees with crystallographic data: cavity needs to be bent out to chelate the cation by macrocyclic nitrogens and carboxylic oxygens. This causes one side location of all the pendant arms relative to macrocycle. Additionally, order of distances with coordinating groups is the same as in the solid Figure 4. The fitted spectra for X-ray absorption of·Bismuth complexes with H3BA3A (a) and H4BATA (b) in the R- and k –spaces. Table 3. Structural parameters refined from fitting the EXAFS Bi L3-edge spectra for Bi-BA3A and Bi-BATA at pH 5–5.5. Coordination shell Bi Bi Bi Bi Bi Bi Bi C/N/O C/N/O C/N/O C/N/O C/N/O C/N/O C/N/O Bi-BA3A (pH 5–5.5) R, Å CN σ, Å2 2.23 � 0.01 2.40 � 0.02 3.03 � 0.02 3.24 � 0.02 3.72 � 0.02 4.41 � 0.04 4.97 � 0.07 3 2 4 4 6 4 5 0.007 � 0.003 0.007 � 0.003 0.007 � 0.003 0.005 � 0.003 0.005 � 0.007 0.005 � 0.007 0.005 � 0.007 Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org 3348 Bi-BATA (pH 5–5.5) R, Å CN σ, Å2 2.17 � 0.01 2.47 � 0.02 3.15 � 0.02 3.45 � 0.06 4.31 � 0.03 4.51 � 0.09 3 2 3 1 6 2 0.0045 � 0.0006 0.007 � 0.003 0.005 � 0.003 0.005 � 0.003 0.005 � 0.007 0.004 � 0.025 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 Figure 5. Optimized geometries of the complexes: BA3ABi (a), BATABi conformation 1 (b) and BATABi conformation 2 (c). Table 4. Interatomic distances [Å] of the metal coordination environment in BA3ABi and BATABi complexes according to optimized geometry. BiBA3A Bi Bi Bi Bi Bi Bi Bi Bi N(1) N(2) N(3) O(4) O(6) O(8) O(1) O(2) BiBATA 2.697 2.706 2.665 2.222 2.192 2.214 2.957 2.984 Bi Bi Bi Bi Bi Bi Bi Bi Bi Bi N(1) N(2) N(3) N(4) O(4) O(6) O(8) O(10) O(1) O(2) Conformation 1 Conformation 2 2.933 2.552 2.691 2.730 2.458 2.322 2.305 2.341 4.785 4.413 2.743 2.495 2.499 2.733 2.388 2.418 2.484 2.420 4.344 2.909 state: acetic oxygens are the closest, then aminogroups and macrocyclic oxygens at the distance almost 3 Å (Table 2, Table 4). Furthermore such environment of bismuth is consistent with fitting of EXAFS spectrum and light C/N/O atoms are found at corresponding distances (Table 3). In general, the structure of complex in crystal form is more “compact” than the calculated gazeous one, resulting in smaller M-ligand distance. For the BATABi complex two conformations were established (Figure 5, Table 4). With regard to larger size of macrocycle cation encapsulates inside the cavity in both geometries. It is in line with earlier published bismuth complex of triacetate derivative of aza-18-crown-6[21] where cation was also embedded into the cavity and chelated by three acetates from both sides of macrocycle. The obtained structures of BATABi differ in an arrangement of pendant arms relative to azacrown cavity. The first one is characterized by three arms (O(4), O(6), O(8)) from one side plus capped by the fourth (O(10)) from another (Figure 5b). In the second conformation arms are altering to chelate cation from both sides with respect to macrocycle: two (O(4), O(8)) from above and two (O(6), O(10)) from below (Figure 5c). Apparently such surrounding on all sides causes elongation of distances with chelating atoms compared to BA3ABi. As soon as coordination sphere is saturated by four aminogroups and four acetates macrocyclic oxygens O(1) and O(2) are distanced from Bi3 + as far as it is possible in the cavity - 4–5 Å. In addition QTAIM[17] analysis implemented into multiwfn package[18] for BATABi complex (Fig S2, Table S1) has shown absence of interaction between cation and macrocyclic oxygens in contrast to BA3ABi. Furthermore regarding effectiveness of BATABi complex[11] we attempted to evaluate the changing of coordination environments of Bi3 + in the BATABi complex at different pH by EXAFS measurements of the solutions at pH 1, 3 and 11 (Figure 6, Table 5) in addition to pH 5-5.5 discussed above. Sample at pH 1 was used as reference to show the spectrum and obtained environment of free cation in presence of Cl- and H4BATA. The obtained distance 2.68 Å corresponds to Bi Cl bond in analogous compounds.[22] Light atoms at 2.33 Å can be attributed to the hydration shell. According to obtained distances decreasing and increasing of pH in relation to coordination at pH 5–5.5 causes extension of the number of the closest neighbors. In the case of pH 3 high CN at 2–3 Å is caused by an incomplete chelation of cation by donor atoms of ligand due to the protonation as it is evident from the speciation diagram (Figure 1b). Cation is partially surrounded by solvent molecules and reflections of plenty light Table 5. Structural parameters refined from fitting the EXAFS Bi L3-edge spectra for Bi-BATA solutions at varied pH. Coordination shell pH1 R, Å CN σ, Å2 pH3 R, Å CN σ, Å2 Bi C/N/O Bi C/N/O Bi C/N/O Bi C/N/O Bi C/N/O Bi C/N/O Bi-Cl 2.33 � 0.01 – 3.18 � 0.03 4.19 � 0.07 4.54 � 0.06 3 – 3 3 4 0.005 � 0.002 – 0.006 � 0.004 0.005 � 0.005 0.003 � 0.0005 2.32 � 0.02 2.53 � 0.02 3.20 � 0.03 4.78 � 0.03 4.64 � 0.07 4 4 4 3 7 0.006 � 0.002 0.004 � 0.001 0.007 � 0.003 0.002 � 0.004 0.005 � 0.011 2.68 � 0.03 1 0.009 � 0.006 – – – Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org 3349 pH5-5.5 R, Å CN σ, Å2 2.17 � 0.01 2.47 � 0.02 3.15 � 0.02 3.45 � 0.06 4.31 � 0.03 4.51 � 0.09 – 3 2 3 1 6 2 – 0.0045 � 0.0006 0.007 � 0.003 0.005 � 0.003 0.005 � 0.003 0.005 � 0.007 0.004 � 0.025 – pH11 R, Å CN σ, Å2 2.36 � 0.03 2.61 � 0.03 3.29 � 0.03 – 4.64 � 0.06 4 6 6 – 8 0.005 � 0.003 0.004 � 0.003 0.005 � 0.002 – 0.007 � 0.007 – – – © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 Figure 6. The fitted spectra for X-ray absorption of Bismuth complexes with H4BATA in the R- (a) and k –spaces (b). atoms are observed in the spectra. Upon pH increase at pH 5–5.5 both monoprotonated and deprotonated complexes exist with prevalence of the latter and cation introduces into the cavity in agreement with DFT calculated structure. Finally at highest pH 11 bismuth cation starts to hydrolyze and external OH groups facilitate gradual escape of the cation from surrounding by ligands’ donor atoms enhancing number of closest light atoms as in the case of lower pH 3.This approach experimentally confirmed structural reorganizations of the BATABi complex upon pH variation in agreement with speciation diagram (Figure 1) obtained by potentiometric studies: deprotonation of the ligand causes encapsulation of the cation while protonated forms as well as hydroxide specie force its escape out of the cage. This way is similar to H4DOTA complexation of cations[23,24] but significantly larger macrocycle of H4BATA provides much less restrictions for deprotonation and consequent embedment of bismuth inside the ligand’s cavity. Because the pH value could influence complexation, we evaluated labeling yields for several concentrations at different pH values, and no significant differences were observed. The labeling yield reached > 96 % at pH 6 and 77–100 μM of H4BATA and > 98 % at pH 7 and 60–100 μM of H3BA3A (Table 6). In vitro stability studies As a preliminary estimation before in vivo behavior analysis, challenge of the labeled compounds [207Bi]LBi with fetal bovine serum was carried out. Serum stability was studied using [207Bi] BATABi and [207Bi]BA3ABi radiolabeled solutions (radiochemical purity 97 %, 1 kBq of 207Bi) mixed with a fetal bovine serum. Serum stabilities were controlled after several time periods at 37 °C. Figure 7 shows the comparative protein binding of radionuclide upon challenging solutions with [207Bi]BA3ABi and [207Bi]BATABi. After 1 h of incubation (~ 1 decay period of 212,213Bi), the [207Bi] Radiolabeling The labeling efficiency of each studied ligand was determined using the TLC technique. The radiolabeling yields of H4BATA with 207Bi have previously been determined in a number of buffer systems[11] and are described here for comparative purposes, and these values as well as the radiolabeling yields of H3BA3A are presented in Table 6. Table 6. Percentage of incorporation of mean values � SD (n = 3). 207 Bi by H4BATA and H3BA3A: c(L), [μM] 10 30 60 100 1000 H3BA3A, pH 6.0 H3BA3A, pH 7.0 H4BATA, pH 6.1 H4BATA, pH 8.0 49 � 5 % 80 � 2 % 88 � 5 % 70 � 5 % 77 � 4 % 89 � 3 % 82 � 3 % 97 � 3 % 91 � 4 % 98 � 3 % 95 � 5 % 78 � 6 % 95 � 5 % 98 � 6 % Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org Figure 7. Transchelation of [207Bi]Bi3 + from complexes with H4EDTA,[21] H3BA3A, and H4BATA[11] by serum proteins. 3350 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 BA3ABi complex demonstrated 23 % 207Bi3+ protein binding and it was 38 % after 2 h (~ 2–3 decay periods of 212,213Bi). Additionally, there was no any significant 207Bi3 + protein binding for the case of [207Bi]BATABi even after 2 days, which indicated remarkable complex stability. This indicates that, in solution, it is more likely that the cation is not included in the cavity of H3BA3A, which is similar to its crystalline structure. For the BATABi complex, with regards to the higher organization on the outer coordination shells revealed by EXAFS, it is reasonable to suggest that inclusion of Bi3 + inside the macrocycle provides effective shielding of the cation from competing agents. Interestingly that despite much higher stability constant complex BA3ABi demonstrates completely the same trend for dissociation in 100-fold excess of serum proteins as complex with pyridine bisamide aza-18-crown-6 triacetate – L6 (logβ = 21.3).[21] This fact might reflect the moderate ability of “opened NOTA” backbone to hold Bismuth cation. In vivo studies Studying the behavior of a complex without a biological vector is an important step for understanding the stability of the complex inside the body. For the case of a stable complex, we expect to observe fast clearance of radioactive substances from the body without significant accumulation in normal tissues. When instability of the complexes is demonstarted in the body, their further conjugation with a biological vector is not reasonable. A total of 6 normal male BALB/c mice were used for the experiment with the [207Bi]BA3ABi complex. The radiolabeled ligand in 100 μL (pH 6.8, 0.01 M MES) was injected intraperitoneally. The mice were then euthanized at 1 and 6 h, and the blood and major organs were harvested, wet weighed, and the radioactivity was measured via gamma-spectrometry. The percent injected dose per gram (% ID/ g) was determined for each tissue (Table S3). All the obtained results are summarized in Figure 8 (a, b). The profile of the biodistribution of complex [207Bi]BA3ABi is similar to that of [207Bi]BATABi and [207Bi]DOTABi. The clearance was also carried out by the kidneys, and it is the organ with the highest radioactivity accumulation compared with other tissues for all complexes. However, much higher accumulation in the kidneys could be observed for the [207Bi]BA3ABi complex: 23 � 9 % of ID/g after 1 h, decreasing to 14 � 3 % of ID/g after 6 h. There was also some accumulation of the H3BA3A complex with 207Bi3 + in the liver 5.0 � 1.1 % of ID/g, and it decreased to 2.7 �0.5 % of ID/g after 6 h. Additionally, the biodistribution profile of the [207Bi]BA3ABi complex was also very similar to those of the [207Bi]EDTABi complex (Figure 8), which is known to be unstable in vivo despite its high stability constant. The accumulation of 207Bi was more pronounced at 6 h post injection (p.i.) of [207Bi]BA3Abi, which indicated slower clearance, and it was also at the same percentage as the complex with H4EDTA. This indicated that the [207Bi]BA3ABi complex released radionuclide in vivo similarly to [207Bi]EDTABi. It is important to notice that according to serum studies and neutral charge of LBi species we could expect the similar biodistribution of [207Bi]L6Bi[21] and [207Bi]BA3ABi. However comparison of biodistribution profiles of these complexes especially at 6 h Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org Figure 8. Biodistribution of 207Bi-labeled compounds: H4EDTA,[21] H3BA3A, H4BATA,[11] and H4DOTA[11] in normal BALB/c Mice at 1 h (a) and 6 h (b) after injection. p.i. clearly shows definitely less stability of [207Bi]BA3ABi in vivo revealing the important role of macrocyclic cavity for Bismuth chelation. The stability of [207Bi]BATABi differed significantly from [207Bi] BA3ABi. A [207Bi]BATABi complex like [207Bi]DOTABi is rapidly excreted from the body, and no significant accumulation of radioactivity in any healthy tissues was observed. The major radioactivity was found in the kidneys (~ 3 % ID/g), which reduced even further after 6 h. Based on this we can conclude that the Bi3 + complex of H4BATA has considerably higher stability than H3BA3A, and it has almost the same in vivo inertness as that of H4DOTA. After summarizing the serum and in vivo behavior of all the considered complexes, the aza-18-crown-6 cavity seems to be more appropriate for Bi3+ particularly in H4BATA, providing markedly better biodistribution of the formed chelate compared to [207Bi]BA3ABi regardless of the nearly similar pBi values. Furthermore, the tetraacetate ligands H4BATA and H4DOTA form complexes of nearly similar in vivo stability, and this demonstrates the effectiveness of negatively charged complex compounds such as BATABi and DOTABi . Additionally, based on the EXAFS data, H4BATA provides Bi3 + a stable long-range environment that allows this complex not only to be stable in the presence of external competitors, such as a 100-fold excess of serum protein, but also 3351 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 ensures fast in vivo clearance and avoids absorption of radionuclides in the organs. Taking into account the nearly similar ionic radii of Na + (1.02 Å) and Bi3+ (1.03 Å) CN6,[25] it is reasonable to suggest inclusion of the Bi3+ cation inside the cavity of the aza-15-crown-5 in H3BA3A. However, we do not observe such inclusion in the crystal structure of BA3ABi and clearly observed that the out-cage location of Bi3 + enables its transchelation, which provides moderate serum stability. It is likely that despite visual suitability of the cavity of the aza-15-crown-5 for Bi3+, additional factors such as the conformation of the azacrown-ligand and the 6 s2 lone pair of Bi3 + influence the resulting out-cage structure of the H3BA3A complex. In addition, by considering all the data,including described NETA and NPTA complexes[12] with Bi3 +, we can assume that 6 donor groups (3 amines and 3 acetates) in H3BA3A is not sufficient for formation of a highly stable complex with the bismuth cation in a competing medium. Conclusion In this work, we studied the complexation of two ligands H4BATA and H3BA3A with bismuth(III). The thermodynamic stability of bismuth(III) complexes of H4BATA and H3BA3A and their pH speciation were analyzed via a potentiometric method. The stability of the studied complexes was reasonably high based on the thermodynamic stability constants and pBi values. However, the results of the serum challenge and in vivo biodistribution show that the structural peculiarities significantly impacted the inertness of complexes in a competing medium. The macrocyclic cavity of aza-18-crown-6 appeared to be the most appropriate for the bismuth cation; in particular, it has four acetate arms and forms a negatively charged complex, which leads H4BATA to be the ideal ligand for Bi3 +. Based on EXAFS measurements in aqueous solutions, the bismuth cation chelated by H4BATA is effectively shielded from the surrounding competitors even in a living organism. The difference in the long-range neighbors of complexes with H3BA3A and H4BATA reflects the ability of the cation to be released from the complex in vitro and in vivo. The H4BATA ligand can be bifunctionalized via the benzene group, providing sufficient distancing of the biomolecule from the cation-coordinating sites to eliminate possible effects on labeling of the bioconjugate. Therefore, H4BATA can be considered as a promising candidate for conjugation with biological vectors (such as peptides and antibodies) for TAT with 212,213Bi radioisotopes. Experimental Section Synthesis of ligands All commercially available reagents were used without further purification. The ligands H3BA3A and H4BATA were prepared as described earlier.[11,26] Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org NMR study 1 H NMR spectra were recorded at 25 °C on Varian Inova 400 spectrometers. Chemical shifts are reported in parts per million (δ) relative to the deuterated solvent used as the internal reference (D2O δ = 4.75). The coupling constant J is given in Hertz. Spectral assignments were based in part on two-dimensional NMR experiments, COSY and NOESY (Figure S3, Figure S4). Samples of the Bi3 + complexes for the NMR measurements were prepared by dissolving the corresponding ligand and 1.5 eq. Bi(NO3)3 in D2O, followed by adjustment of the desired pD with small volumes of concentrated HClO4 or NaOH. Accurate pD measurements in D2O were obtained by direct reading in a D2O solution using a combined glass/AgCl electrode after appropriate calibration procedures using standard buffers. Numbering of the hydrogen and carbon nuclei used to describe the 1H NMR spectra is given in Figure 2. Synthesis of [Bi(BA3A)]: Bi(NO3)3 · 5H2O (5.3 mg, 0.011 mmol) was added to solution of H3BA3A (4.4 mg, 0.010 mmol) in H2O (1 mL), and the pH value of the solution was increased to 7–8. The solution was heated to 90°C with vigorous stirring for 10 min. After cooling to room temperature, the formed precipitate was filtered off. Slow evaporation of the mother solution led to the formation of single crystals suitable for X-ray analysis (5 mg, 71 % yield). 1H NMR (D2O, 400 MHz): 2.52 (t, 2H, H(7e), J = 13.3 Hz), 3.00 (t, 2H, H(6e), J = 13.3 Hz), 3.20 (d, 2H, H(5a), J = 12.9 Hz), 3.41 (t, 2H, H(5e), J = 13.3 Hz), 3.41 (d, 2H, H(8y), J = 16.4 Hz), 3.54 (d, 2H, H(7a), J = 13.3 Hz), 3.65 (d, 2H, H(6a), J = 15.2 Hz), 4.16 (t, 2H, H(4e), J = 10.9 Hz), 4.20 (s, 2H, H(10)), 4.49 (d, 2H, H(4a), J = 10.9 Hz), 6.64 (d, 2H, H(8x), J = 16.8 Hz), 7.05 (m, 4H, H(1,2)). 13C NMR (D2O, 400 MHz): 51.31 (C-7), 54.77 (C-10), 55.09 (C-6), 59.86 (C-5), 64.27 (C-8), 64.82 (C-4), 112.50 (C-2), 122.62 (C-1), 144.28 (C-3), 179.37 (C-9), 185.95 (C11). Synthesis of [Bi(BATA)]: Bi(NO3)3 · 5H2O (5.3 mg, 0.011 mmol) was added to solution of H4BATA (5.4 mg, 0.010 mmol) in H2O (1 mL), and the pH value of the solution was increased to 7–8. The solution was heated to 90°C with vigorous stirring for 10 min. After cooling to room temperature, the formed precipitate was filtered off. Slow evaporation of the mother solution did not lead to the formation of single crystals suitable for X-ray analysis. The product was isolated as a white powder (7.1 mg, 96 % yield). The 1H and 13C NMR spectrum is shown in Figures S6 and S7. Crystal data for [BA3ABi] Single crystals of the bismuth(III) complex BA3A3 were obtained from an aqueous solution via the slow evaporation method. Data collection for [BA3ABi] was performed on a Bruker SMART APEX II diffractometer (λ(MoKα) = 0.71073 Å, ω-scans, 2θ < 61.3°). The colorless crystals of C20H32BiN3O11 (M = 699.46) at 120 K were monoclinic (space group P21/c; a = 10.3041(4) Å; b = 14.4533(5) Å; c = 15.8501(6) Å; β = 103.9340(10)°; V = 2291.07(15) Å3; Z = 4 (Z’ = 1); dcalc = 2.028 g cm 3; μ = 7.763 mm 1; F(000) = 1376). Frames were integrated using the Bruker SAINT software package[27] with a narrow-frame algorithm. A semi-empirical absorption correction was applied with the SADABS[28] program using the intensity data of equivalent reflections. The intensities of 7,047 independent reflections (Rint = 0.0341) out of 30,487 collected were used for structure solution and refinement. The structure was solved using the dual-space approach with the SHELXT program,[29] and then refined via the full-matrix least-squares technique against F2hkl in an anisotropic approximation for non-hydrogen atoms with the SHELXL[30] program. The hydrogen atoms of water molecules were found from difference Fourier synthesis and refined in an isotropic approximation. All other hydrogen atoms were placed in the calculated positions and refined in the riding model with Uiso(H) 3352 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 parameters equal to 1.2 Ueq of the connected carbon atoms. The refinement converged to R1 = 0.0195 (calculated for 6,099 observed reflections with I > 2σ(I)), wR2 = 0.0404, and GOF = 1.024 for all independent reflections. Thermodynamic Stability Studies The setup for potentiometric titrations has been described before.[31] The titrant was a carbonate-free solution of NaOH analytical grade at ca. approximately 0.10 M. The exact concentration of NaOH solution was obtained by application of the Gran’s method upon titration of a previously standardized amounts of HCl and determining the equivalent point by the Gran’s method using the GLEE software.[32] The ionic product of water pKw = 13.78 at 25.0 °C and ionic strength of 0.10 � 0.01 M using KNO3 as a background electrolyte was kept constant. A stock solution of H4BATA and H3BA3A was prepared at ca. 0.01 M. An analytical solution of Bi(NO3)3 was prepared at 0.025 M in 0.7 M aqueous HCl to avoid metal hydrolysis. Potentiometric titrations were run with ca. 0.016 mmol of ligand in a total volume of 16.00 mL at 25.0 � 0.5 °C. Data were collected in the pH range 1.5–9.5/11.0. The [H + ] of the solutions was determined by measurement of the electromotive force of the cell, E = E0 + Q log[H + ] + Ej. The terms E0 and Q were determined by titration of a solution of known hydrogen-ion concentration at the same ionic strength. Each titration consisted of 80–100 equilibrium points and a minimum of two replicates were performed. The stability constants of the complexes were calculated from the electromotive force titration data with the Hyperquad software[32] and speciation diagrams were plotted using the Hyss software.[33] Experimental curves and other details are given in the ESI (Table S4). The overall equilibrium (formation) constants βHhL and βMmHhLl are defined by βHhL = [HhLl]/[H]h[L]l and βMmHhLl = [MmHhLl]/[M]m[H]h[L]l, while stepwise equilibrium constants are given by KMmHhLl = [MmHhLl]/[MmHh–1Ll][H] and correspond to the difference in log units between overall constants of sequentially protonated (or hydroxide) species. EXAFS measurement and data treatment The Bi L3-edge X-ray spectra for sample of aqueous solution containing Bi3 + + H3BA3A and Bi3 + + H4BATA were collected at the beamline “Structural Materials Science”[34] using the equipment of “Kurchatov Synchrotron Radiation Source” (Moscow, Russia). The storage ring with electron beam energy of 2.5 GeV and a current of 80–100 mA was used as the source of radiation. All the spectra were collected in the transmission mode using a Si (111) channelcut monochromator. EXAFS data (χexp(k)) were analyzed using the IFEFFIT data analysis package.[35] EXAFS data reduction used standard procedures for the pre-edge subtraction and spline background removal. The Fourier transformation (FT) of the k2weighted EXAFS functions χexp(k) was calculated over the range of photoelectron wave numbers k = 2–10.0 Å 1. The structural parameters, including interatomic distances (Ri), coordination numbers (Ni) and Debye-Waller factors (σ2), were found by the non-linear fit of theoretical spectra (eq.) to experimental ones. cðkÞ¼S20 n X Ni Fi ðkÞ i¼1 R2i k 2Ri e lðkÞ e 2s2i k2 sinð2kRi þFi ðkÞÞ The theoretical data were simulated using the photoelectron mean free path λ(k), amplitude Fi(k), and phase shift Fi(k) calculated ab initio using the program FEFF6.[36] Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org Computational details All the calculations were performed at the DFT[37,38] level of theory with hybrid density functional PBE0[39] and dispersion correction D3.[40] The atomic orbitals were described with def2-svp[41] basis set (in a case of Bi ion, the inner orbitals were included together with the core into an effective core potential def2-SD[42]). We also used ‘chain-of-spheres’ resolution of identity approximation (RIJCOSX[43–45]) for calculation of the electron repulsion integrals. The preliminary scan of potential energy surface and search of a global energy minimum was performed with a stochastic algorithm described in details in our previous work.[46] All the calculations were done in the orca[47] package. Protocol for radiolabeling with 207Bi All reagents and solvents were purchased from commercially available sources and used as received. The initial concentration of 207Bi was determined by radioactivity counting. To study the labeling efficiency, 30 μL containing 300 Bq [207Bi]BiCl3 in 0.1 M HCl was equilibrated with 1–1000 μM of the ligand in a 0.01 M MES buffer solution at room temperature in a plastic Eppendorf tube with a total volume of 300 μL. The average 207Bi concentration was 0.2 nM. After preparation, the mixtures were analyzed by thin layer chromatography (TLC). The TLC plates (cellulose on Al support, Sigma) were developed in a mixture of 0.9 % NaCl/10 mM NaOH. The plates were cut in half and the radioactivity on each part was measured via gamma spectrometry. The activity was quantified by the 570 keV gamma emission of 207Bi. Autoradiography was performed with a Perkin Elmer Cyclone Plus Storage Phosphor System and the associated software. Percentages of 207 Bi incorporation were deduced from the ratio of the radioactivity intensities at Rf = 0.9 � 0.1 (which corresponds to [207Bi]BA3ABi or [207Bi] BATABi) and Rf ~ 0 (which corresponds to the remaining 207Bi salts). The average values of at least two experiments were used (error estimated: � 6%). In vitro stability study of [207Bi]BA3ABi A solution of the radiolabeled complex [207Bi]BA3ABi with ca. 0.6 mM of ligand was buffered at pH 6.8 with a 0.01 M MES solution at room temperature (RT) (radiochemical purity: 97 %). Fetal bovine serum (triple 0.12 μm sterile filtered) was purchased from HyClone (South Logan, Utah), and all required storage measures were followed. The ratio of the complex [207Bi]BA3ABi solution and serum volume was established at 1 :100 at 37 °C. After 1, 5, 15, 30, 60, 120, 240 min, 1 day, and 2 days an aliquot of each sample was taken; after protein precipitation via ethanol radioactivity of each supernatant, an aliquot of the initial sample with the same volume and geometry was measured by gamma-spectrometry and the percentage of the complex was determined. Additionally, aliquots of the supernatants were analyzed by TLC on cellulose plates with an Al support. In vivo biodistribution study of [207Bi]BA3ABi All in vivo experiments were performed in accordance with EU Directive 2010/63/EU for animal experiments and approved by The Bioethics Commission of Lomonosov Moscow State University (meeting number 125-d held on 28. 01. 2021, protocol No122-a). For the in vivo experiments a total of 12 normal male BALB/c mice were used. The mice were housed with 12-h light/dark cycles with access to water and food ad libitum. A solution of 1 mM H3BA3A and 207Bi3+ (2.0–2.5 kBq) was buffered at pH 6.8 with 0.01 M MES at RT and diluted in a sterile isotonic solution. The radiochemical purity according to TLC reached 97 %. Then, 100 μl of solution was administered to the mice (weight 28–37 g) via intraperitoneal injection as previously described.[48] The mice (3 per data point) were euthanized at 1 and 6 h 3353 © 2021 Wiley-VCH GmbH Full Papers doi.org/10.1002/ejic.202100256 by cervical dislocation. Blood was collected right after euthanasia and mixed with 100 μl of heparin solution. The major organs were harvested and the blood and peritoneal liquid was washed off with 0.9 % NaCl solution, the organs were then wet-weighed and the radioactivity of each was measured using a γ-scintillation counter (with a HPGe-detector GR3818 Canberra Ind.). The percent injected dose per gram (% ID/g) was determined for each tissue, and the values presented are the mean and standard deviation. Deposition Number 2058062 (for [BA3ABi]) contains the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre and Fachinformationszentrum Karlsruhe Access Structures service www.ccdc.cam.ac.uk/structures. Acknowledgements The reported study was funded by the Russian Science Foundation project no. 18-73-10035 (synthesis of ligands and complexation study). E.A.K., A.L.T. and S.N.K. acknowledge support by the Russian Ministry of Science and Education under grant No 075-152019-1891 (EXAFS measurements). X-ray diffraction experiment was performed with financial support from the Ministry of Science and Higher Education of the Russian Federation using equipment at the Center for Molecular Composition Studies of INEOS RAS. The research was carried out using the equipment of the shared research facilities of HPC computing resources at Lomonosov Moscow State University (DFT calculations). Conflict of Interest The authors declare no conflict of interest. Keywords: Benzoazacrown · Bismuth · complexation · EXAFS · DFT [1] L. Marcu, E. Bezak, B. J. Allen Crit. Rev. Oncol. Hematol. 2018, 123, 7–20. [2] D. Cordier, F. Forrer, F. Bruchertseifer, A. Morgenstern, C. Apostolidis, S. Good, J. Müller-Brand, H. Mäcke, J. C. Reubi, A. Merlo Eur. J. Nucl. Med. Mol. Imaging 2010, 37, 1335–1344. [3] A. Morgenstern, C. Apostolidis, C. Kratochwil, M. Sathekge, L. Krolicki, F. Bruchertseifer Curr. Radiopharm. 2018, 11, 200–208. [4] E. W. Price, C. Orvig Chem. Soc. Rev. 2014, 43, 260–290. [5] A. Morgenstern, C. Apostolidis, F. Bruchertseifer Semin. Nucl. Med. 2020, 50, 119–123. [6] B. V. Egorova, O. A. Fedorova, S. N. Kalmykov Russ. Chem. Rev. 2019, 88, 901–924. [7] T. I. Kostelnik, C. Orvig Chem. Rev. 2019, 119, 902–956. [8] M. W. Brechbiel, C. G. Pippin, T. J. Mcmurry, D. Milenic, M. Roselli, D. Colcherb J. Chem. Soc. Chem. Commun. 1991, 1169–1170. [9] A. Hu, S. N. MacMillan, J. J. Wilson J. Am. Chem. Soc. 2020, 142, 13500– 13506. [10] N. A. Thiele, V. Brown, J. M. Kelly, A. Amor-Coarasa, U. Jermilova, S. N. MacMillan, A. Nikolopoulou, S. Ponnala, C. F. Ramogida, A. K. H. Robertson, C. Rodríguez-Rodríguez, P. Schaffer, C. Williams, J. W. Babich, V. Radchenko, J. J. Wilson Angew. Chem. Int. Ed. 2017, 56, 14712–14717; Angew. Chem. 2017, 129, 14904–14909; Angew. Chem. 2017, 129, 14904–14909. Eur. J. Inorg. Chem. 2021, 3344 – 3354 www.eurjic.org [11] E. V. Matazova, B. V. Egorova, E. A. Konopkina, G. Y. Aleshin, A. D. Zubenko, A. A. Mitrofanov, K. V. Karpov, O. A. Fedorova, Y. V. Fedorov, S. N. Kalmykov MedChemComm 2019, 10, 1641–1645. [12] H.-S. Chong, D. E. Milenic, K. Garmestani, E. D. Brady, H. Arora, C. Pfiester, M. W. Brechbiel Nucl. Med. Biol. 2006, 33, 459–467. [13] H. S. Chong, H. A. Song, N. Birch, T. Le, S. Lim, X. Ma Bioorg. Med. Chem. Lett. 2008, 18, 3436–3439. [14] E. Csajbok, Z. Baranyai, I. Banyai, B. Erno, R. Kiraly, A. Mueller-Fahrnow, J. Platzek, B. Raduechel, M. Schaefer Inorg. Chem. 2003, 42, 2342–2349. [15] J. F. Desreux Inorg. Chem. 1980, 19, 1319–1324. [16] I. J. Bruno, J. C. Cole, M. Kessler, J. Luo, W. D. S. Motherwell, L. H. Purkis, B. R. Smith, R. Taylor, R. I. Cooper, S. E. Harris, A. G. Orpen J. Chem. Inf. Comput. Sci. 2004, 44, 2133–2144. [17] R. F. W. Bader Atoms in Molecules: A Quantum Theory. Oxford Univ. Press; 1990. [18] T. Lu, F. Chen J. Comput. Chem. 2012, 33, 580–592. [19] C. F. Baes, R. E. Mesmer The Hydrolysis of Cations. New York: John Wiley & Sons, Inc; 1976. [20] V. Stavila, R. L. Davidovich, A. Gulea, K. H. Whitmire Coord. Chem. Rev. 2006, 250, 2782–2810. [21] B. V. Egorova, E. V. Matazova, A. A. Mitrofanov, G. Y. Aleshin, A. L. Trigub, A. D. Zubenko, O. A. Fedorova, Y. V. Fedorov, S. N. Kalmykov Nucl. Med. Biol. 2018, 60, 1–10. [22] B. E. Etschmann, W. Liu, A. Pring, P. V. Grundler, B. Tooth, S. Borg, D. Testemale, D. Brewe, J. Brugger Chem. Geol. 2016, 425, 37–51. [23] J. Moreau, E. Guillon, J. C. Pierrard, J. Rimbault, M. Port, M. Aplincourt Chem. A Eur. J. 2004, 10, 5218–5232. [24] Y. H. Jang, M. Blanco, S. Dasgupta, D. a. Keire, J. E. Shively, W. a G. Iii, J. Am. Chem. Soc. 1999, 6142–6151. [25] R. D. Shannon Acta Crystallogr. Sect. A 1976, 32, 751–767. [26] G. Y. Aleshin, B. V. Egorova, A. B. Priselkova, L. S. Zamurueva, S. Y. Khabirova, A. D. Zubenko, V. A. Karnoukhova, O. A. Fedorova, S. N. Kalmykov Dalton Trans. 2020, 49, 6249–6258. [27] Bruker 2013 v.8.34 A. [28] L. Krause, R. Herbst-Irmer, G. M. Sheldrick, D. Stalke J. Appl. Crystallogr. 2015, 48, 3–10. [29] G. M. Sheldrick Acta Crystallogr. Sect. A 2015, 71, 3–8. [30] G. M. Sheldrick Acta Crystallogr. Sect. C 2015, 71, 3–8. [31] Y. V. Fedorov, O. A. Fedorova, S. N. Kalmykov, M. S. Oshchepkov, Y. V. Nelubina, D. E. Arkhipov, B. V. Egorova, A. D. Zubenko Polyhedron 2017, 124, 229–236. [32] P. Gans, A. Sabatini, A. Vacca Talanta 1996, 43, 1739–1753. [33] L. Alderighi, P. Gans, A. Ienco, D. Peters, A. Sabatini, A. Vacca, V. Maragliano, Coord. Chem. Rev. 1999, 184, 311–318. [34] A. A. Chernyshov, A. A. Veligzhanin, Y. V. Zubavichus Nucl. Instrum. Methods Phys. Res. Sect. A 2009, 603, 95–98. [35] M. Newville J. Synchrotron Radiat. 2001, 8, 96–100. [36] S. I. Zabinsky, J. J. Rehr, A. Ankudinov, R. C. Albers, M. J. Eller Phys. Rev. B 1995, 52, 2995–3009. [37] P. Hohenberg, W. Kohn Phys. Rev. 1964, 136, B864-B871. [38] W. Kohn, L. J. Sham Phys. Rev. 1965, 140, A1133-A1138. [39] M. Ernzerhof, G. E. Scuseria J. Chem. Phys. 1999, 110, 5029–5036. [40] S. Grimme, J. Antony, S. Ehrlich, H. Krieg J. Chem. Phys. 2010, 132. [41] A. Schäfer, H. Horn, R. Ahlrichs J. Chem. Phys. 1992, 97, 2571–2577. [42] X. Cao, M. Dolg J. Chem. Phys. 2001, 115, 7348–7355. [43] J. L. Whitten J. Chem. Phys. 1973, 58, 4496–4501. [44] R. A. Kendall, H. A. Früchtl Theor. Chem. Accounts Theory, Comput. Model. (Theoretica Chim. Acta) 1997, 97, 158–163. [45] F. Neese, F. Wennmohs, A. Hansen, U. Becker Chem. Phys. 2009, 356, 98– 109. [46] A. Mitrofanov, N. Andreadi, P. Matveev, G. Zakirova, N. Borisova, S. Kalmykov, V. Petrov J. Mol. Liq. 2021, 325, 115098. [47] F. Neese WIREs Comput. Mol. Sci. 2018, 8. [48] E. Machholz, G. Mulder, C. Ruiz, B. F. Corning, K. R. Pritchett-Corning J. Visualization 2012, 1–8. Manuscript received: March 29, 2021 Revised manuscript received: July 26, 2021 Accepted manuscript online: July 27, 2021 3354 © 2021 Wiley-VCH GmbH