Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2023 Nov 10;62(50):20733–20744. doi: 10.1021/acs.inorgchem.3c02211

Metal–Ligand Interactions in Scandium Complexes with Radiopharmaceutical Applications

Attila Kovács 1,*
PMCID: PMC10731654  PMID: 37949439

Abstract

graphic file with name ic3c02211_0009.jpg

The radioisotopes of scandium (43Sc, 44Sc, and 47Sc) are potential candidates for use in imaging and therapy both separately and as elementally matched pairs for radiotheranostics. In the present study the bonding interactions of Sc3+ with 18 hepta- to decadentate ligands are compared using density functional theory (DFT) calculations. The bonding analysis is based on the natural bond orbital (NBO) model. The main contributions to the bonding were assessed using natural energy decomposition analysis (NEDA). Most of the ligands have anionic character (charges from 2– to 8−); thus the electrical term determines the major differences in the interaction energies. However, interesting features were found in the covalent contributions manifested by the ligand → Sc3+ charge transfer (CT) interactions. Significant differences could be observed in the energetic contributions of the N and O donors to the total CT.

Short abstract

Density functional theory calculations in conjunction with energy decomposition analysis were used to uncover the bonding characteristics of Sc3+ complexes. The analysis confirmed the comparable stabilities of complexes with hepta- and octadentate ligands. The primary ligand donors are anionic oxygen and aromatic nitrogen atoms.

Introduction

Scandium radioisotopes (43Sc, 44Sc, and 47Sc) are important candidates for use in imaging and radioimmunotherapy. Their main advantages are emitted low- or middle-energy β-particles and conveniently short half-lives.

44Sc produces middle-energy β+ radiation with a high branching ratio (Eβ+av = 632 keV, 94%) and a physical half-life of t1/2 = 3.97 h.1 This large t1/2 compared with those of the most common β+ emitters, 18F (1.83 h) and 68Ga (1.13 h), facilitates acquiring positron emission tomography (PET) images at later time points than with the latter conventional isotopes. The advantages are a higher image quality,2 convenient investigation of pharmacokinetics of slower-circulating bioconjugates, and cost-effective centralized production and regional distribution of the isotope.

47Sc is a high branching β emitter (Eβ av = 162 keV, 100%) with a significant half-life (t1/2 = 3.35 d).1 This low-energy radiation is appropriate for treatment of small tumors as well as cancer metastasis,3 making 47Sc a promising competitor of the clinically established 177Lu radionuclide (Eβav = 134 keV, t1/2 = 6.65 d).1 In addition, the emitted γ-rays (Eγ = 159 keV, Iγ = 68%) have an ideal energy for single photon emission computed tomography (SPECT) imaging.3

The above Sc isotopes are perfectly suitable for radiotheranostics, in which method the same molecular targeting vectors are labeled partly with a diagnostic, partly with a therapeutic radionuclide.4,5 Currently, the pair of 68Ga and 177Lu is used in clinics for PET imaging and therapy.6,7 However, the two elements have different coordination chemistries8 leading to somewhat different in vivo kinetics and receptor binding affinity, thus decreasing the therapeutic efficiency.9 From this point of view, the ideal solution would be the combination of different radioisotopes of the same chemical element like the matched pair 44Sc and 47Sc.

Additional useful Sc radioisotopes include 43Sc (t1/2 = 3.89 h, Eβ +av = 476 keV, 88%),1 which is free from high-energy γ emission and therefore can be suitable for PET imaging.10 Last, the 44mSc isotope (t1/2 = 58.6 h), an isomeric state of 44Sc, has been suggested as a potential in vivo 44Sc generator allowing longer pharmacokinetic studies.11

Currently, the primary chelating ligand for Sc3+ is 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetate (DOTA).12 Disadvantages of this ligand are, however, its high affinity to other metal ions13 and the slow radiolabeling kinetics. There is a high demand for more suitable ligands which could increase the efficiency of Sc radiotheranostics.

In the present work the structural and bonding properties of Sc3+ with various cyclic and acyclic chelating ligands are studied in the isolated molecules using DFT calculations. The small ionic radius of Sc3+ may imply preferred interaction with small ligands with coordination numbers (CNs) of <8. However, a comparison of available stability constants of Sc3+ with octadentate and smaller hexadentate ligands proved the preference for octacoordination.14,15 On the other hand, heptacoordination can be reasonably stable.16 Therefore, in the present study the bonding characteristics with selected potentially octa- or heptadentate ligands are assessed. They contain various donor groups with the main donor atoms including neutral N and O and anionic (mostly carboxylate) O. The metal–ligand interactions are analyzed using the natural bond orbital (NBO) model.1719

Ligands

The cyclen-based ligands presented in Figure 1 compose an important group of chelators in radiopharmaceutics. The 12-membered ring of cyclen (1,4,7,10-tetraazacyclododecane) forms a semirigid basis of the ligand cavity with (generally four) pendant arms filling the conformational space around the captured metal ions. The cyclen ring itself can establish tetradentate coordination with its four N donor atoms.

Figure 1.

Figure 1

Cyclen-based ligands probed with Sc3+ in the present study. H3PSMA′ and H3DOTATATE′ are truncated models of the very large original ligands.

The most known ligand from the cyclen-based group is H4DOTA (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid). Due to its good performance for metal ions, it achieved the title of “gold standard” for a number of trivalent radioisotopes, including 111In, 177Lu, 86/90Y, 213Bi, 225Ac, and 44/47Sc.12 The crystal structure of the Sc complex as well as the thermodynamic stability in solution (log K = 30.79) was reported by Pniok et al.20 The ligand is generally applied in alkaline aqueous solutions; thus the complexes formed between DOTA4– and trivalent metal ions have a charge of 1–. In the complexes generally octacoordination is established through the cyclen N and carboxylate O donors. However, due to the small size of the pendant arms the metals are not fully surrounded, thus allowing coordination of an additional small ligand at the partly open side between the COO groups of the complex. In aqueous solutions this is generally a H2O molecule;21,22 its role in complexation of the Sc3+ ion is probed here too.

Three extended H4DOTA derivatives are included in the present study in order to estimate the effect of popular linker moieties (which link the complex with the targeting vector) on the bonding properties. These linkers are substituents in the outer sphere of the tetraazacyclododecane ring in order to avoid direct interference with metal binding. We probed the p-SCN-Bn (4-isothiocyanatophenyl) methyl substituent23 in position 2 as well as truncated H3PSMA and H3DOTATATE ligands, in which one COOH group of H4DOTA is replaced by (CO)NH–CH2–C6H11 and (CO)NH–CH(−CH2–C6H5)–(CO)NH–CH3 groups, respectively (cf. Figure 1). Note that the kinetic and thermodynamic stabilities of these latter complexes were found to be slightly lower than those of the H4DOTA ones, but still sufficiently high for in vivo application.

Neutral ligands are represented in this study by 1,4,7,10-tetrakis(3-pyridazylmethyl)-1,4,7,10-tetraazacyclododecane (Lpyd). The N donors of the pendant arms facilitate octadentate coordination to the metal ion, whereas the large size of the arms can shield the metal and sterically hinder a coordination of additional small molecules (e.g., H2O). This ligand was shown recently to be effective for the borderline Lewis acid Bi3+,24 and it would be of interest to see how the bonding occurs in detail with the harder Lewis acid Sc3+.

The H2MeDO2PA (6,6′-((4,10-dimethyl-1,4,7,10-tetraazacyclododecane-1,7-diyl) bis(methylene))dipicolinic acid) ligand has only two pendant arms, but it achieves octacoordination by the pyridine N and carboxylate O donors of the picolinic acid arms. This ligand was found to be superior to H4DOTA in complexing Bi3+;25,26 thus it would be interesting to see the relation in the case of Sc3+.

The H4DOTPA (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetrapropionic acid) molecule contains pendant arms elongated by a methylene group, and in this way it has a larger flexibility with respect to H4DOTA. This larger flexibility seemed to be a slight advantage for bonding with the Bi3+ and Ac3+ ions as compared to H4DOTA.23

In the H8DOTMP ligand (((1,4,7,10-tetraazacyclododecane-1,4,7,10-tetrayl)tetrakis(methylene))tetraphosphonic acid) the carboxylate oxygen donors of H4DOTA are replaced by phosphonate ones. This chelator has been considered for a bone-seeking agent with various radioisotopes (111In, 166Ho, 153Sm, 177Lu, 212Bi).2729 In this study both the form with the fully deprotonated (charge 8−) and that with the half-deprotonated (charge 4−) tetraphosphonic acid were probed. The two forms facilitate a comparison of the coordination interactions in strong and moderately alkaline environments.

Figure 2 presents some additional cyclic ligands probed in the present study. From them H4AAZTA (1,4-bis(carboxymethyl)-6-(bis(carboxymethyl))-amino-6-methylperhydro-1,4-diazepine) already has a history with Sc3+, forming highly stable complexes (log K = 27.7).16 This heptadentate ligand coordinates by four carboxylate O and three tertiary amine N donors. The backbone is flexible enough for room-temperature radiolabeling, while the semirigid cyclic moiety results in inertness comparable to DOTA.

Figure 2.

Figure 2

Other applied cyclic ligands. From H2bispa2 two heptadentate ligands (L2, L3) were derived by modification of the top picolinic acid moiety; see text.

H2macropa (N,N′-bis((6-carboxy-2-pyridil)methyl)-4,13-diaza-18-crown-6) is a recently introduced competitor for H4DOTA for actinides and lanthanides.3032 This decadentate ligand can efficiently capture metal ions with the N and O donors in the crown and picolinate arms. Though H2macropa was reported to prefer larger-size metals,30 a comparative bonding analysis of its complex with Sc3+ can be of interest.

Bispidine-based ligands showed high complex stability, inertness, and fast complexation kinetics33,34 and can simply be conjugated to peptides and antibodies.3537 Another advantage of this ligand family is the easy variation of O- and/or N-donor sets on the rigid bispidine scaffold producing CN = 4–8.33,34,38,39 The recently introduced octadentate H2bispa2 (6,6′-((9- hydroxy-1,5-bis(methoxycarbonyl)-2,4-di(pyridin-2-yl)-3,7-diazabicyclo[3.3.1]nonane-3,7-diyl) bis(methylene))dipicolinic acid) provided excellent performances with medically relevant metal ions In3+, Bi3+, Lu3+, and La3+ including fast complex formation under ambient conditions.33,40 Two other picolinate-/acetate-based heptadentate bispa-type ligands40 are also included in the present study in order to check whether they might be more suitable for the small Sc3+ ion. They are derived from H2bispa2 by removing the COOH group from the top picolinic acid (see Figure 2) and by replacing this top picolinic acid with a COOH group, denoted as HL2 and H2L3, respectively.

Acyclic chelators are kinetically somewhat less stable than macrocyclic ones (like H4DOTA), but they have the advantage that the formation of their complexes at room temperature is much faster. In the present study five octadentate acyclic chelators (Figure 3) were probed. Three of them have been reported for fast complex formation with Sc3+ and very high labeling yields.41

Figure 3.

Figure 3

Acyclic ligands probed with Sc3+ in the present study.

H4EGTA (ethylene glycol-bis(2-aminoethyl ether)-N,N,N′,N′-tetraacetic acid, also known as egtazic acid) showed very good complexation properties to Sc3+, outperforming several acyclic ligands, and thus was suggested to be a suitable precursor for Sc radiopharmaceuticals.41 The strong complexes are formed by binding via the four carboxylic and two ether O atoms, completing the coordination by two tertiary amine N’s.

H4BAPTA (6-dentate 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid) is compatible with H4EGTA for the binding of Ca2+ ions42 but for Sc3+ showed a worse performance than its parent H4EGTA.41 Surely, the structure of H4BAPTA is less flexible due to substitution of two ethylene moieties by benzyl rings. The present bonding analysis can elucidate the details behind the weaker complexation.

H5DTPA (1,1,4,7,7-diethylenetriaminepentaacetic acid) belongs to the most efficient acyclic chelators12 due primarily to the five carboxylate O donors and efficient encapsulation of metal ions. The coordination sphere is completed by three tertiary amine N’s (cf. Figure 3). This ligand has been found to exert great affinity for Sc3+ (log K = 27.43)20 and was suggested to be a suitable precursor for Sc radiopharmaceuticals.41 The observed dissociation aptitude in solution was alleviated in its more preorganized (semirigid) derivative, CHX-A″-H5DTPA (2-aminoethyl-trans-[S,S]-cyclohexane-1,2-diamine-N,N,N′,N″,N″-pentaacetic acid) while maintaining rapid labeling kinetics. The latter ligand is now widely used with radioimmunoconjugates.12,43

3,4,3-LI(1,2-HOPO), N,N′-1,4-butanediylbis(N-(3-(((1,6-dihydro-1-hydroxy-6-oxo-2-pyridinyl)carbonyl)amino)propyl)-1,6-dihydro-1-hydroxy-6-oxo-2-pyridinecarboxamide (denoted as HOPO below), has demonstrated strong affinity for hard metal 3+ and 4+ ions44,45 including Sc3+,46 with log KScL = 25.16.47 Advantages of this ligand include strong bonding with the hard O-donating hydroxypyridinone groups, complex formations at lower pH than H4DOTA, and fast binding kinetics.

In the discussion below the abbreviations of the deprotonated ligands are used according to the complexing anionic forms (DOTA, etc.).

Results and Discussion

Structural Characteristics

The octacoordinated Sc3+ ion has a relatively small effective ionic radius: 0.870 Å.48 It is smaller by 0.256 Å than Pr3+ (1.126 Å48), where the switch of the DOTA conformers in complexes of the lanthanide (Ln) row occurs. Trivalent ions with ionic radii larger than that of Pr3+ are known to prefer the twisted square antiprismatic (TSAP) conformer of DOTA, whereas the smaller lanthanides favor the square antiprismatic (SAP) one.8,4951 These two conformers are presented in Figure 4. The SAP structure has slightly smaller cavity and thus can form stronger interactions with the smaller ions. It should be noted that in the crystal of K[Sc(DOTA)](H6DOTA)Cl2·4H2O the TSAP conformer has been identified, which may be stabilized by the interaction of the four COO groups with K+ linking the H6(DOTA)Cl2 moiety to Sc(DOTA).20

Figure 4.

Figure 4

Two conformers of the Sc(DOTA) complex omitting hydrogen atoms for clarity. Moieties toward the back are gradually faded.

In agreement with expectations, in the present B3LYP/6-31G** calculations the SAP conformer of ScDOTA proved to be more stable by 15.2 kJ/mol than the TSAP conformer. A similar stability relation was computed for the DOTA-p-NCS-Bn, DOTATATE′, and PSMA′ ligands. In contrast, the other cyclen-based ligands with larger pendant arms, DOTPA, DOTMP, and Lpyd, preferred the TSAP conformer by 15–50 kJ/mol at the B3LYP/6-31G** level.

Another ligand with significant conformational varieties is macropa, from which the Δ(λδλ)(λδλ) and Δ(δλδ)(δλδ) conformers, denoted by LDL and DLD, respectively, were probed with Sc3+ in the present study. For Ln, the LDL form was found to form the most stable complexes with smaller Ln3+ (Lu3+),30 whereas DLD prefers larger ions (e.g., La3+, Ac3+).30,52 In agreement with the above trend, the Sc(macropaLDL)+ complex was computed at the B3LYP/6-31G** level to be more stable by 15.9 kJ/mol than the Sc(macropaDLD)+ conformer. The distortion of the C2 symmetry to C1 in the macropaDLD complex is likely the consequence of the lesser ability of this macropa conformer to efficiently encapsulate Sc3+.

Some of the probed ligands are known to allow the coordination of an additional small ligand, like H2O. This is due to insufficient encapsulation of the metal ion, leaving a site with considerable free space in the coordination sphere of the metal. The optimized structures of the present Sc3+ complexes were thoroughly inspected from this point of view. The conclusion of this visual inspection is that small ligands can easily coordinate to the significant free site of the cyclen-type and AAZTA complexes, whereas the other ligands encapsulate Sc3+ more tightly, leaving not enough space for attack by a small ligand. The effect of a H2O ligand at the ninth coordination site on the bonding properties was assessed in the Sc(DOTASAP) complex; similar effects can be expected for the other related complexes too.

Table 1 compiles the symmetries of the optimized structures. The well-known C4 symmetry of DOTA complexes8,23,4951,53,54 is preserved in the derivatives with Lpyd, DOTPA, and DOTMP. Similarly to literature data on Bi3+ complexes, optimized structures with C2 symmetry were obtained for the MeDO2PA, macropaLDL, and BAPTA complexes with Sc3+.23 In contrast, the C2 symmetry reported for Ln(macropaDLD)+ complexes30,52 was destroyed with the smaller Sc3+ ion.

Table 1. Selected Structural Characteristicsa.

complex CN symmetry Sc–Oav Sc–Nav
DOTATSAP 8 C4 2.099 (4) 2.655 (4)
DOTASAP 8 C4 2.099 (4) 2.617 (4)
DOTASAP + H2O 9 C2 2.157 (5) 2.740 (4)
DOTA-p-NCS-Bn 8 C1 2.097 (4) 2.631 (4)
PSMA′ 8 C1 2.122 (4) 2.587 (4)
DOTATATE′ 8 C1 2.122 (4) 2.573 (4)
Lpyd 8 C4 2.422 (8)
MeDO2PA 8 C2 2.053 (2) 2.455 (6)
DOTPA 8 C4 2.069 (4) 2.747 (4)
DOTMP 8 C4 2.097 (4) 2.661 (4)
DOTMP″ 8 C4 2.059 (4) 3.244 (4)
AAZTA 7 C1 2.091 (4) 2.468 (3)
macropaLDL 10 C2 2.339 (6) 2.730 (4)
bispa2 8 C1 2.048 (2) 2.483 (6)
L3 7 C1 2.006 (2) 2.382 (5)
L2 7 C1 1.985 (1) 2.347 (6)
EGTA 8 C1 2.230 (6) 2.474 (2)
BAPTAb 8 C2 2.016 (6) 2.687 (2)
DTPA 8 C1 2.165 (5) 2.599 (3)
CHX-A″-DTPA 8 C1 2.161 (5) 2.597 (3)
HOPO 8 C1 2.231 (8)
a

The Sc3+ complexes are denoted by the ligand names given in the first column. The PSMA′ and DOTATATE′ notations mean the truncated ligands to 73 and 77 atoms, respectively; DOTMP and DOTMP″ mean the DOTMP4– and DOTMP8– ligands, respectively. The average Sc–O and Sc–N bond distances are given in angstroms. The number of coordinating O and N atoms are given in parentheses.

b

Two Sc–O distances around 3.4 Å in the Sc(BAPTA) complex are too large for a significant interaction. They were excluded from the average data in order to reflect the very strong character of the relevant Sc–O interactions.

The average Sc–O and Sc–N distances are given in Table 1, whereas the individual bond distances are compiled in Figure 5. The values depicted in Figure 5 as well as the Cartesian coordinates of the optimized structures are given in the Supporting Information. We can distinguish O donors with formal negative charge (carboxylate in most ligands and deprotonated hydroxypyridinone in HOPO, denoted as O) and formally neutral (ether O in macropa, EGTA, BAPTA; C=O in PSMA′, DOTATATE′, and HOPO; denoted as O). The formally neutral N donors can be separated as aliphatic (in cyclen and macropa rings, denoted as Ncyc) or aromatic (in the picolinate and pyridine/piperidine groups, denoted as Nar).

Figure 5.

Figure 5

Distributions of Sc–O and Sc–N distances in the studied complexes. For the notation of ligands, see footnote a of Table 1. The red and blue lines indicate the average Sc–O and Sc–N values, respectively.

The primary Sc–O distances are around 2.1 Å, whereas those of the Sc–Nar and Sc–Ncyc are around 2.3 and 2.6 Å, respectively (cf. Figure 5). Note that the two Sc–O distances around 3.4 Å in the Sc(BAPTA) complex are too large for a significant interaction; therefore they were excluded from the average data.

The data indicate generally stronger interactions with the O donors than with N, which is reasoned by the anionic character of most O donors in the complexes. The relatively large Sc–Oav values in the macropa and EGTA complexes can be attributed to the neutral ether O’s in these ligands, which form very weak interactions with Sc3+. The carbonyl oxygens in the HOPO ligand establish intermediate interactions due partly to their better donor abilities with respect to ether O and to their more flexible position in the ligand. Moreover, in HOPO the C=O groups have Sc–O distances comparable to those of the anionic N–O groups implying similar interaction strengths of the two types of O donors.

From the two types of Sc–N interactions, the ones with Nar have generally shorter distances (see Figure 5). As the Sc–N interactions are generally weaker than the Sc–O ones, the latter ones play the major role in the structure of the complexes. Consequently, the bond distances of the weaker Sc–N interactions show a larger variation in Figure 5.

The H2O ligand at the ninth coordination site slightly increases the Sc–Ncyc distance with respect to that in the parent Sc(DOTA) complex as it pulls Sc3+ slightly closer to the ninth coordination site.

The effects of pendant arms on the Sc–Ncyc interaction can be analyzed in the first half of Figure 5 (from DOTA to DOTMP). This effect on the Sc–Ncyc bond distances is generally small. A striking exception can be observed in the Sc(DOTMP)5– molecule, where the Sc3+ ion is pulled significantly away from the cyclen moiety by the strong electrostatic interactions with the PO32– groups. The Sc–N distance of 3.244 Å suggests marginal interaction with N; thus this complex can rather be referred to as tetracoordinated Sc3+. (This marginal Sc–N interaction is reflected also in the CT data, vide infra.) In contrast, the Sc–N distances in the tetraprotonated Sc(DOTMP) are in the range of those in the complexes with DOTA and other DOTA derivatives. A small but significant increase in the Sc–Ncyc distance can be observed also in Sc(DOTPA), where the pendant arm increased by a CH2 group results in an increased cavity affecting primarily the positions of the O donors.

Comparison of the Sc–ligand bond distances in the complexes with DOTASAP and DOTATSAP can provide a clue to the stability relations. The Sc–O distances are the same whereas the Sc–N distances are shorter by 0.04 Å in the Sc(DOTASAP) complex (cf. Table 1 and Figure 5). This refers to comparable Sc–O interactions in the two conformers and indicates that the larger stability of the Sc(DOTASAP) complex is enforced by its stronger Sc–N interactions. The latter feature is facilitated by the smaller cavity of the SAP structure.

The effect of water solvent on the Sc–donor distances was investigated on selected complexes using the polarizable continuum model (PCM).55,56 A figure comparing the average Sc–N and Sc–O distances of the isolated and solvated structures is given in the Supporting Information (Figure S1). The general effect of the water solvent is a slight increase of the Sc–O distances and a minor decrease of the Sc–N ones. The longest Sc–N bonds (with average distances above 2.6 Å) shortened more significantly. The polar solvent changes the polarization of the carboxylate groups and thus reduces the attraction toward Sc3+. This gives the possibility to the N donors for a strengthened interaction with Sc3+.

Natural Energy Decomposition Analysis (NEDA)

Briefly, NEDA57,58 is an energy partitioning procedure for molecular interactions applicable for self-consistent field (SCF) wave functions and DFT charge densities. The total interaction energy (ΔEint) between the appropriately selected fragments consists of electrical interaction (EL), charge transfer (CT), and core repulsion (CORE) contributions:

graphic file with name ic3c02211_m001.jpg

The electrical term, EL = ES + POL + SE, consists of classical electrostatic (ES) and polarization interactions (POL + SE), where SE is the linear response self-energy (energy penalty) of polarization. CT is evaluated as the difference between the energies of the total and localized charge densities. The CORE contribution, CORE = XC + DEF – SE, results from intermolecular exchange–correlation interactions (XC) and deformation (DEF), the latter being the energy difference of the perturbed and relaxed monomer densities.

Metal–DOTA interactions have previously been analyzed in several studies.8,24,54,5963 The main results from the present NEDA study on the isolated complexes are compiled in Table 2. The reported efficient bonding of the DOTA ligand to Sc3+ is reflected in the high interaction energy of the Sc(DOTA) complex in the present set. The other complexes in the present set with 4– charged free ligands have either the same or lower computed interaction energies.

Table 2. Selected Results from the NEDA Analysisa of the Isolated Sc Complexes.

complexb qcc ΔEINT ΔEEL ΔECT ΔECORE EELd
DOTATSAP 1– –7093 –6947 –2033 1887 77.4
DOTASAP 1– –7109 –6969 –2070 1929 77.1
DOTASAP + H2Oe 1– –6440 –6496 –1934 1990 77.1
DOTA-p-NCS-Bn 1– –6936 –6699 –2060 1823 76.5
PSMA′ 0 –6159 –6041 –2092 1974 74.3
DOTATATE′ 0 –6141 –6017 –2089 1966 74.2
Lpyd 3+ –3364 –3085 –1753 1474 63.8
MeDO2PA 1+ –5207 –5088 –2081 1963 71.0
DOTPA 1– –7080 –6974 –1931 1825 78.3
DOTMP 1– –6839 –6705 –2014 1881 76.9
DOTMP″ 5– –10481 –10452 –1675 1646 86.2
AAZTA 1– –7084 –6986 –2163 2066 76.4
macropaLDL 1+ –5213 –4993 –1950 1729 71.9
bispa2 1+ –5211 –5099 –2026 1914 71.6
L3 1+ –5154 –5120 –2106 2072 70.9
L2 2+ –4238 –4165 –1969 1896 67.9
EGTA 1– –7149 –7011 –2247 2108 75.7
BAPTA 1– –6952 –7050 –2051 2149 77.5
DTPA 2– –8075 –7872 –2110 1907 78.9
CHX-A″-DTPA 2– –8044 –7844 –2120 1920 78.7
HOPO 1– –6933 –6628 –2438 2134 73.1
a

Energy data (kJ/mol) according to ΔEINT = ΔEEL + ΔECT + ΔECORE. ΔEINT means the total interaction energy between the two fragments Sc3+ and ligand consisting of electrical interaction (EL), charge transfer (CT), and core repulsion (CORE) contributions.

b

The Sc3+ complexes are denoted by the ligand names given in the first column. The PSMA′ and DOTATATE′ notations mean the truncated ligands to 73 and 77 atoms, respectively; DOTMP and DOTMP″ mean the DOTMP4– and DOTMP8– ligands, respectively.

c

Charges of the isolated complex molecules.

d

EEL = 100·ΔEEL/(ΔEEL + ΔECT).

e

NEDA results of the two-fragment model Sc(H2O)3+ + DOTA4–.

The trends in the total interaction energies of the complexes (ΔEINT), the electrical contributions (ΔEEL), and formal charges of the free ligands are graphically demonstrated in Figure 6. The total interaction energies between the Sc3+ and the ligand fragments vary between −3000 and −11 000 kJ/mol. The trend is primarily determined by the electrical term (EL) due to approximate cancellation of the attractive charge transfer (CT) and repulsive CORE contributions having similar values but opposite signs. The EL contribution is obviously strongly related to the charges of the free ligands: it is the largest with the highly negative DOTMP8– and the smallest with the neutral Lpyd. This latter significant computed value (−3400 kJ/mol), in spite of the neutral net charge of Lpyd, originates mainly from the interaction of Sc3+ with the negatively polarized N atoms. The roughly linear relation of ΔEINT and free ligand charges is well reflected by the gradual increase of ΔEint from neutral Lpyd < MeDO2PA2–, mac2– < DOTA4– derivatives < DTPA5– < DOTMP8– (cf. Figure 6). On the other hand, the markedly different ligand charges do not considerably influence the CT and CORE contributions, as they vary only within 200 kJ/mol (cf. Table 2).

Figure 6.

Figure 6

Total interaction energies (ΔEINT) and electrical contributions (ΔEEL) of the complexes from the NEDA analysis compared with the formal charges (qL) of the free ligands. For the notation of ligands, see footnote a of Table 1.

The ratio of the attractive EL and CT interactions is also strongly related to the charges of the free ligands. The dominant (around 70%) EL interactions have a larger ratio in the complexes with highly negative ligands than in those with less negative ones at the cost of CT (cf. Table 2).

Figure 6 provides also a graphical overview on the minor differences of the various derivatives. Attachment of H2O, like any additional coordination at the ninth coordination site, decreases slightly the interaction energy of the DOTA complex. In both the two-fragment (Sc(H2O)3+ + DOTA4–) and three-fragment (Sc3+ + DOTA4– + H2O) models, ΔEINT is smaller than the related Sc3+ + DOTA4– interaction energy in the Sc(DOTA) complex. This result is in line with the endothermic dissociation to Sc(DOTA) + H2O from the present B3LYP/6-31G** calculations including a correction for BSSE as well as with literature X-ray diffraction results on K[Sc(DOTA)](H6DOTA)Cl2·4H2O lacking any ninth-coordinated ligand.20

The interaction of Sc3+ with DOTA is weakened also by the substituents p-NCS-Bn, PSMA′, and DOTATATE′. From these latter substituents, the p-NCS-Bn one is the most favorable for the complex. It is attached to a cyclen carbon atom and thus leaves all the O and N donors in operation. The slight effect may mainly be attributed to steric interactions. In contrast, the PSMA′ and DOTATATE′ fragments substitute one carboxyl group of DOTA by an amide moiety: accordingly, the strong interaction of Sc3+ with the anionic carboxyl O is replaced by a significantly weaker one with a polarized C=O bond of the amide group. As a consequence, ΔEINT is decreased by ca. 14%.

The interaction energy of the DOTA complex is preserved upon increasing the pendant arms by CH2 groups in Sc(DOTPA). The lengthening/weakening of the Sc–N bonds noted in the previous section seems to be efficiently compensated by the shortening/strengthening of the Sc–O bonds. In contrast, the slightly longer Sc–N bonds (with preserved Sc–O bond distances) result in a slight decrease of ΔEINT for the Sc(DOTMP) complex.

Essentially the same ΔEINT as that with DOTA was computed for the heptadentate AAZTA complex, supporting the experimental observation16 and confirming that octacoordination is not a strict requirement for strong metal–ligand interaction in the case of Sc3+.

From the acyclic ligands, a ΔEINT comparable to that of Sc(DOTA) is achieved with EGTA, whereas slightly smaller ones are achieved with the BAPTA and HOPO ligands. Thus, the shorter carboxylate O–Sc distances upon benzyl substitution in Sc(BAPTA) are not enough to increase ΔEINT with respect to Sc(EGTA) as they insufficiently compensate the considerable weakening of the Sc–N and ether O–Sc interactions (see these longer bonds in Figure 5) in the former complex.

There seems to be no gain for the Sc–ligand interaction with introducing a cyclohexyl group in the DTPA5– ligand either. The decreased flexibility of the ligand in CHX-A″-DTPA5– decreased the scattering of the Sc–O and Sc–N distances (cf. Figure 5) and decreased slightly ΔEINT with respect to that of Sc(DTPA)2–.

The complexes with free ligand charges of 2– (MeDO2PA, macropa, bispa2, L3) have very close ΔEINT’s. The marginally smaller value of L3 further supports the potential of acyclic heptadentate ligands for complexing Sc3+.

Charge Transfer Interactions

Figure 7 compares the energy contribution of CT interactions from the NEDA analysis with the total transferred amount of electrons to Sc3+ and natural populations of the acceptor Sc 4s and Sc 3d orbitals from NBO analysis. There is a reasonably good correlation between the CT energies and transferred electrons. As is known for transition metals, the main electron acceptors are the d orbitals, hence 3d for Sc3+, which thus determines the trend for the totally transferred electrons. The population of the 4s orbitals is ca. 25% of those of 3d with small nonsystematic variations.

Figure 7.

Figure 7

Compilation of the energy contributions of CT interactions in the complexes (ΔECT) from the NEDA analyses, the total transferred amount of electrons to Sc3+ (∑CT), and natural populations of the acceptor Sc 4s and Sc 3d orbitals from NBO analysis. For the notation of ligands, see footnote a of Table 1.

The CT interactions were analyzed in more detail using the second-order perturbation theory approach in the frame of the NBO model. This approach can give the energy contributions of CT between each relevant natural bond orbital and thus facilitates a separation of different types of donors. The contributions of these various donor types in the Sc complexes are summarized in Figure 8.

Figure 8.

Figure 8

Energetics of CT interactions in the complexes from second-order perturbation theory analysis of the Fock matrix in NBO basis: variation of the total CT energy (ΔECT) and % contributions from the four type of donors: anionic O (O), neutral O, aliphatic N (Ncyc), and aromatic N (Nar). The data of Sc(DOTA)(H2O) refer to the three-fragment model: Sc3+ + DOTA4– + H2O. For the notation of ligands, see footnote a of Table 1.

The curve of the total CT energy from second-order perturbation theory analysis agrees qualitatively with the one from the NEDA analysis (cf. Figures 6 and 7). There are only minor quantitative differences due to the different model theories.

Figure 8 demonstrates the CT predominance of carboxylic O donors in the studied Sc complexes. The lower (%) energetic contribution in the complexes with MeDO2PA, macropa, bispa2, L2, and L3 is in agreement with the small number (one or two) of COO groups in these ligands. Yet, their cumulative contribution is in most cases higher than that of the (in larger number available) N donors. Interestingly, in the HOPO complex the formally neutral C=O and the formally anionic N–O donors of the hydroxypyridinone group have nearly identical CT energies. This can be explained by the considerable delocalization of the negative charge in the (O=C—N—O) moiety (cf. Figure 3) leading to a highly polarized C=O group while at the same time preserving its carbonyl character with a bond distance of 1.26 Å. In contrast, the formally neutral C=O (PSMA′, DOTATATE′) and ring O donors in the other ligands (macropa, EGTA, BAPTA) have low-energy CT contributions.

According to the expectations, the H2O ligand in Sc(DOTA)(H2O) has a weak (9.6%) CT to Sc3+. Yet, this interaction is considerably stronger than a single Sc–N one (3.95%, summed up to 15.8% of the four N’s). The data in Figure 8 confirm the weakening of both the Sc–O and Sc–N interactions upon H2O coordination with respect to the Sc(DOTA) parent complex.

CT from N donors is generally low (around 20%) except when the ligand contains aromatic N in pyridazine or picoline groups (Lpyd, macropa, bispa2, and derivatives). In the Sc(DOTMP)5– complex the above-mentioned extraordinary move of Sc3+ toward the PO32– groups results in marginal N → Sc CT interaction (altogether 6.2% by the four N’s). The similarly weak (6.7%) N → Sc charge donation in the Sc(BAPTA) complex in spite of the reasonable Sc–N proximity is mainly due to the delocalization of the N lone pairs with the benzene rings in this ligand.

Conclusions

In the present DFT study the bonding interactions of the pharmaceutically important Sc3+ ion with 18 hepta- to decadentate ligands in isolated complex molecules have been assessed. The theoretical analysis was based on natural energy decomposition analysis and second-order perturbation energies of the Fock matrix from the NBO model. The latter results facilitated a differentiation between the various O and N donors for the CT interactions.

The most stable complexes in terms of interaction energies were formed with octa- and heptadentate ligands. With appropriate donors no significant difference was found between the two coordination forms. In contrast, decadentate coordination (with macropa) had no advantage in complex formation. The steric effects originating from the small size of Sc3+ and the weak coordination efficiency of ether O donors appeared in several longer Sc–donor distances in this complex.

The computed total interaction energies are differentiated in magnitude of a few thousand kilojoules per mole by the charges of the free ligands from 0 to 8–. In contrast, the covalent charge transfer interactions vary within a few hundred kilojoules per mole only. In agreement with the expectations, in all the complexes (even with the neutral Lpyd ligand) the electrical term has the major contribution amounting to 64–86% of the total interaction energy.

More detailed information was obtained on the CT interactions, where both the number of electrons transferred to the acceptor Sc atomic orbitals (3d, 4s) and the energetic contribution of the various O and N donors could be assessed. ΔECT follows qualitatively the trend in the amount of electrons transferred to Sc3+, in particular to the 3d orbitals. The 4s orbitals received ca. 20% of the donated electrons and show marginal variations in the complexes.

The energetic analysis of the CT interactions revealed the overwhelming preference of the carboxylate O donors. From N’s, those incorporated in aromatic rings are more efficient donors in spite of their smaller (less negative) natural charges compared to the cyclen N’s.

Computational Details

The computations were carried out with the Gaussian 09 suite of programs64 using the B3LYP exchange–correlation functional65,66 in conjunction with the 6-31G** basis set.6772 The B3LYP functional was extended with the D3 version of Grimme’s dispersion correction applying the original D3 damping function.73 The SuperFine grid, containing 175 radial shells and 974 angular points (175,974) per shell for H, C, O and N and 250,974 for P, S, and Sc, was applied for integration accuracy. For visualization and construction of some initial geometries, GaussView 5 software74 was applied.

The initial structures for geometry optimizations were taken from the following literature data: complexes with DOTA, Lpyd, MeDO2PA, DOTPA, and DOTMP from computed structures of analogous Bi complexes;23 from computed structures of Lu(macropa)+,52 In(bispa2)+,75 In(L3)+ and In(L2)2+;76 from the crystal structures of NaGd(AAZTA),77 H2Bi(DTPA) and H2Bi(CHX-A″-DTPA),78 KSc(HOPO);46 manually generated from the crystal structures of Cd complexes with methoxy-tetrakisquinoline analogues of EGTA and BAPTA.79 The applied fragments of the large PSMA and DOTATATE ligands were generated manually starting from the computed Sc(DOTA) geometries. In addition to geometry optimizations of the isolated complexes, optimizations in aqueous solution using the polarizable continuum model (PCM)55,56 were also performed.

The natural atomic charges, valence orbital populations, and second-order perturbation energies were evaluated on the basis of the natural bond orbital model17 on the isolated complexes. The metal–ligand interactions were further explored with natural energy decomposition analysis57,58 using the NBO 5.9 code80 incorporated in Firefly software.81 This software is partially based on the GAMESS (US)82 source code. The choice of the B3LYP/6-31G** level of theory for this study was primarily determined by the limited availability of exchange–correlation functionals in the Firefly software. A comparison of the computed and available experimental solid-state20,22 Sc–O and Sc–N bond distances of Sc(DOTA), Sc(DTPA), and Sc(HOPO) complexes is provided in Table S4 together with a brief discussion in the Supporting Information.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.3c02211.

  • Average Sc–N and Sc–O distances in isolated and solvated structures of selected complexes; tables with data presented in Figures 4, 6, and 7 and comparing experimental and computed Sc–O and Sc–N bond distances; Cartesian coordinates of relevant optimized structures (PDF)

Author Contributions

All the calculations, analyses, and manuscript preparation have been carried out by A.K.

The calculations have been carried out using resources provided by the affiliation of the author (JRC Karlsruhe). No external funding was received.

The author declares no competing financial interest.

Supplementary Material

ic3c02211_si_001.pdf (294KB, pdf)

References

  1. NuDat 3 Database. National Nuclear Data Center, Brookhaven National Laboratory, Brookhaven, NY, USA, 2023. https://www.nndc.bnl.gov/nudat3/chartNuc.jsp.
  2. Eppard E.; de la Fuente A.; Benešová M.; Khawar A.; Bundschuh R. A.; Gärtner F. C.; Kreppel B.; Kopka K.; Essler M.; Rösch F. Clinical Translation and First In-Human Use of [44Sc]Sc-PSMA-617 for PET Imaging of Metastasized Castrate-Resistant Prostate Cancer. Theranostics 2017, 7, 4359–4369. 10.7150/thno.20586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Müller C.; Bunka M.; Haller S.; Köster U.; Groehn V.; Bernhardt P.; van der Meulen N.; Türler A.; Schibli R. Promising Prospects for 44Sc-/47Sc-Based Theragnostics: Application of 47Sc for Radionuclide Tumor Therapy in Mice. J. Nucl. Med. 2014, 55, 1658–1664. 10.2967/jnumed.114.141614. [DOI] [PubMed] [Google Scholar]
  4. Velikyan I. Molecular Imaging and Radiotherapy: Theranostics for Personalized Patient Management. Theranostics 2012, 2, 424–426. 10.7150/thno.4428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Baum R. P.; Kulkarni H. R. THERANOSTICS: From Molecular Imaging Using Ga-68 Labeled Tracers and PET/CT to Personalized Radionuclide Therapy - The Bad Berka Experience. Theranostics 2012, 2, 437–447. 10.7150/thno.3645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Kuo P.; Hesterman J.; Rahbar K.; Kendi A. T.; Wei X. X.; Fang B.; Adra N.; Armstrong A. J.; Garje R.; Michalski J. M.; Ghebremariam S.; Brackman M.; Wong C.; Benson T.; Vogelzang N. J. [68Ga]Ga-PSMA-11 PET baseline imaging as a prognostic tool for clinical outcomes to [177Lu]Lu-PSMA-617 in patients with mCRPC: A VISION substudy. J. Clin. Oncology 2022, 40, 5002–5002. 10.1200/JCO.2022.40.16_suppl.5002. [DOI] [Google Scholar]
  7. Dadgar H.; Jafari E.; Ahmadzadehfar H.; Rekabpour S. J.; Ravanbod M. R.; Kalantarhormozi M.; Nabipour I.; Assadi M. Feasibility and therapeutic potential of the 68Ga/177Lu-DOTATATE theranostic pair in patients with metastatic medullary thyroid carcinoma. Ann. Endocrinol. 2023, 84, 45–51. 10.1016/j.ando.2022.08.001. [DOI] [PubMed] [Google Scholar]
  8. Viola-Villegas N.; Doyle R. P. The coordination chemistry of 1,4,7,10-tetraazacyclododecane-N,N’,N″,N’″-tetraacetic acid (H4DOTA): Structural overview and analyses on structure-stability relationships. Coord. Chem. Rev. 2009, 253, 1906–1925. 10.1016/j.ccr.2009.03.013. [DOI] [Google Scholar]
  9. Reubi J. C.; Schär J.-C.; Waser B.; Wenger S.; Heppeler A.; Schmitt J. S.; Mäcke H. R. Affinity profiles for human somatostatin receptor subtypes SST1–SST5 of somatostatin radiotracers selected for scintigraphic and radiotherapeutic use. Eur. J. Nucl. Med. 2000, 27, 273–282. 10.1007/s002590050034. [DOI] [PubMed] [Google Scholar]
  10. Domnanich K. A.; Eichler R.; Müller C.; Jordi S.; Yakusheva V.; Braccini S.; Behe M.; Schibli R.; Türler A.; van der Meulen N. P. Production and separation of 43Sc for radiopharmaceutical purposes. EJNMMI Radiopharm. Chem. 2017, 2, 14. 10.1186/s41181-017-0033-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Huclier-Markai S.; Kerdjoudj R.; Alliot C.; Bonraisin A. C.; Michel N.; Haddad F.; Barbet J. Optimization of reaction conditions for the radiolabeling of DOTA and DOTA-peptide with 44m/44Sc and experimental evidence of the feasibility of an in vivo PET generator. Nucl. Med. Biol. 2014, 41, e36–e43. 10.1016/j.nucmedbio.2013.11.004. [DOI] [PubMed] [Google Scholar]
  12. Kostelnik T. I.; Orvig C. Radioactive Main Group and Rare Earth Metals for Imaging and Therapy. Chem. Rev. 2019, 119, 902–956. 10.1021/acs.chemrev.8b00294. [DOI] [PubMed] [Google Scholar]
  13. Cacheris W. P.; Nickle S. K.; Sherry A. D. 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. 10.1021/ic00253a038. [DOI] [Google Scholar]
  14. Majkowska-Pilip A.; Bilewicz A. Macrocyclic complexes of scandium radionuclides as precursors for diagnostic and therapeutic radiopharmaceuticals. J. Inorg. Biochem. 2011, 105, 313–320. 10.1016/j.jinorgbio.2010.11.003. [DOI] [PubMed] [Google Scholar]
  15. Domnanich K. A.; Müller C.; Farkas R.; Schmid R. M.; Ponsard B.; Schibli R.; Türler A.; van der Meulen N. P. 44Sc for labeling of DOTA- and NODAGA-functionalized peptides: preclinical in vitro and in vivo investigations. EJNMMI Radiopharm. Chem. 2017, 1, 8. 10.1186/s41181-016-0013-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Nagy G.; Szikra D.; Trencsényi G.; Fekete A.; Garai I.; Giani A. M.; Negri R.; Masciocchi N.; Maiocchi A.; Uggeri F.; Tóth I.; Aime S.; Giovenzana G. B.; Baranyai Z. AAZTA: An Ideal Chelating Agent for the Development of 44Sc PET Imaging Agents. Angew. Chem., Int. Ed. 2017, 56, 2118–2122. 10.1002/anie.201611207. [DOI] [PubMed] [Google Scholar]
  17. Reed A. E.; Curtiss L. A.; Weinhold F. Intermolecular interactions from a natural bond orbital, donor-acceptor viewpoint. Chem. Rev. 1988, 88, 899–926. 10.1021/cr00088a005. [DOI] [Google Scholar]
  18. Glendening E. D.; Landis C. R.; Weinhold F. NBO 6.0: Natural Bond Orbital Analysis Program. J. Comput. Chem. 2013, 34, 1429–1437. 10.1002/jcc.23266. [DOI] [PubMed] [Google Scholar]
  19. Weinhold F.; Landis C. R.; Glendening E. D. What is NBO analysis and how is it useful?. Int. Rev. Phys. Chem. 2016, 35, 399–440. 10.1080/0144235X.2016.1192262. [DOI] [Google Scholar]
  20. 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. 10.1002/chem.201402041. [DOI] [PubMed] [Google Scholar]
  21. Aime S.; Barge A.; Botta M.; Fasano M.; Ayala J. D.; Bombieri G. Crystal structure and solution dynamics of the lutetium(III) chelate of DOTA. Inorg. Chim. Acta 1996, 246, 423–429. 10.1016/0020-1693(96)05130-4. [DOI] [Google Scholar]
  22. Benetollo F.; Bombieri G.; Calabi L.; Aime S.; Botta M. Structural Variations Across the Lanthanide Series of Macrocyclic DOTA Complexes: Insights into the Design of Contrast Agents for Magnetic Resonance Imaging. Inorg. Chem. 2003, 42, 148–157. 10.1021/ic025790n. [DOI] [PubMed] [Google Scholar]
  23. Kovács A.; Varga Z. Metal–ligand interactions in complexes of cyclen-based ligands with Bi and Ac. Struct. Chem. 2021, 32, 1719–1731. 10.1007/s11224-021-01816-9. [DOI] [Google Scholar]
  24. Wilson J. J.; Ferrier M.; Radchenko V.; Maassen J. R.; Engle J. W.; Batista E. R.; Martin R. L.; Nortier F. M.; Fassbender M. E.; John K. D.; Birnbaum E. R. Evaluation of nitrogen-rich macrocyclic ligands for the chelation of therapeutic bismuth radioisotopes. Nucl. Med. Biol. 2015, 42, 428–438. 10.1016/j.nucmedbio.2014.12.007. [DOI] [PubMed] [Google Scholar]
  25. Lima L. M. P.; Beyler M.; Oukhatar F.; Le Saec P.; Faivre-Chauvet A.; Platas-Iglesias C.; Delgado R.; Tripier R. H2Me-do2pa: an attractive chelator with fast, stable and inert natBi3+ and 213Bi3+ complexation for potential α-radioimmunotherapy applications. Chem. Commun. 2014, 50, 12371–12374. 10.1039/C4CC05529B. [DOI] [PubMed] [Google Scholar]
  26. Lima L. M. P.; Beyler M.; Delgado R.; Platas-Iglesias C.; Tripier R. Investigating the Complexation of the Pb2+/Bi3+ Pair with Dipicolinate Cyclen Ligands. Inorg. Chem. 2015, 54, 7045–7057. 10.1021/acs.inorgchem.5b01079. [DOI] [PubMed] [Google Scholar]
  27. Das T.; Chakraborty S.; Sarma H. D.; Banerjee S. 177Lu-DOTMP: A viable agent for palliative radiotherapy of painful bone metastasis. Radiochim. Acta 2008, 96, 55–61. 10.1524/ract.2008.1464. [DOI] [PubMed] [Google Scholar]
  28. Simón J.; Frank R. K.; Crump D. K.; Erwin W. D.; Ueno N. T.; Wendt R. E. A preclinical investigation of the saturation and dosimetry of 153Sm-DOTMP as a bone-seeking radiopharmaceutical. Nucl. Med. Biol. 2012, 39, 770–776. 10.1016/j.nucmedbio.2011.12.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Yousefnia H.; Jalilian A. R.; Zolghadri S.; Mirzaei A.; Bahrami-Samani A.; Mirzaii M.; Ghannadi M. Development of 111In DOTMP for dosimetry of bone pain palliation agents. J. Radioanal. Nucl. Chem. 2015, 304, 911–916. 10.1007/s10967-014-3911-6. [DOI] [Google Scholar]
  30. Roca-Sabio A.; Mato-Iglesias M.; Esteban-Gómez D.; Tóth E.; 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. 10.1021/ja808534w. [DOI] [PubMed] [Google Scholar]
  31. Thiele N. A.; Brown V.; Kelly J. M.; Amor-Coarasa A.; Jermilova U.; MacMillan S. N.; Nikolopoulou A.; Ponnala S.; Ramogida C. F.; Robertson A. K. H.; Rodríguez-Rodríguez C.; Schaffer P.; Williams C. Jr; Babich J. W.; Radchenko V.; Wilson J. J. An Eighteen-Membered Macrocyclic Ligand for Actinium-225 Targeted Alpha Therapy. Angew. Chem., Int. Ed. 2017, 56, 14712–14717. 10.1002/anie.201709532. [DOI] [PubMed] [Google Scholar]
  32. Thiele N. A.; MacMillan S. N.; Wilson J. J. Rapid Dissolution of BaSO4 by Macropa, an 18-Membered Macrocycle with High Affinity for Ba2+. J. Am. Chem. Soc. 2018, 140, 17071–17078. 10.1021/jacs.8b08704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Comba P.; Jermilova U.; Orvig C.; Patrick B. O.; Ramogida C. F.; Rück K.; Schneider C.; Starke M. Octadentate Picolinic Acid-Based Bispidine Ligand for Radiometal Ions. Chem.—Eur. J. 2017, 23, 15945–15956. 10.1002/chem.201702284. [DOI] [PubMed] [Google Scholar]
  34. Comba P.; Kerscher M.; Rück K.; Starke M. Bispidines for radiopharmaceuticals. Dalton Trans. 2018, 47, 9202–9220. 10.1039/C8DT01108G. [DOI] [PubMed] [Google Scholar]
  35. Singh G.; Zarschler K.; Hunoldt S.; Martínez I. I. S.; Ruehl C. L.; Matterna M.; Bergmann R.; Máthé D.; Hegedüs N.; Bachmann M.; Comba P.; Stephan H. Versatile Bispidine-Based Bifunctional Chelators for 64CuII-Labelling of Biomolecules. Chem.—Eur. J. 2020, 26, 1989–2001. 10.1002/chem.201904654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Choudhary N.; Dimmling A.; Wang X.; Southcott L.; Radchenko V.; Patrick B. O.; Comba P.; Orvig C. Octadentate Oxine-Armed Bispidine Ligand for Radiopharmaceutical Chemistry. Inorg. Chem. 2019, 58, 8685–8693. 10.1021/acs.inorgchem.9b01016. [DOI] [PubMed] [Google Scholar]
  37. Ndiaye D.; Sy M.; Pallier A.; Même S.; de Silva I.; Lacerda S.; Nonat A. M.; Charbonnière L. J.; Tóth E.; et al. Unprecedented Kinetic Inertness for a Mn2+-Bispidine Chelate: A Novel Structural Entry for Mn2+-Based Imaging Agents. Angew. Chem., Int. Ed. 2020, 59, 11958–11963. 10.1002/anie.202003685. [DOI] [PubMed] [Google Scholar]
  38. Comba P.; Schiek W. Fit and misfit between ligands and metal ions. Coord. Chem. Rev. 2003, 238–239, 21–29. 10.1016/S0010-8545(02)00294-1. [DOI] [Google Scholar]
  39. Nonat A. M.; Roux A.; Sy M.; Charbonnière L. J. 2,4-Substituted bispidines as rigid hosts for versatile applications: from κ-opioid receptor to metal coordination. Dalton Trans. 2019, 48, 16476–16492. 10.1039/C9DT03480C. [DOI] [PubMed] [Google Scholar]
  40. Bruchertseifer F.; Comba P.; Martin B.; Morgenstern A.; Notni J.; Starke M.; Wadepohl H. First-Generation Bispidine Chelators for 213BiIII Radiopharmaceutical Applications. ChemMedChem. 2020, 15, 1591–1600. 10.1002/cmdc.202000361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Połosak M.; Piotrowska A.; Krajewski S.; Bilewicz A. Stability of 47Sc-complexes with acyclic polyamino-polycarboxylate ligands. J. Radioanal. Nucl. Chem. 2013, 295, 1867–1872. 10.1007/s10967-012-2188-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Ricci A. J.; Wu Y. C.; Fettiplace R. The Endogenous Calcium Buffer and the Time Course of Transducer Adaptation in Auditory Hair Cells. J. Neurosci. 1998, 18, 8261. 10.1523/JNEUROSCI.18-20-08261.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Nayak T. K.; Regino C. A. S.; Wong K. J.; Milenic D. E.; Garmestani K.; Baidoo K. E.; Szajek L. P.; Brechbiel M. W. PET imaging of HER1-expressing xenografts in mice with 86Y-CHX-A″-DTPA-cetuximab. Eur. J. Nucl. Med. Mol. Imaging 2010, 37, 1368–1376. 10.1007/s00259-009-1370-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Daumann L. J.; Tatum D. S.; Snyder B. E. R.; Ni C.; Law G.-I.; Solomon E. I.; Raymond K. N. New Insights into Structure and Luminescence of EuIII and SmIII Complexes of the 3,4,3-LI(1,2-HOPO) Ligand. J. Am. Chem. Soc. 2015, 137, 2816–2819. 10.1021/ja5116524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Deri M. A.; Ponnala S.; Kozlowski P.; Burton-Pye B. P.; Cicek H. T.; Hu C.; Lewis J. S.; Francesconi L. C. p-SCN-Bn-HOPO: A Superior Bifunctional Chelator for 89Zr ImmunoPET. Bioconjugate Chem. 2015, 26, 2579–2591. 10.1021/acs.bioconjchem.5b00572. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Phipps M. D.; Cingoranelli S.; Bhupathiraju N. V. S. D. K.; Younes A.; Cao M.; Sanders V. A.; Neary M. C.; Devany M. H.; Cutler C. S.; Lopez G. E.; Saini S.; Parker C. C.; Fernandez S. R.; Lewis J. S.; Lapi S. E.; Francesconi L. C.; Deri M. A. Sc-HOPO: A Potential Construct for Use in Radioscandium-Based Radiopharmaceuticals. Inorg. Chem. 2023, 10.1021/acs.inorgchem.2c03931. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Carter K. P.; Deblonde G. J. P.; Lohrey T. D.; Bailey T. A.; An D. D.; Shield K. M.; Lukens W. W.; Abergel R. J. Developing scandium and yttrium coordination chemistry to advance theranostic radiopharmaceuticals. Commun. Chem. 2020, 3, 61. 10.1038/s42004-020-0307-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Shannon R. D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Crystallogr. 1976, A32, 751–767. 10.1107/S0567739476001551. [DOI] [Google Scholar]
  49. Aime S.; Botta M.; Fasano M.; Marques M. P.; Geraldes C. F.; Pubanz D.; Merbach A. E. Conformational and Coordination Equilibria on DOTA Complexes of Lanthanide Metal Ions in Aqueous Solution Studied by 1H-NMR Spectroscopy. Inorg. Chem. 1997, 36, 2059–2068. 10.1021/ic961364o. [DOI] [PubMed] [Google Scholar]
  50. Cosentino U.; Villa A.; Pitea D.; Moro G.; Barone V.; Maiocchi A. Conformational Characterization of Lanthanide(III)-DOTA Complexes by ab Initio Investigation in Vacuo and in Aqueous Solution. J. Am. Chem. Soc. 2002, 124, 4901–4909. 10.1021/ja017666t. [DOI] [PubMed] [Google Scholar]
  51. Peters J. A.; Djanashvili K.; Geraldes C. F. G. C.; Platas-Iglesias C. The chemical consequences of the gradual decrease of the ionic radius along the Ln-series. Coord. Chem. Rev. 2020, 406, 213146. 10.1016/j.ccr.2019.213146. [DOI] [Google Scholar]
  52. Kovács A. Theoretical Study of Actinide Complexes with Macropa. ACS Omega 2020, 5, 26431–26440. 10.1021/acsomega.0c02873. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Vaira M. D.; Stoppioni P. Theoretical investigation on the geometries of DOTA and DOTA-like complexes and on the transition states of their conformational equilibria. New J. Chem. 2002, 26, 136–144. 10.1039/b106168m. [DOI] [Google Scholar]
  54. Frimpong E.; Skelton A. A.; Honarparvar B. DFT study of the interaction between DOTA chelator and competitive alkali metal ions. J. Mol. Graph. Modelling 2017, 76, 70–76. 10.1016/j.jmgm.2017.06.025. [DOI] [PubMed] [Google Scholar]
  55. Tomasi J.; Mennucci B.; Cammi R. Quantum mechanical continuum solvation models. Chem. Rev. 2005, 105, 2999–3093. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
  56. Scalmani G.; Frisch M. J. Continuous surface charge polarizable continuum models of solvation. I. General formalism. J. Chem. Phys. 2010, 132, 114110. 10.1063/1.3359469. [DOI] [PubMed] [Google Scholar]
  57. Glendening E. D.; Streitwieser A. Natural energy decomposition analysis: An energy partitioning procedure for molecular interactions with application to weak hydrogen bonding, strong ionic, and moderate donor–acceptor interactions. J. Chem. Phys. 1994, 100, 2900–2909. 10.1063/1.466432. [DOI] [Google Scholar]
  58. Glendening E. D. Natural Energy Decomposition Analysis: Extension to Density Functional Methods and Analysis of Cooperative Effects in Water Clusters. J. Phys. Chem. A 2005, 109, 11936–11940. 10.1021/jp058209s. [DOI] [PubMed] [Google Scholar]
  59. Stasiuk G. J.; Long N. J. 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. 10.1039/c3cc38507h. [DOI] [PubMed] [Google Scholar]
  60. Pujales-Paradela R.; Rodríguez-Rodríguez A.; Gayoso-Padula A.; Brandariz I.; Valencia L.; Esteban-Gómez D.; Platas-Iglesias C. On the consequences of the stereochemical activity of the Bi(III) 6s2 lone pair in cyclen-based complexes. The [Bi(DO3A)] case. Dalton Trans. 2018, 47, 13830–13842. 10.1039/C8DT02602E. [DOI] [PubMed] [Google Scholar]
  61. Stein B. W.; Morgenstern A.; Batista E. R.; Birnbaum E. R.; Bone S. E.; Cary S. K.; Ferrier M. G.; John K. D.; Pacheco J. L.; Kozimor S. A.; Mocko V.; Scott B. L.; Yang P. Advancing Chelation Chemistry for Actinium and Other +3 f-Elements, Am, Cm, and La. J. Am. Chem. Soc. 2019, 141, 19404–19414. 10.1021/jacs.9b10354. [DOI] [PubMed] [Google Scholar]
  62. Morgenstern A.; Lilley L. M.; Stein B. W.; Kozimor S. A.; Batista E. R.; Yang P. Computer-Assisted Design of Macrocyclic Chelators for Actinium-225 Radiotherapeutics. Inorg. Chem. 2021, 60, 623–632. 10.1021/acs.inorgchem.0c02432. [DOI] [PubMed] [Google Scholar]
  63. Kovács A. Theoretical Study of Actinide(III)-DOTA Complexes. ACS Omega 2021, 6, 13321–13330. 10.1021/acsomega.1c01292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; Nakatsuji H.; Caricato M.; Li X.; Hratchian H. P.; Izmaylov A. F.; Bloino J.; Zheng G.; Sonnenberg J. L.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E.; Kudin K. N.; Staroverov V. N.; Keith T.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Rega N.; Millam J. M.; Klene M.; Knox J. E.; Cross J. B.; Bakken V.; Adamo C.; Jaramillo J.; Gomperts R.; Stratmann R. E.; Yazyev O.; Austin A. J.; Cammi R.; Pomelli C.; Ochterski J. W.; Martin R. L.; Morokuma K.; Zakrzewski V. G.; Voth G. A.; Salvador P.; Dannenberg J. J.; Dapprich S.; Daniels A. D.; Farkas O.; Foresman J. B.; Ortiz J. V.; Cioslowski J.; Fox D. J.. Gaussian 09, rev. D.01; Gaussian, Inc.: Wallingford, CT, 2010. https://gaussian.com/. [Google Scholar]
  65. Becke A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. 10.1063/1.464913. [DOI] [Google Scholar]
  66. Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
  67. Ditchfield R.; Hehre W. J.; Pople J. A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724–728. 10.1063/1.1674902. [DOI] [Google Scholar]
  68. Francl M. M.; Pietro W. J.; Hehre W. J.; Binkley J. S.; Gordon M. S.; DeFrees D. J.; Pople J. A. Self-consistent molecular orbital methods. XXIII. A polarization-type basis set for second-row elements. J. Chem. Phys. 1982, 77, 3654–3665. 10.1063/1.444267. [DOI] [Google Scholar]
  69. Gordon M. S.; Binkley J. S.; Pople J. A.; Pietro W. J.; Hehre W. J. Self-consistent molecular-orbital methods. 22. Small split-valence basis sets for second-row elements. J. Am. Chem. Soc. 1982, 104, 2797–2803. 10.1021/ja00374a017. [DOI] [Google Scholar]
  70. Hariharan P. C.; Pople J. A. The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 1973, 28, 213–222. 10.1007/BF00533485. [DOI] [Google Scholar]
  71. Hehre W. J.; Ditchfield R.; Pople J. A. 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. 10.1063/1.1677527. [DOI] [Google Scholar]
  72. Rassolov V. A.; Pople J. A.; Ratner M. A.; Windus T. L. 6-31G* basis set for atoms K through Zn. J. Chem. Phys. 1998, 109, 1223–1229. 10.1063/1.476673. [DOI] [Google Scholar]
  73. Grimme S.; Antony J.; Ehrlich S.; Krieg H. A consistent and accurate ab initio parameterization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
  74. Dennington R.; Keith T.; Millam J.. GaussView, ver. 5; Semichem Inc.: Shawnee Mission, KS, 2009. https://gaussian.com/. [Google Scholar]
  75. Kovács A. Metal-Ligand Bonding in Bispidine Chelate Complexes for Radiopharmaceutical Applications. Struct. Chem. 2023, 34, 5–15. 10.1007/s11224-022-01902-6. [DOI] [Google Scholar]
  76. Kovács A. Theoretical study of heptadentate bispidine ligands for radiopharmaceutic applications. Comput. Theor. Chem. 2022, 1212, 113716. 10.1016/j.comptc.2022.113716. [DOI] [Google Scholar]
  77. Aime S.; Bombieri G.; Cavallotti C.; Giovenzana G. B.; Imperio D.; Marchini N. An unusual gadolinium ten-coordinated dimeric complex in the series of MRI contrast agents: Na[Gd(H2O) AAZTA]·3H2O. Inorg. Chim. Acta 2008, 361, 1534–1541. 10.1016/j.ica.2007.09.029. [DOI] [Google Scholar]
  78. Brechbiel M. W.; Gansow O. A.; Pippin C. G.; Rogers R. D.; Planalp R. P. Preparation of the Novel Chelating Agent N-(2-Aminoethyl)-trans-1,2-diaminocyclohexane- N,N’,N″-pentaacetic Acid (H5CyDTPA), a Preorganized Analogue of Diethylenetriaminepentaacetic Acid (H5DTPA), and the Structures of BiIII (CyDTPA)2– and BiIII (H2DTPA) Complexes. Inorg. Chem. 1996, 35, 6343–6348. 10.1021/ic951326p. [DOI] [Google Scholar]
  79. Mikata Y.; Kaneda M.; Konno H.; Matsumoto A.; Sato S.; Kawamura M.; Iwatsuki S. Methoxy-substituted tetrakisquinoline analogs of EGTA and BAPTA for fluorescence detection of Cd2+. Dalton Trans. 2019, 48, 3840–3852. 10.1039/C8DT04735A. [DOI] [PubMed] [Google Scholar]
  80. Glendening E. D.; Badenhoop J. K.; Reed A. E.; Carpenter J. E.; Bohmann J. A.; Morales C. M.; Weinhold F.. NBO 5.9; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, US, 2011. https://nbo6.chem.wisc.edu/index.htm. [Google Scholar]
  81. Granovsky A. A.Firefly, ver. 8; http://classic.chem.msu.su/gran/firefly/index.html.
  82. Schmidt M. W.; Baldridge K. K.; Boatz J. A.; Elbert S. T.; Gordon M. S.; Jensen J. H.; Koseki S.; Matsunaga N.; Nguyen K. A.; Su S.; Windus T. L.; Dupuis M.; Montgomery Jr J. A. General atomic and molecular electronic structure system. J. Comput. Chem. 1993, 14, 1347–1363. 10.1002/jcc.540141112. [DOI] [Google Scholar]

Associated Data

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

Supplementary Materials

ic3c02211_si_001.pdf (294KB, pdf)

Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES