Abstract
Both scandium-44 and yttrium-86 are popular PET isotopes with appropriate half-lives for immuno-positron emission tomography (immuno-PET) imaging. Herein, a new bifunctional H4pypa ligand, H4pypa-phenyl-NCS, is synthesized, conjugated to a monoclonal antibody, TRC105, and labeled with both radionuclides to investigate the long-term in vivo stability of each complex. While the 44Sc-labeled radiotracer exhibited promising pharmacokinetics and stability in 4T1-xenograft mice (n=3) even upon prolonged interactions with blood serum proteins, the progressive bone uptake of the 86Y-counterpart indicated in vivo demetallation, obviating H4pypa as a suitable chelator for Y3+ ion in vivo. The solution chemistry of [natY(pypa)]− was studied in detail and the complex found to be thermodynamically stable in solution with a pM value 22.0, ≥3 units higher than those of the analogous DOTA- and CHX-A”-DTPA-complexes; the 86Y-result in vivo was therefore most unexpected. To explore further this in vivo lability, Density Functional Theory (DFT) calculation was performed to predict the geometry of [Y(pypa)]− and the results were compared with those for the analogous Sc- and Lu-complexes; all three adopted the same coordination geometry (i.e. distorted capped square antiprism), but the metal-ligand bonds were much longer in [Y(pypa)]− than in [Lu(pypa)]− and [Sc(pypa)]−, which could indicate that the size of the binding cavity is too small for the Y3+ ion, but suitable for both the Lu3+ and Sc3+ ions. Considered along with results from [86Y][Y(pypa-phenyl-TRC105)], it is noted that when matching chelators with radionuclides, chemical data such as the thermodynamic stability and in vitro inertness, albeit useful and necessary, do not always translate to in vivo inertness, especially with the prolonged blood circulation of the radiotracer bound to a monoclonal antibody. Although H4pypa is a nonadentate chelator, which theoretically matches the coordination number of the Y3+ ion, we show herein that its binding cavity, in fact, favors smaller metal ions such as Sc3+ and Lu3+ and further exploitation of the Sc-pypa combination is desired.
Graphical Abstract

H4pypa was conjugated to an antibody via a newly synthesized H4pypa-phenyl-NCS; promising immuno-PET imaging with 44Sc was demonstrated
Introduction
Medical imaging is a vital part of drug discovery in oncology. Advancing at a rapid pace, Single-Photon Computerized Tomography (SPECT) and Positron-Emission Tomography (PET) provide information complementary to the anatomical images generated by X-ray Computed Tomography (CT), thereby allowing non-invasive imaging of the tumor microenvironment with excellent sensitivity.1 Both rely on the detection of the photons emitted (SPECT, 100–250 keV) or indirectly generated (PET, 511 keV) during the radioactive decay,2,3 but the latter generally has higher sensitivity and spatial resolution (< 2–3 mm vs. 6–8 mm).4,5,6
Positron-emitting radionuclides are either non-metallic or metallic; the former are incorporated into the carrier via a covalent linkage,7 while the latter are coordinated to a bifunctional multidentate chelating ligand, or chelator, conjugated to a targeting molecule (i.e. antibody, antibody fragment, peptide, etc).2,3,8 The combination of a radionuclide and a bio-vector requires an appropriate match of the radioactive and biological half-lives. The monoclonal antibody is a promising targeting vector with high binding selectivity and recognition of the cellular target, and thus, it is ideal for immuno-PET imaging when combined with an appropriately long-lived positron-emitting radionuclides.9,10 Two attractive PET isotopes are scandium-44 (t1/2 = 3.97 h) and yttrium-86 (t1/2 = 14.7 h).11,12 Scandium-44 has a very high positron-branching ratio (Iβ+ = 94%, Eβ+avg = 632 keV)11,13 and can be produced by either a titanium-44/scandium-44 generator or cyclotron irradiation of a calcium-44 target (44Ca(p,n)44Sc).14,15 The latter allows production in large quantity to address clinical needs. Furthermore, both preclinical and clinical studies have proposed scandium-44 as a better imaging surrogate for lutetium-177 (t1/2 ~ 6.64 d).13,16,17 Although lutetium-177 emits an imageable γ-ray, using it in low-dose dosimetry study can be unreliable as high therapeutic dose can follow different pharmacokinetics.18 Most importantly, the therapeutic isotopologue, scandium-47 (t1/2 = 3.35 d, Eβ−avg=162 keV), possesses highly favorable decay characteristics for therapy.19 These two constitute a chemically equivalent theranostic pair that permits direct translation of the imaging data to the radiation doses of the therapeutic version.20,21,22 As for yttrium-86, the high energy β+-particles and γ-rays are the major concerns since it impacts the image resolution and complicates the logistics.23 Also, the emission of a plethora of γ-rays increases the gamma-coincidences, and subsequently the quantitative errors, which are further amplified in tissues with high attenuation, such as bone, posing difficulties in detection of bone metastases.20,24 Nonetheless, yttrium-86 is still useful in pre-therapy “scout” imaging for the extensively used therapeutic yttrium-90 (t1/2 = 2.67 d, β−).8,25,26
Free scandium-44 and yttrium-86 are rapidly taken up in bone and liver; therefore, a carefully tailored coordination chemistry approach for a chelator is required for specific delivery of the radiation dose to the tumor site.27,28 An appropriate ligand should not only bind the radiometal ion with great thermodynamic stability, but also should result from a synthetically friendly scheme with high functional versatility because the chemical properties of the linker can significantly alter the properties of the whole radiotracer, and consequently the biodistribution. When a temperature- and pH-sensitive monoclonal antibody is involved, the radiolabeling conditions are limited to ~RT and pH = 6–7.29 The tetracarboxylato-macrocyclic DOTA (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid, Chart 1) is incompatible with these conditions, albeit dominating the chelation for a myriad of metal ions, including those of scandium-44 and yttrium-86.3,4,8 In this regard, chelators that combine fast radiolabeling kinetics at room temperature and high stability are desired. The recently reported nonadentate H4pypa demonstrated promising coordination chemistry with scandium-44, with an advantage of RT-radiolabeling.30 As for yttrium-86, although CHX-A”-DTPA (Chart 1) often replaces DOTA in radioimmunoconjugates due to more efficient radiolabeling at ambient temperature,28,31,32 it was shown that the 88Y-labeled p-nitrobenzyl derivatives of CHX-A”-DTPA steadily dissociated in serum over time, while yttrium-88 gradually accumulated in the cortical bone of femurs after the labeled antibody-conjugates was injected into mice.33 Another human serum stability study with 90Y-labeled CHX-A”-DTPA-Trastuzumab also proved transchelation of radioactivity with time.31 Others such as H4octapa, the NETA- and the pyclen-based chelators (Chart 1) have each been investigated to a different extent.34,31,35,36,37,38,39,40 Despite a few promising findings, some of the chelators were derived from rather complicated synthetic routes with limited functional variations, hindering kg production and clinical translation.37 Aside from that, it will be advantageous to employ the same chelator-conjugate for both imaging and therapeutic radionuclides since the resulting radiotracers are more probable to have similar chemical properties.
Chart 1.

Chemical structures of subject ligands.
Considering all these factors, the goal of this report is to explore and evaluate the possibility of employing H4pypa in immuno-PET imaging using scandium-44 and/or yttrium-86. H4pypa is a nonadentate picolinate-carboxylate chelator built around a central pyridine unit that connects the pendant arms, increases the overall complex rigidity, and provides a convenient locus for bifunctionalization. Its nine-coordinating binding cavity has demonstrated high affinity for a range of small- to medium-sized trivalent metal ions (e.g. Sc3+, In3+, Lu3+) at room temperature,30,41 a property that provides versatility in switching treatment modality and complexing both diagnostic and therapeutic radionuclides such as scandium-44 and lutetium-177. Since H4pypa possesses high affinity for Lu3+ at room temperature,41 while Y3+ and Lu3+ ions have rather similar physical properties and chelator selectivity,28 it was anticipated that H4pypa would be a suitable chelating agent for yttrium-86 as well; however, detailed investigations are still required. For radiopharmaceutical applications, complex properties such as the thermodynamic stability, stability in the presence of blood serum proteins, radiolabeling efficiency under mild conditions, as well as the biological stability and properties are important considerations. In order to conjugate H4pypa to an antibody (TRC105, an anti-CD105 monoclonal antibody that inhibits angiogenesis and tumor growth),42,43 a N-isothiocyanate (NCS)-bifunctionalized H4pypa (H4pypa-phenyl-NCS) was developed via a facile synthetic route and labeled with both scandium-44 and yttrium-86 after conjugation. Biodistribution and PET/CT imaging studies were performed to compare their pharmacokinetics. Prior to the animal studies, the basic coordination chemistry of the [Y(pypa)]− complex anion was investigated, including the thermodynamic stability via a series of potentiometric and spectrophotometric titrations. Furthermore, density-functional theory (DFT) calculation predicted the complex geometry, while the radiolabeling efficiency and the stability in mouse serum were demonstrated with yttrium-86.
Results and Discussion
Synthesis and Characterization
A N-hydroxysuccinimide (NHS)-functionalized pypa (tBu4pypa-alkyl-NHS) has previously been reported for peptide conjugation on solid-phase with high coupling efficiency.41 However, because tert-butyl carboxylate groups are supposed to be hydrolyzed after bioconjugation using trifluoroacetic acid (TFA), it is not compatible with the planned antibody conjugation. Fortunately, the para-hydroxyl (p-OH) group in the central pyridine in compound 7 renders H4pypafunctionally versatile. Theoretically, any spacer with a leaving group can be integrated through nucleophilic substitution with the p-OH moiety, engendering a much simpler linker-switching procedure because the bifunctional precursor (compound 7) can be produced in bulk and repeatedly used for attachment of different linkers, helpful for preliminary screening. Herein, with an intention to incorporate a reactive N-isothiocyanate (NCS) group into H4pypa for antibody-coupling, commercially available N-boc-2-(4-aminophenyl)ethanol was selected as the starting material of the spacer (Scheme 1), and tosylated with p-tosyl chloride in tetrahydrofuran (THF) and 6 M sodium hydroxide (NaOH) aqueous solution at room temperature (compound 8, 71%). The p-tosyl leaving group allowed the linker to be coupled to compound 7, reproduced following published protocol.41 The linker was used in 0.1–0.2 equivalent in excess, stirred with compound 7 and anhydrous potassium carbonate (K2CO3) (4 equiv) at room temperature for 24–48 h to ensure complete alkylation. Due to the instability of compound 9 on silica columns, isolation was not performed, and indeed, not required, if all compound 7 was alkylated because the excess linkers would not interfere the following reactions and could be removed during the HPLC purification in the later step. An important key for this reaction is that the chelator should be stirred with the base vigorously for about an hour prior to the addition of the linker. After that, the tert-butyl groups and the tert-butyloxycarbonyl (boc) group in compound 9 were simultaneously removed with TFA in dichloromethane (DCM) (1:1) at room temperature overnight to give H4pypa-phenyl-NH2 (compound 10, 50%), which was isolated using the reverse-phase high-performance liquid chromatography (HPLC) as a single peak at tR = 17.8 min (A: H2O/0.1% TFA B: acetonitrile (ACN)/0.1% TFA, 5–60% B over 40 min, 10 mL/min). The purified aniline compound was finally activated with thiophosgene (CSCl2) in a vigorously stirred biphasic solution of hydrochloric acid (5%, aq)/glacial acetic acid (4:1) and chloroform (CHCl3) at room temperature overnight. The resulting product, H4pypa-phenyl-NCS (compound 11), was isolated with HPLC using the gradient above (tR = 35.0 min), and then lyophilized to a fluffy white solid in 30% yield.
Scheme 1.

Reagents and conditions. i) SOCl2, MeOH, RT-60 °C, 26 h, >99%; ii) BnBr, ACN, K2CO3, 60 °C, overnight, 64%; iii) NaBH4, dry MeOH, RT, overnight, 82%; iv) PBr3, dry CHCl3/dry ACN, 60 °C, overnight, 70%; v) K2CO3, dry ACN, RT, 24 h, 73%; vi) Pd/C (10% w/w), H2 (g), MeOH, RT, overnight; vii) K2CO3, dry THF, RT, 24–48 h, 90%; viii) p-TsCl, THF, 6 M NaOH, 71%; ix) TFA/DCM, overnight, 50%; x) CSCl2, 1 M HCl/ glacial AcOH/CHCl3, RT, 24 h, 30%
Complexation and Characterization
Non-bifunctional H4pypa was synthesized following the published protocol,41 and then complexed with yttrium(III) perchlorate hexahydrate (Y(ClO4)3·6 H2O, 1.1 equiv) in water (pH ~ 7) at room temperature for 1 h to form [Y(pypa)]−. The complex was characterized by different nuclear magnetic resonance (NMR) spectroscopic techniques at ambient temperature (Figures 1 and S19–S22) and high-resolution electrospray-ionization mass spectrometry (HR-ESI-MS) (Figure S23). Characterizations of the [Sc(pypa)]− complex were reported previously and two isomeric species were identified in a temperature-dependent conformational equilibrium.30 As the temperature increased from 25˚C to 85˚C, the equilibrium position further shifted towards the major isomer with sharper 1H NMR peaks.30 This geometric isomerism, however, was not observed with [Y(pypa)]− which formed a single species in neutral aqueous solution without observable fluxionality on the NMR timescale, indicated by the sharp and well-defined 1H NMR signals (Figure S19). Furthermore, the [Y(pypa)]− complex appeared to have 2-fold rotational symmetry about the central pyridine, evidenced by five chemically distinct aromatic 1H signals in a pattern concordant with the symmetric ligand (two triplets: δ 8.23 (Hh) and δ 7.87 (Hf) ppm; three doublets: δ 8.08 (Hg), 7.80 (He), and 7.45 (Hd) ppm). Also, there were no additional 13C NMR peaks compared to the spectrum of H4pypa (Figure S20). The sharp diastereotopic coupling peaks in the aliphatic region confirmed the metal coordination (Figure S19) and the correlations between the diastereotopic 1H were realized by COSY experiments (Figure S22), showing 2J – couplings between Ha and Ha’, Hb and Hb’, as well as between the two peaks merged in Hc. Hf (t, 1H) represented the para-H atom in the central pyridyl ring and underwent 3J – coupling with the ortho-Hd. Similarly, the para-Hh atoms (t, 2H) in two chemically equivalent picolinate arms showed the expected 3J – correlations with both neighboring ortho-Hg and -He atoms. Hb and Hc signals were assigned to the methylene-H atoms on the pyridyl backbone and both picolinate arms, respectively, confirmed by the intense 4J – coupling with the neighboring ortho-H in the pyridines. The most upfield methylene-Ha(a’) signals were assigned to the acetic arms.
Figure 1.

The chemical structure of [Y(pypa)]− with 1H NMR assignments - reference to Figure 2.
Comparing the 1H NMR spectra of [M(pypa)]− M = Sc, Lu and Y (Figure 2), the coordination environment of the Y3+-complex was more similar to that of Lu3+; both exhibited 2-fold rotational symmetry about the central pyridine (Figure 2).41 The major difference between the two was the diastereotopic behavior of the 1H signals at δ ~ 4.66 ppm which was significantly weaker in [Lu(pypa)]− than in [Y(pypa)]−. Nonetheless, both Lu3+- and Y3+-pypa complexes were vastly different from the Sc3+ analogue, which existed in solution as two asymmetric isomers with broader 1H NMR signals at room temperature. Even at 85˚C, the [Sc(pypa)]− complex was not isomerically pure. These observations could be due to the size similarity between the Lu3+ and Y3+ ions, while both are larger than the Sc3+ ion (Sc3+ (0.87 Å) < Lu3+ (0.98 Å) < Y3+ (1.02 Å), 8-coordinated).44 Regarding the isomerism of the [Sc(pypa)]− complex, it is noteworthy that the previous biodistribution studies using the isomeric mixture of the 44Sc-labeled H4pypa-peptidomimetic-conjugate targeting prostate-specific membrane antigen demonstrated high overall in vivo stability (the isomers could not be separated).30 However, since the biological properties are also heavily dependent on the bioconjugation method and the conjugated targeting vector, it is essential to evaluate the pharmacokinetics of each bioconjugate, even if the same radiometal-chelate is used.
Figure 2.

1H NMR spectra (400 MHz, D2O) of H4pypa (RT), [Y(pypa)]− (RT), [Lu(pypa)]− (RT) and [Sc(pypa)]− (RT & 85˚C).
DFT Calculations
The geometry of the [Y(pypa)]− anion (Figure 3A) was calculated by DFT and compared to those of [Sc(pypa)]− and [Lu(pypa)]− (Figure S25). The calculated bond lengths suggested nine-coordination for all three complexes, consistent with the 1H NMR spectra where diastereotopic splitting of the methylene-H atoms were observed (Figure 2). All three pypa-complexes shared the same stable geometry of distorted capped square antiprism (Figure 3B). As reported previously, the smaller Sc3+ ion also complexed with H4pypa in another conformation ~22.4 kJ/mol less stable than the distorted capped square antiprism.30 The small energy difference allowed both [Sc(pypa)]− isomers to co-exist in solution and interconvert in a temperature-dependent manner, as confirmed in the 1H NMR spectra (Figure 2).30 The average bond lengths of the less stable [Sc(pypa)]− geometry were expectedly longer than the more stable one.30 In order to directly compare the structural information of [M(pypa)]− (M = Sc, Lu and Y), the less stable [Sc(pypa)]− conformer was not considered due to its completely different geometry, and only the bond lengths (Table 1) and the X-M-Y bond angles (Table 2) of the shared geometry were concerned. As shown in Table 1, the bond distances between the metal ion and the donor atoms increased in tandem with the ionic radii (i.e. Sc3+ < Lu3+ < Y3+), indicating weaker coordination bonds in the [Y(pypa)]− complex among all,44 while the X-M-Y bond angles were smaller for the bigger metal ions (Table 2). The observations were consistent with the trend reported by Bazargan et al. based on the Cambridge Structural Database.45 The flexibility and the relationship between the bond length and the bond angle allow the binding cavity of the ligand to match itself with different metal ions.45
Figure 3. (A).

DFT calculated geometry for the [Y(pypa)]− anion. (Geometries calculated for the [Sc(pypa)]− and [Lu(pypa)]− anions are in Figure S25). (B) [Y(pypa)]− coordination environment showing only Y3+ and the donor atoms (A perspective looking through the central pyridine).
Table 1.
DFT-calculated metal-donor bond lengths for the [M(pypa)]− (M = Sc, Lu, Y) anions.
| Donor atom | Bond length (Å) | |||
|---|---|---|---|---|
| [Sc(pypa)]− | [Lu(pypa)]− | [Y(pypa)]− | ||
| Backbone pyr-N | N1 | 2.4960 | 2.5150 | 2.5405 |
| Backbone-N | N2 | 2.5661 | 2.6058 | 2.6354 |
| Backbone-N | N3 | 2.5647 | 2.6057 | 2.6353 |
| Picolinate-N | N4 | 2.3495 | 2.4495 | 2.4803 |
| Picolinate-N | N5 | 2.3500 | 2.4501 | 2.4793 |
| Acetate-COO | O1 | 2.2115 | 2.3514 | 2.3723 |
| Acetate-COO | O2 | 2.2149 | 2.3494 | 2.3753 |
| Picolinate-COO | O3 | 2.1872 | 2.3341 | 2.3647 |
| Picolinate-COO | O4 | 2.1925 | 2.3375 | 2.3618 |
Table 2.
DFT-calculated X-M-Y bond angles for the [M(pypa)]− (M = Sc, Lu, Y) anions.
| X | Y | M | ||
|---|---|---|---|---|
| Sc3+ | Lu3+ | Y3+ | ||
| Backbone pyr-N | Backbone-N | 65.89˚ | 65.49˚ | 64.76˚ |
| Backbone pyr-N | Backbone-N | 65.93˚ | 65.36˚ | 64.93˚ |
| Picolinate-N | Backbone-N | 67.74˚ | 66.69˚ | 66.02˚ |
| Picolinate-N | Backbone-N | 67.67˚ | 66.59˚ | 66.19˚ |
| Backbone-N | Acetate-COO | 70.83˚ | 68.99˚ | 68.28˚ |
| Backbone-N | Acetate-COO | 70.75˚ | 69.03˚ | 68.17˚ |
| Picolinate-N | Picolinate-COO | 70.38˚ | 67.55˚ | 66.73˚ |
| Picolinate-N | Picolinate-COO | 70.30˚ | 67.42˚ | 66.82˚ |
Based on that, the ideal M-N bond length is ~2.5 Å and the ideal N-M-N bond angle is ~69˚ in a five-membered NCCNM chelate ring, which is preferred by the larger metal ions (i.e. ionic radius ~1 Å),46,47,48 the coordination bonds between the metal ion and the N atoms of two tertiary amines are slightly more favorable in [Sc(pypa)]− (~2.56 Å) compared to [Y(pypa)]− (~2.64 Å) (Table 1), so as the N-M-N bond angles (Table 2). Despite the differences, all three metal ions demonstrated strong preferences for the oxygen donor atoms (i.e. significantly shorter M-O bonds) due to their hard nature.
Solution Thermodynamics
The stability constant of a metal complex is an important thermodynamic parameter, despite the lack of predictability of the kinetic inertness. Because the metal complexation reaction always occurs in competition with the protonation equilibria of the ligand, knowledge of the protonation constants of the ligand is mandatory to determine the stability constant of its metal complexes. H4pypa was previously determined to possess nine protonation sites, the most basic of which deprotonated with a pKa = 7.78 (protonated tertiary amines in the pyridyl backbone), suggesting that at physiological pH (~7.4), almost all the donor atoms are free for chelation. Complex formation equilibria with more basic chelators, such as DOTA (Chart 1, highest pKa = 12.6049) and CHX-A”-DTPA (Chart 1, 12.3050) will occur with higher competition between the protons and the basic sites on the ligand. In this light, a more comprehensive parameter that allows for a better comparison of the affinity of different chelators for a metal ion is pM. The pM value is widely adopted for this purpose and is defined under physiologically relevant conditions (−log [M]free at [L] = 10 μM, [M] = 1 μM and pH = 7.4). It accounts not only for the stability of the metal complexes, but also for the ligand basicity (pKa values) and denticity, and can provide a more accurate prediction of the in vivo thermodynamic stability and speciation of metal-scavenging at physiological pH = 7.4.51
To investigate the complexation equilibria of the Y3+-H4pypa system, a series of titration studies were conducted. Since the complexation between the Y3+ ion and H4pypa was complete at pH ≤ 2, below the electrode threshold (Figure S26), direct determination of the thermodynamic stability constant of [Y(pypa)]− (log KY(pypa)) was not feasible. Instead, a ligand-ligand competition method with EDTA as competitor (Chart 1) and acidic in-batch UV spectrophotometric titrations (Figure S27) was applied to determine the formation constant of the protonated species ([Y(Hpypa)]) prior to those of [Y(pypa)]− and [Y2(OH)(pypa)2]3- using direct potentiometric titration. Potentiometric and spectrophotometric experimental data were refined using the Hyperquad201352 and HypSpec201453 programs. The thermodynamic stability constant of [Y(pypa)]− (log K[Y(pypa)]−) was finally calculated to be 21.60(1) (0.16 M NaCl, 25˚C) (Table 3). Although it is 2.49 and 3.10 units lower than those of the DOTA- and CHX-A”-DTPA- complexes, respectively,34,50 the pY value (22.0), a more accurate assessment, as described above, was ≥3 units higher.34,50 The difference was partially a result of the drastically lower basicity of H4pypa. When comparing to other M3+-H4pypa systems (M3+ = Sc3+, In3+, Lu3+ and La3+),30,41 pY value was consistent with the inverse correlation between the size of the metal ion and the metal sequestering ability of H4pypa (except the Sc3+-complex).30,41 With similar ionic radii (Y3+ = 1.075 Å vs. Lu3+ = 1.032 Å, 9-coordinated),44 pY is also very close to pLu, implying similar thermodynamic stability in both systems (Figure 4). Another interesting observation is that with larger metal ions such as Y3+ and La3+, H4pypa tends to form a 2:2 metal:ligand-hydroxide species under basic conditions, rather than the 1:1 complex as seen in the systems with the Sc3+, In3+ and Lu3+ ions (Table 3).
Table 3.
Stepwise stability constants (log K) of H4pypa with Sc3+, Lu3+, Y3+ and La3+ ions.
| Equilibrium reaction | Sc3+ | Lu3+ | Y3+ | La3+ |
|---|---|---|---|---|
| M3+ + L ⇆ ML | 26.98(1)30 | 22.20(2)41 | 21.60(1) (a) | 19.74(3)41 |
| ML + H+ ⇆ MHL | 3.55(2)30 | 3.60(6)41 | 2.14(5)(b) 2.29(c) | 3.24(5)41 |
| M(OH)L + H+ ⇆ ML | 10.34(3)30 | 10.77(8)41 | - | - |
| M2L2(OH) + H+ ⇆ M2L2 | - | - | 37.23(a) | 34.40(6)41 |
| pM(d) | 27.130 | 22.641 | 22.0 | 19.941 |
from direct UV-potentiometric titration method at 25 °C, I = 0.16 M NaCl;
from EDTA potentiometric competition titration method at 25 °C, I = 0.16 M NaCl;
from direct UV-acidic competition titration method at 25 °C, I ≠ 0.16 M NaCl;
pM is defined as −log [M]free at [L] = 10 μM, [M] = 1 μM and pH = 7.4.
Figure 4.

pM vs. ionic radius44 for M3+ and ligands of interest.
Radiolabeling of [86Y][Y(pypa)]− and Mouse Serum Challenge Experiments
Because the radiotracer is usually formulated with very low chelator concentrations (<10−5 M), it is important to ascertain the affinity of a chelator for a radiometal ion under dilution. In this regard, one common useful study is the concentration-dependent radiolabeling experiment in which a constant amount of radioactivity is added to a series of diluted chelator solutions and then incubated for a desired period. With an intention to use H4pypa in immuno-PET imaging, the labeling experiment of H4pypa with yttrium-86 was performed under antibody-compatible conditions (pH = 7 and ambient temperature) using 0.2 M HEPES (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid) buffer and ~4.1 MBq of yttrium-86. The mixtures were agitated at room temperature for 7, 15, 30 and 60 min, and then the radiochemical yield percentage (RCY%) of [86Y][Y(pypa)]− was determined with silica aluminum-backed thin layer chromatography (TLC) plates, developed in diethylenetriamine-pentaacetic acid (DTPA) buffer (0.1 M, pH = 5.5). Attempts to identify the complex by the radioactive high-performance liquid chromatography (HPLC) failed as [86Y][Y(pypa)]− decomplexed in the HPLC column, probably due to the fast acid-assisted dissociation kinetics at low concentrations of the metal complex in the acidic mobile phases (ACN/H2O/0.1% TFA solution, pH ~ 2) (Figure S28). Regarding the radio-metalation, the efficiency was concentration-dependent (96±1% at 10−5 M vs. 88±3 at 10−6 M after 15 min) (Figure 5A and Table S1) and that at the lower concentrations (i.e. 10−6 M and 10−7 M) gradually improved with time (83±2%, 7 min vs. 93±1, 60 min at 10−6 M; 15±2%, 7 min vs. 26±2, 60 min at 10−7 M) (Table S1). Furthermore, the apparent molar activity of the [86Y][Y(pypa)]− complex (the radioactivity was corrected to the end-of-bombardment (EoB)) significantly increased with time, from 79 GBq/μmol at 7 min to 105 GBq/μmol at 60 min, meaning that more radioactivity could be used to achieve quantitative RCY if the complexation time is longer. The efficient radiolabeling at room temperature is an advantage of non-macrocyclic/acyclic chelator and was also seen with CHX-A”-DTPA. A radiolabeling study of yttrium-90 and CHX-A”-DTPA-DUPA-Pep (a peptide conjugate) conducted by Benjamin et al. obtained ~98% RCY in 5 min at room temperature (25 μg of peptide, pH = 5.5 and 3.7–3.8 MBq).54 On the other hand, quantitative radiolabeling with DOTATOC in 10 min required heating at 100˚C (pH = 4),55 which is not compatible with the monoclonal antibody, but an expected requirement for a constrained macrocycle. The radiolabeling results for [44Sc][Sc(pypa)]− were previously reported,30 and similar to [86Y][Y(pypa)]−, the radiochemical yields were both concentration- and time-dependent (95±0% at 10−5 M vs. 90±2 at 10−6 M after 5 min; 9±3%, 5 min vs. 21±1, 60 min at 10−7 M, pH = 5.5, 2.9 MBq). In spite of the similar radiolabeling behavior, only [44Sc][Sc(pypa)]− could be isolated in the acidic HPLC system as a single sharp peak at 13.4 min (same stationary and mobile phases as [86Y][Y(pypa)]− were used), implying much slower acid-assisted dissociation kinetics at pH ~ 2 than for the 86Y-analog.30 Nevertheless, the efficient radiolabeling of H4pypa at room temperature with both radionuclides renders it a promising alternative to the “gold standard” chelator, DOTA, in building immuno-constructs.55,56
Figure 5. (A).

Concentration-dependent radiolabeling studies of H4pypa with 44Sc in NH4OAc buffer (0.5 M, pH=5.5) and 86Y in HEPES buffer (0.5 M, pH=7) at room temperature over 15 min. (B) Mouse serum stability of [86Y][Y(pypa)]−over 48 h at 37˚C.
After the successful radiolabeling results, it was deemed prudent to test the stability of the [86Y][Y(pypa)]− complex in mouse serum. In the study, the labeled complex was incubated in mouse serum at 37˚C over 48 h, and the intact percentage of the complex was determined by silica aluminum-backed TLC plates, developed in DTPA buffer (0.1 M, pH = 5.5). To validate the experiment, a control sample containing only yttrium-86 in the labeling buffer and mouse serum was prepared, showing only one radioactivity spot at the solvent front on the developed TLC plate, which confirmed that both the free and the transchelated yttrium-86 were moved by DTPA. The study showed that ~97% of [86Y][Y(pypa)]− remained intact after 2 days which is encouraging (Figure 5B and Table S2).
PET/CT Imaging and Biodistribution Studies
Following injection of 44Sc- or 86Y-labeled H4pypa-phenyl-TRC105 (~18 and 11 MBq/mouse, respectively) into 4T1-xenograft-bearing mice (n = 3), serial PET/CT images (maximum intensity projection, MIP) were acquired at 30 min, 4 h, 8 h and 18 h post-injection (p.i.) for the 44Sc-labeled tracer, as well as 30 min, 4 h, 24 h and 48 h p.i. for the 86Y-counterpart (Figures 6–8). A higher dose of the 44Sc-labeled radiotracer was injected in order to acquire the late-timepoint images. Radioactivity uptakes of selected organs were analyzed with quantitative regions-of-interest (ROI) imaging and plotted in Figure 7 in % injected-dose-per-gram (% ID/g). The ex vivo biodistribution results at the last time point are shown in Figure 8 and Table S3 which confirmed the accuracy of the image-based ROI quantification. Both radioimmunoconjugates showed expected uptakes in the blood, tumor, liver, kidney and spleen due to the antibody circulation, metabolism and excretion. High uptake in lung was seen in other reported TRC105-constructs, perhaps due to the prolonged residence in the blood pool.57,58,59 The tumor was delineated shortly after 4 h post-injection (p.i.) with 10.2±1.6% ID/g and 8.1±0.6% ID/g for 44Sc- and 86Y-constructs, respectively. In the case of 44Sc, the tumor uptake increased by ~6% ID/g from 4 to 18 h p.i. (16.1±1.9% ID/g), but only ~1% ID/g for [86Y][Y(pypa-phenyl-TRC105)] between 4 and 24 h p.i, and then stabilized over the following 24 h (9.1±0.9% ID/g at 48 h p.i.). Additionally, for the 44Sc-tracer, the radioactivity in blood decreased in tandem with that in the liver without significant change in the bone uptake throughout the course of study, indicating the long-term in vivo stability of the [44Sc][Sc(pypa)]− complex. Unfortunately, for the 86Y-counterpart, despite the clearance in the blood pool, the bone uptake jumped by nearly 3-fold from 30 min (2.1±0.3% ID/g) to 48 h p.i. (5.7±0.5% ID/g), almost double compared to that of the [86Y][Y(DTPA-TRC105)] (~3% ID/g at 48 h p.i.).58 The liver activity also increased by ~2.9% ID/g within the first 4 h. Taken together, it suggests that [86Y][Y(pypa-phenyl-TRC105)] is unstable in vivo. The in vivo transchelation of the 86Y-construct could also explain the significantly lower tumor accumulation compared to that of the 44Sc-counterpart. This outcome was not expected as [86Y][Y(pypa)]− appeared to be stable in mouse serum, with high thermodynamic stability comparable to the Lu3+-H4pypa system which was stable in vivo.41 A possible explanation could be that the conjugation changed the coordination environment of the [86Y][Y(pypa)]− complex. Furthermore, since the radiopharmaceutical is usually injected in a very low concentration and is further diluted as it is distributed in the blood volume, the rate of dissociation of the complex, rather than the thermodynamic equilibrium, will dictate the in vivo stability.3,60,61 This is further enhanced when there are orders-of-magnitude more concentrated endogenous ligands (e.g. transferrin) competing with the chelator-conjugate in binding the radiometal ions.3 Therefore, the disappointing pharmacokinetics of [86Y][Y(pypa-phenyl-TRC105)] are likely a result of the insufficient kinetic inertness and metabolic stability of the complex. The result with [44Sc][Sc(pypa-phenyl-TRC105)] was deemed a great success. The tumor was visualized at the highest contrast at 18 h p.i. (Figure 6A). The biodistribution pattern of [44Sc][Sc(pypa-phenyl-TRC105)] closely resembled that of [86Y][Y(DTPA-TRC105)] with uptake in blood, lung, liver, kidney, spleen and tumor, and no abnormal bone and liver retentions which manifest signs of in vivo stability (Figure 7A),58 evidencing the aptness of H4pypa in scavenging 44Sc.
Figure 6.

PET/CT MIP Images of (A) [44Sc][Sc(pypa-phenyl-TRC105)] (B) [86Y][Y(pypa-phenyl-TRC105)] at different post-injection time points.
Figure 8.

Ex vivo biodistribution data of [44Sc][Sc(pypa-phenyl-TRC105)] (18 h p.i.) and [86Y][Y(pypa-phenyl-TRC105)] (48 h p.i.).
Figure 7.

Quantitative ROI analysis of the in vivo PET imaging data of (A) [44Sc][Sc(pypa-phenyl-TRC105)] (B) [86Y][Y(pypa-phenyl-TRC105)] at different post-injection time points.
Conclusion
The hypothesis of this work was to explore H4pypa in immuno-PET imaging with scandium-44 and yttrium-86, which eventually could be useful for theranostic applications with long-lived therapeutic radionuclides such as lutetium-177. H4pypa previously demonstrated favorable radiolabeling properties and complex stability with scandium-44 and lutetium-177.30,41 Due to the similar chemical properties between Lu3+ and Y3+ ions,28 H4pypa was anticipated to hold promise for the Y3+ ion as well. The [Y(pypa)]− complex presented a single symmetric species with the DFT-calculated bond lengths longer than both [Sc(pypa)]− and [Lu(pypa)]− complexes. Although the Y3+-H4pypa system was thermodynamically favorable (pY value of 22.0, more than 3 units higher than those of the DOTA- and CHX-A”-DTPA-complexes34,50), and the [86Y][Y(pypa)]− complex appeared to be stable in mouse serum as well, when [86Y][Y(pypa-phenyl-TRC105)] was injected into a cohort of 4T1-xenograft mice (n = 3), progressive bone accumulation was observed due to the undesirable release of free yttrium-86 in vivo. On the other hand, the pharmacokinetics of [44Sc][Sc(pypa-phenyl-TRC105)] were highly favorable with long-term in vivo stability. Although the previously reported [44Sc][Sc(pypa-C7-PSMA617)] showed promising biological stability as well,30 this finding is important because, firstly, H4pypa-phenyl-NCS was a new bifunctional chelator which could lead to different metal coordination properties. In addition, antibody-conjugates possess much longer blood half-life (1–3 weeks) which prolonged the interactions between the radio-complex and the blood serum proteins, while the PSMA-conjugates tend to be excreted very quickly;62 therefore, the in vivo stability, particularly at late timepoints, would be more accurate when examined with an antibody-construct. The long-term in vivo stability of [44Sc][Sc(pypa)]− encourages future immuno-PET applications using a 44Sc-labeled pypa-phenyl-antibody, especially in the dosimetry study for the long-lived therapeutic radionuclides such as scandium-47 and lutetium-177, which can be conveniently chelated with the same bioconjugate to attain more comparable biodistributions.
Experimental Section
Materials and Methods
All solvents and reagents were purchased from commercial suppliers (TCI America, Alfa Aesar, AK Scientific, Sigma-Aldrich, Fisher Scientific, Fluka) and were used as received. Deionized water was filtered through the PURELAB Ultra Mk2 system. 1H, 13C{1H}, 1H-13C{1H} HSQC and COSY NMR spectra were recorded at ambient temperature on Bruker AV400 instruments, as specified; the NMR spectra are expressed on the δ scale and were referenced to residual solvent peaks. Low-resolution (LR) mass spectrometry was performed using a Waters ZG spectrometer with an ESCI electrospray/chemical-ionization source, and high-resolution electrospray-ionization mass spectrometry (HR-ESI-MS) was performed on a Micromass LCT time-of-flight instrument at the Department of Chemistry, University of British Columbia. Analyses of radiolabeled compounds were performed with both thin layer chromatography (TLC) (i.e. silica aluminum-backed TLC plates and iTLC-impregnated with silica gel (iTLC-SG) strips) purchased from Agilent Technologies. The radio-TLC scanner model was the Cyclone Storage Phosphor System (Cyclone Plus) from Perkin Elmer and the HPLC system was Dionex Ultimate 3000. DIONEX Acclaim C18 5 μm 120 Å column (250 mm × 4.60 mm) was used for separation of free radioactivity and radio-complex. 44Sc and 86Y were provided by the UW-Madison Cyclotron Lab as 0.1 M HCl solutions.
Dimethyl 4-hydroxypyridine-2,6-dicarboxylate (1)
Thionyl chloride (SOCl2) (9.50 mL, 0.130 mol, 5 equiv) was added slowly using a syringe to a stirred suspension of chelidamic acid monohydrate (5.28 g, 26.2 mmol, 1 equiv) in methanol (60 mL) in a two-neck round-bottom flask at 0°C. The mixture was stirred at room temperature for 24 h and then refluxed for an additional 2 h. The solvent was removed under reduced pressure gently at room temperature and then deionized water was added at 0˚C. The mixture was neutralized with aqueous 1 M K2CO3 solution and the precipitate was filtered by vacuum filtration, and then washed with 50% MeOH in water (~10 mL). The white precipitate was dried under reduced pressure to give a white solid (5.54 g, >99%). 1H NMR (400 MHz, 298 K, (CD3)2SO): δ 6.74 (s, 2H), 3.72 (s, 6H). 13C NMR (100 MHz, 298 K, (CD3)2SO): δ 165.7, 149.2, 116.6, 52.7. LR-ESI-MS: calcd. for [C9H9NO5 + Na]+ 234.0; found [M + Na]+ 234.2
Dimethyl 4-(benzyloxy)pyridine-2,6-dicarboxylate (2)
To a round-bottom flask with a stirred solution of compound 1 (1.65 g, 7.82 mmol, 1 equiv) in dry acetonitrile (ACN) was added anhydrous K2CO3 (2.19 g, 15.8 mmol, 2.02 equiv) and benzyl bromide (1.02 mL, 8.60 mmol, 1.1 equiv). The reaction mixture was refluxed overnight at 60˚C. K2CO3 was filtered out by vacuum filtration and then washed with DCM. The filtrate was concentrated in vacuo and then purified through a silica column (CombiFlash Rf automated column system, 24 g gold silica column, A: DCM B: MeOH, 0–5% B). The product fractions were rotary-evaporated to give a white powder (1.51 g, 64 %). 1H NMR (400 MHz, 298 K, CDCl3): δ 7.90 (s, 2H), 7.44–7.38 (m, 5H), 5.23 (s, 2H), 4.01 (s, 6H). 13C NMR (100 MHz, 298 K, CDCl3): δ 150.0, 129.0, 128.9, 127.9, 115.0, 71.0, 53.4. LR-ESI-MS: calcd. for [C16H15NO5 + Na]+ 324.1; found [M + Na]+ 324.1
(4-(Benzyloxy)pyridine-2,6-diyl)dimethanol (3)
To a round-bottom flask with a stirred solution of compound 2 (8.74 g, 29.0 mmol, 1 equiv) in dry MeOH (90 mL) was added sodium borohydride (NaBH4) (3.29 g, 87.1 mmol, 3 equiv) in three portions over 30 min at 0˚C. The reaction mixture was stirred at room temperature. After 24 h, the mixture was diluted with CHCl3 (50 mL) and then quenched with saturated aqueous NaHCO3 solution (50 mL). The organic phase was separated and the bulk of MeOH in the aqueous layer was removed in vacuo to give an aqueous solution which was extracted with CHCl3 (50 mL × 4). The combined organic phases were dried over anhydrous sodium sulfate (Na2SO4), and then clarified by filtration. The filtrate was rotary-evaporated to give a white solid (5.86 g, 82%).1H NMR (400 MHz, 298 K, CDCl3): δ 7.42–7.35 (m, 5H), 6.79 (s, 2H), 5.12 (s, 2H), 4.70 (s, 4H). 13C NMR (100 MHz, 298 K, CDCl3): δ 184.4, 166.5, 162.7, 160.6, 149.6, 135.6, 128.9, 128.6, 127.6, 117.2, 111.8, 107.7, 106.5, 106.1, 105.2, 70.2, 64.5. LR-ESI-MS: calcd. for [C14H15NO3 + Na]+ 268.1; found [M + Na]+ 268.2
4-(Benzyloxy)-2,6-bis(bromomethyl)pyridine (4)
Compound 3 (1.76 g, 12.6 mmol, 1 equiv) was suspended in dry ACN/CHCl3 (40 mL, 50:50 v/v) in a three-neck round-bottom flask. Phosphorus tribromide (PBr3) (3.60 mL, 37.9 mmol, 3 equiv) in CHCl3 (5 mL) was added dropwise using a dropping funnel to the stirred solution of compound 3 at 0˚C over 15 min. The reaction mixture was stirred at 60˚C for 18 h and then saturated aqueous Na2CO3 solution was added slowly to quench the reaction at 0˚C. The aqueous phase was extracted with CHCl3 (50 mL × 3). The combined organic phases were dried over anhydrous Na2SO4, and then clarified by filtration. The filtrate was rotary-evaporated to yield a colorless oil which later solidified to a white solid (3.28 g, 70%). 1H NMR (400 MHz, 298 K, CDCl3): δ 7.43 (m, 5H), 7.36 (s, 2H), 5.37 (s, 2H), 4.95 (s, 4H). 13C NMR (100 MHz, 298 K, CDCl3): δ 170.9, 154.5, 133.2, 129.5, 129.3, 128.3, 113.2, 73.0, 25.3. LR-ESI-MS: calcd. for [C14H1379Br2NO + H]+ 369.9; found [M(79Br) + H]+ 369.9
Tert-butyl 6-(((2-(tert-butoxy)-2-oxoethyl)amino)methyl)picolinate (5).
To a round-bottom flask with a stirred solution of tert-butyl 6-formylpicolinate41 (0.500 g, 2.40 mmol, 1 equiv) in dry MeOH (20 mL) was added tert-butyl glycinate (0.320 g, 2.40 mmol, 1 equiv). The mixture was stirred for 1 h at room temperature and then sodium cyanoborohydride (NaBH3CN) (0.310 g, 4.87 mmol, 2 equiv) was added. The reduction reaction was continued for 3 h at room temperature before quenching with saturated NaHCO3 in water (10 mL) and then extraction with DCM (20 mL × 3). The combined organic phases were dried over anhydrous Na2SO4, and then clarified by filtration. The filtrate was concentrated in vacuo and the residue was purified through a silica column (CombiFlash Rf automated column system, 12 g gold silica column, A: DCM B: MeOH, 0–5% B). The product fractions were combined and rotary-evaporated to give a pale yellow oil (0.550 g, 70%).1H NMR (400 MHz, 298 K, CDCl3): δ 7.76 (d, J = 7.7 Hz, 1H), 7.65 (t, J = 7.7 Hz, 1H), 7.43 (d, J = 7.6 Hz, 1H), 3.91 (s, 2H), 3.26 (s, 2H), 1.48 (s, 9H), 1.32 (s, 9H).13C NMR (100 MHz, 298 K, CDCl3): δ 171.1, 163.9, 159.7, 148.6, 137.2, 124.9, 123.0, 81.9, 81.0, 54.2, 51.0, 27.9. LR-ESI-MS: calcd. for [C17H26N2O4 + H]+ 323.2; found [M + H]+ 323.1
Di-tert-butyl-6,6’-((((4-(benzyloxy)pyridine-2,6-diyl)bis(methylene))bis((2-(tert-butoxy)-2-oxoethyl)azanediyl))bis(methylene))dipicolinate (6)
Compound 4 (0.400 g, 1.30 mmol, 1 equiv), K2CO3 (595 mg, 4.31 mmol, 3.3 equiv) and KI (434 mg, 2.61 mmol, 2 equiv) were added sequentially to a stirred solution of compound 5 (0.837 g, 2.60 mmol, 2 equiv) in dry ACN (15 mL) in a round-bottom flask. The mixture was stirred at 30˚C for 24 h. K2CO3 was removed by centrifugation and then washed with DCM (10 mL × 3). The combined supernatants were concentrated in vacuo and then purified with a silica column (CombiFlash Rf automated column system, 12 g gold silica column, A: DCM B: MeOH, 0–5% B). The product fractions were rotary-evaporated to give a pale-yellow oil (0.670 g, 73 %). 1H NMR (400 MHz, 298 K, CDCl3): δ 7.92–7.61 (m, 6H), 7.52–7.30 (m, 5H), 7.12 (s, 2H), 5.11 (s, 2H), 4.03 (s, 4H), 3.86 (s, 4H), 3.33 (s, 4H), 1.57 (s, 18H), 1.43 (s, 18H). 13C NMR (100 MHz, 298 K, CDCl3): δ 170.5, 166.2, 164.0, 160.2, 148.6, 137.2, 136.1, 128.5, 128.1, 127.7, 125.6, 123.3, 123.0, 107.8, 81.8, 80.9, 69.7, 64.4, 59.8, 56.1, 53.4, 28.0. LR-ESI-MS: calcd. for [C48H63N5O9 + Na]+ 876.5; found [M + Na]+ 876.6
Di-tert-butyl-6,6’-((((4-hydroxypyridine-2,6-diyl)bis(methylene))bis((2-(tert-butoxy)-2-oxoethyl)azanediyl))bis(methylene))dipicolinate (7)
Compound 6 (0.170 g, 0.200 mmol) was dissolved in dry MeOH (20 mL) in a three-neck round-bottom flask and saturated with N2(g). Pd/C (10% w/w, 0.1 equiv) was added under a stream of N2(g). The flask was purged with N2(g), followed by H2(g) from a balloon. The mixture was stirred vigorously at room temperature overnight under H2 atmosphere, and then Pd/C was filtered off through a pre-wet (MeOH) Celite bed, washed with MeOH (10 mL × 5). The filtrate was rotary-evaporated to a pale-yellow oil (0.150 g) and used without further purification. LR-ESI-MS: calcd. for [C41H57N5O9 + H]+ 764.4; found [M + H]+ 764.6
4-((Tert-butoxycarbonyl)amino)phenethyl 4-methylbenzenesulfonate (8)
N-boc-2-(4-aminophenyl)ethanol (1.97 g, 8.28 mmol, 1 equiv) was dissolved in THF (12 mL) and cooled to 0°C with an ice-water bath. NaOH aqueous solution (6 M, 11.9 mL) was added, followed by dropwise addition of para-tosyl chloride (3.16 g, 0.0169 mol, 2 equiv) in THF (24 mL) under N2(g). After stirring at 0˚C for 1 h, the reaction mixture was warmed to room temperature and further stirred overnight. The mixture was extracted with DCM (30 mL × 3). The combined organic phases were washed with 1 M NaOH aqueous solution (40 mL × 2) and deionized water (40 mL × 2), and then dried over magnesium sulfate (MgSO4). The mixture was clarified with filtration, evaporated in vacuo and then purified through a silica column (CombiFlash Rf automated column system, 24 g gold silica column, A: DCM B: MeOH, 0–5% B). The product fractions were rotary-evaporated to give a white solid (2.30 g, 71%). 1H NMR (400 MHz, 298 K, CDCl3): δ 7.68 (d, J = 8.3 Hz, 2H), 7.25 (dd, J = 13.4, 8.7 Hz, 4H), 7.01 (d, J = 8.4 Hz, 2H), 6.45 (s, 1H), 4.16 (t, J = 7.0 Hz, 2H), 2.89 (t, J = 7.0 Hz, 2H), 2.43 (s, 3H), 1.51 (s, 9H).13C NMR (100 MHz, 298 K, CDCl3): δ 152.8, 144.8, 137.3, 133.1, 130.7, 129.9, 129.6, 128.0, 118.8, 80.7, 70.8, 34.8, 28.5, 21.7. LR-ESI-MS: calcd. for [C20H25NO5S + H]+ 392.1; found [M + H]+ 392.1
Tetramethyl 6,6’,6”,6”’-((((4-(4-((tert-butoxycarbonyl)amino)phenethoxy)pyridine-2,6-diyl)bis(methylene))bis(azanetriyl))tetrakis(methylene))tetrapicolinate (9)
To a round-bottom flask with a stirred solution of compound 7 (72.1 mg, 0.0944 mmol, 1 equiv) in dry THF (1 mL) was added anhydrous K2CO3 (52.1 mg, 0.378 mmol, 4 equiv). The mixture was stirred vigorously for 1 h at 25˚C before the addition of compound 8 (42.1 mg, 0.108 mmol, 1.14 equiv). The mixture was stirred for 48 h at 25˚C when compound 7 was completely consumed. The solvent was evaporated in vacuo, and the residue was resuspended in DCM (6 mL). K2CO3 was removed by centrifugation and washed with DCM twice (~5 mL each). The combined organic phases were washed with saturated NaHCO3 in water (10 mL × 2), water (10 mL × 2) and brine (10 mL × 2), and then dried over anhydrous MgSO4. The drying agent was filtered off and the filtrate was concentrated in vacuo to a yellow oil. The product was confirmed by MS and then used without further purification in the next step. LR-ESI-MS: calcd. for [C54H74N6O11 + Na]+ 1005.5; found [M + K]+ 1005.9
6,6’,6”,6”’-((((4-(4-Aminophenethoxy)pyridine-2,6-diyl)bis(methylene))bis(azanetriyl))tetra-kis(methylene))tetrapicolinic acid (10)
Compound 9 (166 mg, 0.169 mmol, 1 equiv) was dissolved in TFA/DCM (1:1) (4 mL). The mixture was stirred overnight vigorously at room temperature, and then concentrated to dryness in vacuo. The crude product was re-dissolved in deionized water, and then purified through reverse phase HPLC (A: ACN/0.1% TFA, B: H2O/0.1% TFA, 5–60% A over 40 min, 10 mL/min, tR = 17.8 min). The combined product fractions were dried in vacuo to give a yellow oil (55.7 mg, 50%).1H NMR (400 MHz, 298 K, D2O): δ 8.23 (t, J = 7.9 Hz, 2H), 7.97 (d, J = 7.8 Hz, 2H), 7.91 (d, J = 7.7 Hz, 2H), 7.51 (d, J = 8.4 Hz, 2H), 7.43 (d, J = 8.5 Hz, 2H), 6.91 (s, 2H), 4.49 (s, 4H), 4.34 (t, J = 5.8 Hz, 2H), 4.28 (s, 4H), 3.82 (s, 4H), 3.17 (t, J = 5.8 Hz, 2H). 13C NMR (100 MHz, 298 K, D2O): δ 13C NMR (101 MHz, D2O) δ 174.1, 169.6, 164.9, 154.1, 152.6, 146.5, 144.3, 139.5, 131.0, 128.7, 128.4, 125.2, 123.3, 110.7, 70.6, 57.8, 57.2, 33.9.. LR-ESI-MS: calcd. for [C33H34N6O9 + H]+ 659.7; found [M + H]+ 659.4.
H4pypa-phenyl-NCS (11).
Compound 10 (68.0 mg, 0.103 mmol, 1 equiv) was dissolved in 1 M HCl/ glacial acetic acid (2 mL, 4:1 v/v) in a round-bottom flask; thiophosgene (CSCl2) (119 μL, 1.55 mmol, 15 equiv) in CHCl3 (2 mL) was then added dropwise using a Pasteur pipette to the stirred mixture which was stirred vigorously at room temperature overnight. After the reaction completed, CHCl3 was removed with a Pasteur pipette. The aqueous phase was further washed with CHCl3 (1 mL) which was removed by a Pasteur pipette. The process was repeated 4 times and then the residue was injected onto reverse phase HPLC (A: ACN/0.1% TFA, B: H2O/0.1% TFA, 5–60% A over 40 min, 10 mL/min, tR = 35 min). The product fractions were combined and lyophilized to give a fluffy white solid (21.7 mg, 30 %). 1H NMR (400 MHz, 298 K, CD3CN:D2O 1:1): δ 8.62 – 8.50 (m, 4H), 8.27 (d, J = 7.3 Hz, 2H), 7.90 (d, J = 8.4 Hz, 2H), 7.84 (d, J = 8.4 Hz, 2H), 7.50 (s, 2H), 4.90 (s, 6H), 4.31 (s, 4H), 3.65 (t, J = 6.3 Hz, 2H). 13C NMR (100 MHz, 298 K, CD3CN:D2O 1:1): δ 173.5, 165.9, 155.0, 142.2, 131.2, 131.2, 128.7, 126.5, 126.5, 125.4, 111.2, 58.7, 56.6, 56.6, 56.2, 56.2. HR-ESI-MS: calcd. for [C34H32N6O9S+H]+ 701.2030; found [M+H]+ 701.2028
Na[natY(pypa)]
H4pypa · 2 TFA· 2 H2O (21.8 mg, 4.16 × 10−5 mol, 1 equiv) was dissolved in D2O (1 mL) in a scintillation vial and 0.1 M NaOD solution (aq) was added to adjust the pH to 7; Y(ClO4)3 · 6 H2O (45.4 μL, 4.58 × 10−5 mol, 1.1 equiv) was then added. The mixture was stirred at room temperature for 1 h and used to confirm the complexation by LR-ESI-MS and NMR. 1H NMR (400 MHz, 298 K, D2O): δ 8.23 (t, J = 7.8 Hz, 2H), 8.08 (d, J = 7.7 Hz, 2H), 7.87 (t, J = 7.8 Hz, 1H), 7.80 (d, J = 7.8 Hz, 2H), 7.45 (d, J = 7.8 Hz, 2H), 4.71 – 4.58 (m, 4H), 4.40 (d, J = 14.7 Hz, 2H), 4.05 (d, J = 14.8 Hz, 2H), 3.97 (m, J = 17.0 Hz, 2H), 3.47 (d, J = 17.0 Hz, 2H).. 13C NMR (100 MHz, 298 K, D2O): δ 178.7, 172.5, 156.7, 152.9, 151.1, 141.8, 140.2, 125.6, 124.1, 123.0, 64.6, 63.2. HR-ESI-MS: calcd. for [C25H21N5O889Y + 2Na]+ 654.0244; found [M + 2Na]+ 654.0251.
DFT Calculation
All DFT simulations were performed as implemented in the Gaussian 09 revision D.01 suite of ab initio quantum chemistry programs (Gaussian Inc., Wallingford, CT). B3LYP functional,63,64 and the effective core potentials LanL2DZ basis sets for scandium65,66 and yttrium63,67,68 were applied to optimize the structural geometry in the presence of water solvent (IEF PCM as implemented in G09) without the use of symmetry constraints. Normal self-consistent field (SCF) and geometry convergence criteria were conducted for all the calculations. The calculated structures were visualized using Mercury 4.1
Solution Thermodynamics
All potentiometric titrations were carried out with a Metrohm Titrando 809 and a Metrohm Dosino 800 with a Ross combined electrode. A 20 mL and 25˚C thermostated glass cell with an inlet-outlet tube for nitrogen gas (purified through a 10% NaOH solution to exclude any CO2 prior to and during the course of the titration) was used as a titration cell. The electrode hydrogen ion concentration was calibrated daily by direct titration of HCl (aq) with freshly prepared NaOH aqueous solution and the results were analyzed with the Gran procedure69 in order to obtain the standard potential E° and the ionic product of water pKw at 25˚C and 0.16 M NaCl (aq) used as a supporting electrolyte. Solutions were titrated with carbonate-free NaOH (aq, 0.16 M) that was standardized against freshly recrystallized potassium hydrogen phthalate aqueous solution. In the study of complex formation equilibria, the determination of the stability constants of the [Y(Hpypa)] species was carried out by two different methods. The first method used UV-vis spectrophotometric measurements on a set of solutions containing 1:1 metal to ligand molar ratio ([H4pypa] = [Y]3+ = 1.33 × 10−4 M) and different amounts of standardized HCl (aq) and NaCl (aq) to set the ionic strength constant at 0.16 M when possible. The equilibrium H+ concentration in this UV in-batch titration procedure at low pH solutions (2 ≥ pH ≤ 0) was calculated from solution stoichiometry, not measured with a glass electrode. For the solutions of high acidity, the correct acidity scale H0 was used.70 The spectral range was 200–400 nm at 25˚C and 1 cm path length. The molar absorptivities of all the protonated species of H4pypa calculated with HypSpec201453 from the protonation constant experiments41 were included in the calculations. The second method used competition pH-potentiometric titrations with EDTA as a ligand competitor and the composition of the solutions was [Y]3+ = Y3+ ~ 1.51 × 10−3 M, [H4pypa] ~ 6.88 × 10−4 M and [EDTA] ~ 1.51 × 10−3 M at 25˚C and I = 0.16 M NaCl (aq). The stability constants for the complexes formed by EDTA and Y3+ were taken from the literature.71 Direct pH-potentiometric titrations of the Y3+-H4pypa systems were also carried out. The Y3+ metal ion solution was prepared by adding the atomic absorption (AA) standard solution to a H4pypa solution of known concentration in the 1:1 metal to ligand molar ratio. Ligand and metal concentrations were in the range of 8.24 × 10−4 M. The exact amount of acid present in the AA standard solution was determined by Gran’s method69 titrating equimolar solutions of Y(III) and Na2H2-EDTA. Each titration consisted of 100–150 equilibrium points in the pH range 1.6–11.5; equilibration time for titrations was up to 5 min for metal complex titrations. Three replicates of each titration were performed. Relying on the stability constants for the species [Y(Hpypa)] obtained by the two different methods, the fitting of the direct potentiometric titrations was possible and yielding the stability constants in Table 3. All the potentiometric measurements were processed using the Hyperquad201352 software while the obtained spectrophotometric data were processed with the HypSpec201453 program. Proton dissociation constants corresponding to hydrolysis of Y(III) aqueous ions included in the calculations were taken from Baes and Mesmer.72 The overall equilibrium (formation) constants log β referred to the overall equilibria: pM + qH + rL ⇆ MpHqLr (the charges are omitted), where p might also be 0 in the case of protonation equilibria and q can be negative for metal-hydroxo species. Stepwise equilibrium constants log K correspond to the difference in log units between the overall constants of sequentially protonated (or hydroxide) species. The parameter used to calculate the metal scavenging ability of a ligand towards a metal ion, pM, is defined as −log [Mn+]free at [ligand] = 10 mM and [Mn+] = 1 μM at pH = 7.4.73
Production and Radiochemical Separation of Yttrium-86
Yttrium-86 (t1/2 = 14.7 h, 34% β+, Eβ+max = 3.2 MeV) was produced in a 16 MeV GE PETtrace biomedical cyclotron using enriched [86Sr][SrCO3] targets of pressed powder. Following irradiation, the radiochemical isolation of the yttrium-86 was performed by single column extraction chromatography, as previously described.74 Briefly, the target material was dissolved in 5 mL of 9 M HCl (aq), and loaded onto a column filled with a resin functionalized with N,N,N′,N′-tetrakis-2-ethylhexyldiglycolamide (branched DGA, Eichrom). Subsequent washes were carried out using 9 M HCl (aq, 15 mL) and 0.5 M HNO3 (aq, 15 mL) to remove bulk strontium and other trace metal contaminants. No-carrier-added yttrium-86 was eluted with 0.1 M HCl (aq, 4 × 0.3 mL). The activity was assayed in a Capintec CRC-15R dose calibrator (setting #850/2).
Production and Radiochemical Separation of Scandium-44
Scandium-44 was cyclotron-produced using natCa[p,n]4xSc nuclear reactions on pressed targets of metallic calcium (300–350 mg). Target preparation was performed in air, and rapidly mounted in the cyclotron to reduce calcium oxidation. Irradiations were performed at 20 μA for 1 h with direct water cooling, and a 12.7 μm Nb foil was used to degrade the beam energy from the nominal 16 MeV to 14.1 MeV. Under these conditions, a scandium-44 production yield of 0.4 mCi/μAh was obtained through the reaction natCa(p,n)44Sc. Isolation of the produced scandium-44 was carried out by single column extraction chromatography using a N,N,N’,N’-tetrakis-2-ethylhexyldiglycolamide functionalized resin (DGA-branched, Eichrom).65 The target was dissolved in 9 M HCl (aq, 10 mL) and passed through a 1 mL fritted solid phase extraction (SPE) tube filled with the DGA resin (~120 mg), loading the scandium-44 and eluting bulk Ca2+. Remaining Ca2+ was removed by rinsing the column with 4 M HCl (aq, 20 mL). Next, a 12 mL wash with 1 M HNO3 (aq) was performed to elute possible trace metal contaminants such as Zn, Fe and Cu. Finally, the scandium-44 was eluted in a small volume using 0.1 M HCl (aq, 4 × 500 μL fractions). The radionuclidic and chemical purities were confirmed by high purity germanium (HPGe) gamma spectrometry and microwave plasma atomic emission spectroscopy (MP-AES), respectively.
Radiolabeling Studies
An aliquot of ligand solution (H4pypa) (10 μL) was mixed with HEPES solution (0.5 M, pH = 7, 87 μL), followed by [86Y][YCl3] in HCl (aq) (4.14 MBq, 3 μL). The reaction mixtures were incubated at ambient temperature over the desired time period; 3 μL was then spotted on a silica aluminum-backed TLC plate, and developed in DTPA buffer (0.1 M, pH = 5.5). The TLC plate was read by a TLC reader, showing the free metal ion migrated to the solvent front while the complex stayed at the baseline. The areas of both peaks were used to calculate RCY%.
In vitro Mouse Serum Challenge
To the radiolabeled sample (100 μL, 1.4 MBq), an equal volume (100 μL) of the mouse serum was added. The mixture was incubated at 37˚C and 5 μL aliquots were collected at desired time points (0.5 h, 4 h, 24 h, 48 h). Each aliquot was spotted onto a silica aluminum-backed TLC plate next to the control spot (yttrium-86 in buffer with serum), and then developed in DTPA buffer (0.1 M, pH = 5.5). The TLC plate was read by the TLC reader. The free metal migrated to the solvent front while the complex stayed at the baseline. The areas of both peaks were used to calculate % intact.
Bioconjugation of H4pypa-phenyl-TRC105 and Radiolabeling with Yttrium-86 and Scandium-44
TRC105 (TRACON Pharmaceuticals) was prepared for radiolabeling through conjugation with H4pypa-phenyl-NCS. The pH of the TRC105 solution was adjusted to ~9.0 with 0.1 M sodium carbonate buffer and H4pypa-phenyl-NCS chelator was added in a 10:1 (chelator:antibody) molar ratio. After reacting for 2 h at room temperature, H4pypa-phenyl-TRC105 was purified by size exclusion chromatography (PD-10, GE-Healthcare) using phosphate-buffered saline (PBS) as the mobile phase. For animal studies, 2.7 μg of H4pypa-phenyl-TRC105 per MBq of yttrium-86 (or scandium-44) was added to the activity solution in sodium acetate buffer (0.1 M, pH = 5.5) and incubated at 37˚C for 30 min. After labeling, [AE][E(pypa-phenyl-TRC105)] (AE = 86Y, 44Sc) was purified by size exclusion chromatography using PBS as the mobile phase. Radiochemical yields (RCY) were measured via iTLC-SG plates developed in DTPA buffer (0.1 M, pH = 5.5) for the 86Y-conjugate and sodium citrate buffer (0.4 M, pH = 4.5) for the 44Sc-conjugate. The radiochemical yields were consistently above 90% for both radiotracers.
PET/CT Imaging and Biodistribution Studies
Animal experiments were conducted with the approval of the University of Wisconsin Institutional Animal Care and Use Committee (IACUC) and in strict accordance with the NIH guidelines for the care and use of laboratory animals (NIH Publication No. 85–23 Rev. 1985) and the relevant guidelines and regulations (protocol ID: B00000657). 4T1 cells were cultured in RPMI 1640 growth medium (Invitrogen) with a 10% fetal bovine serum (FBS) supplement. During culturing, cells were incubated at 37˚C with 5% CO2. Tumors were grafted in four-week-old female Balb/c mice by subcutaneous injection of 2–3 × 106 cells, suspended in 100 μL of 1:1 mixture of RPMI 1640 and Matrigel (BD Biosciences).
For imaging studies, mice bearing 4T1 murine breast cancer tumors were injected via the tail vein with 11–18 MBq of either [86Y][Y(pypa-phenyl-TRC105)] or [44Sc][Sc(pypa-phenyl-TRC105)]. PET scans of 20 million coincidence events per mouse were obtained using an Inveon PET/CT scanner (Siemens) at different post-injection timepoints. Measurements of regions-of-interest (ROI) in PET images were performed using the Inveon Research Workspace (Siemens).
Following the final imaging timepoint, mice were euthanized via CO2 asphyxiation, and major organs were excised, wet-weighed, and radioactive content was measured using a gamma counter (PerkinElmer). Results of PET ROI analysis and biodistribution studies are presented as %ID/g.
Supplementary Material
Acknowledgements
We thank the team of animal technicians and veterinarians for helpful assistance in the University of Wisconsin. We gratefully acknowledge funding from NSERC CREATE IsoSiM at TRIUMF for PhD research stipend (LL), both NSERC and CIHR for financial support via a Collaborative Health Research Project (CHRP to CO and PS), NSERC Discovery (CO, VR, PS) and financial support from the National Cancer Institute of the National Institutes of Health under Award Number T32CA009206 (Eduardo Aluicio-Sarduy). TRIUMF receives federal funding via a contribution agreement with the National Research Council of Canada.
Footnotes
The authors declare no competing financial interest.
Supporting Information
NMR spectra of compounds 1 - 11 and [Y(pypa)]−; high-resolution mass spectra of [Y(pypa)]−and H4pypa-phenyl-NCS; DFT calculated geometry for the [Sc(pypa)]− and [Lu(pypa)]− anions; distribution diagram of the Y3+-H4pypa system; representative spectra of the in-batch UV-titration of the Y3+-pypa system as the pH is raised; radiochemical-yield data, intact % of the complex upon mouse serum challenge and radio-HPLC chromatographs of [86Y][Y(pypa)]− complex and free yttrium-86; ex vivo biodistribution studies data (%ID/g) of both [86Y][Y(pypa-phenyl-TRC105)] at 48 h p.i. and [44Sc][Sc(pypa-phenyl-TRC105)] at 18 h p.i.
References
- 1.Rudin M and Weissleder R, Nat. Rev. Drug Discovery, 2003, 2, 123–131. [DOI] [PubMed] [Google Scholar]
- 2.Ramogida CF and Orvig C, Chem. Commun, 2013, 49, 4720–4739. [DOI] [PubMed] [Google Scholar]
- 3.Price EW and Orvig C, Chem. Soc. Rev, 2014, 43, 260–290. [DOI] [PubMed] [Google Scholar]
- 4.Zeglis BM and Lewis JS, Dalton Trans, 2011, 40, 6168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Rahmim A and Zaidi H, Nucl. Med. Commun, 2008, 29, 193–207. [DOI] [PubMed] [Google Scholar]
- 6.van Dongen GAMS, Poot AJ and Vugts DJ, Tumor Biol, 2012, 33, 607–615.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Hao G, Singh AN and Liu W and Sun X, Curr. Top. Med. Chem, 2010, 10, 1096–1112. [DOI] [PubMed] [Google Scholar]
- 8.Kostelnik TI and Orvig C, Chem. Rev, 2019, 119, 902–956. [DOI] [PubMed] [Google Scholar]
- 9.Wu AM, J. Nucl. Med, 2009, 50, 2–5. [DOI] [PubMed] [Google Scholar]
- 10.Verel I, Visser GWM and van Dongen GA, J. Nucl. Med, 2005, 46, 164S–171S. [PubMed] [Google Scholar]
- 11.Hernandez R, Valdovinos HF, Yang Y, Chakravarty R, Hong H, Barnhart TE and Cai W, Mol. Pharmaceutics, 2014, 11, 2954–2961. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Banerjee SR, Foss CA, Pullambhatla M, Wang Y, Srinivasan S, Hobbs RF, Baidoo KE, Brechbiel MW, Nimmagadda S, Mease RC, Sgouros G and Pomper MG, J. Nucl. Med, 2015, 56, 628–634. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Umbricht CA, Benešová M, Schmid RM, Türler A, Schibli R, van der Meulen NP and Müller C, EJNMMI Res., 2017, 7, 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Severin GW, Engle JW, Valdovinos HF, Barnhart TE and Nickles RJ, Appl. Radiat. Isot, 2012, 70, 1526–1530. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Filosofov DV, Loktionova NS and Rösch F, Radiochim. Acta, 2010, 98, 149–156. [Google Scholar]
- 16.Müller C, Bunka M, Reber J, Fischer C, Zhernosekov K, Türler A and Schibli R, J. Nucl. Med, 2013, 54, 2168–2174. [DOI] [PubMed] [Google Scholar]
- 17.Singh A, van der Meulen NP, Müller C, Klette I, Kulkarni HR, Türler A, Schibli R and Baum RP, Cancer Biother. Radiopharm, 2017, 32, 124–132. [DOI] [PubMed] [Google Scholar]
- 18.Pfestroff A, Luster M, Jilg CA, Olbert PJ, Ohlmann CH, Lassmann M, Maecke HR, Ezziddin S and Bodei L, Eur. J. Nucl. Med. Mol. Imaging, 2015, 42, 1971–1975. [DOI] [PubMed] [Google Scholar]
- 19.Müller C, Bunka M, Haller S, Köster U, Groehn V, Bernhardt P, van der Meulen N, Türler A and Schibli R, J. Nucl. Med, 2014, 55, 1658–1664. [DOI] [PubMed] [Google Scholar]
- 20.Rösch F, Herzog H and Qaim S, Pharmaceuticals, 2017, 10, 56. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Srivastava SC, Semin. Nucl. Med, 2012, 42, 151–163. [DOI] [PubMed] [Google Scholar]
- 22.Herzog H, Rösch F, Stöcklin G, Lueders C, Qaim SM and Feinendegen LE, J. Nucl. Med, 1993, 34, 2222–2226. [PubMed] [Google Scholar]
- 23.Aluicio-Sarduy E, Ellison PA, Barnhart TE, Cai W, Nickles RJ and Engle JW, J. Labelled Compd. Radiopharm, 2018, 61, 636–651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Conti M and Eriksson L, EJNMMI Phys, 2016, 3, 8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Bodei L, Cremonesi M, Grana C, Rocca P, Bartolomei M, Chinol M and Paganelli G, Eur. J. Nucl. Med. Mol. Imaging, 2004, 31, 1038–1046. [DOI] [PubMed] [Google Scholar]
- 26.Helisch A, Fórster GJ, Reber H, Buchholz HG, Arnold R, Göke B, Weber MM, Wiedenmann B, Pauwels S, Haus U, Bouterfa H and Bartenstein P, Eur. J. Nucl. Med. Mol. Imaging, 2004, 31, 1386–1392. [DOI] [PubMed] [Google Scholar]
- 27.Brechbiel MW, Nucl QJ. Med. Mol. Imaging, 2008, 52, 166–173. [PMC free article] [PubMed] [Google Scholar]
- 28.Khozeimeh Sarbisheh E and Price EW, in Radiopharmaceutical Chemistry, Springer International Publishing, Cham, 2019, pp. 359–370. [Google Scholar]
- 29.Vermeer AW and Norde W, Biophys. J, 2000, 78, 394–404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Li L, de G. Jaraquemada-Peláez M, Aluicio-Sarduy E, Wang X, Jiang D, Sakheie M, Kuo H-T, Barnhart TE, Cai W, Radchenko V, Schaffer P, Lin K-S, Engle JW, Bénard F and Orvig C, Inorg. Chem, 2020, 59, 1985–1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Price EW, Edwards KJ, Carnazza KE, Carlin SD, Zeglis BM, Adam MJ, Orvig C and Lewis JS, Nucl. Med. Biol, 2016, 43, 566–576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Kobayashi H, Wu C, Yoo TM, Sun BF, Drumm D, Pastan I, Paik CH, Gansow OA, Carrasquillo JA and Brechbiel MW, J. Nucl. Med, 1998, 39, 829–836. [PubMed] [Google Scholar]
- 33.Camera L, Kinuya S, Garmestani K, Wu C, Brechbiel MW, Pai LH, McMurry TJ, Gansow OA, Pastan I, Paik CH and Carrasquillo JA, J. Nucl. Med, 1994, 35, 882–889. [PubMed] [Google Scholar]
- 34.Le Fur M, Beyler M, Molnár E, Fougère O, Esteban-Gómez D, Tircsó G, Platas-Iglesias C, Lepareur N, Rousseaux O and Tripier R, Chem. Commun, 2017, 53, 9534–9537. [DOI] [PubMed] [Google Scholar]
- 35.Price EW, Cawthray JF, Adam MJ and Orvig C, Dalton Trans, 2014, 43, 7176–7190. [DOI] [PubMed] [Google Scholar]
- 36.Kang CS, Chen Y, Lee H, Liu D, Sun X, Kweon J, Lewis MR, and Chong H-S, Nucl. Med. Biol, 2015, 42, 242–249. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Chong H-S, Song HA, Kang CS, Le T, Sun X, Dadwal M, Lee H, Lan X, Chen Y and Dai A, Chem. Commun, 2011, 47, 5584–5586. [DOI] [PubMed] [Google Scholar]
- 38.Kang CS, Sun X, Jia F, Song HA, Chen Y, Lewis M and Chong H-S, Bioconjugate Chem, 2012, 23, 1775–1782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Le Fur M, Beyler M, Molnár EM, Fougè O, Esteban-Gómezgómez D, Tircsó G, Platas-Iglesias C, Lepareur N, Rousseaux O and Tripier R, Inorg. Chem 2018, 57, 2051–2063. [DOI] [PubMed] [Google Scholar]
- 40.Le Fur M, Beyler M, Lepareur N, Fougère O, Platas-Iglesias C, Rousseaux O and Tripier R, Inorg. Chem, 2016, 55, 8003–8012. [DOI] [PubMed] [Google Scholar]
- 41.Li L, de G. Jaraquemada-Peláez M, Kuo H-T, Merkens H, Choudhary N, Gitschtaler K, Jermilova U, Colpo N, Uribe-Munoz C, Radchenko V, Schaffer P, Lin K-S, Bénard F and Orvig C, Bioconjugate Chem, 2019, 30, 1539–1553. [DOI] [PubMed] [Google Scholar]
- 42.Karzai FH, Apolo AB, Cao L, Madan RA, Adelberg DE, Parnes H, McLeod DG, Harold N, Peer C, Yu Y, Tomita Y, Lee M-J, Lee S, Trepel JB, Gulley JL, Figg WD and Dahut WL, BJU Int, 2015, 116, 546–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Mendelson DS, Gordon MS, Rosen LS, Hurwitz H, Wong MK, Adams BJ, Alvarez D, Seon BK, Theuer CP and Leigh BR, J. Clin. Oncol, 2010, 28, 3013–3013. [Google Scholar]
- 44.Shannon RD, Acta Crystallogr., Sect. A: Found. Adv, 1976, 32, 751–767. [Google Scholar]
- 45.Bazargan M, Mirzaei M, Franconetti A and Frontera A, Dalton Trans, 2019, 48, 5476–5490. [DOI] [PubMed] [Google Scholar]
- 46.Melton DL, VanDerveer DG and Hancock RD, Inorg. Chem 2006, 45, 9306–9314. [DOI] [PubMed] [Google Scholar]
- 47.Hancock RD, Melton DL, Harrington JM, McDonald FC, Gephart RT, Boone LL, Jones SB, Dean NE, Whitehead JR and Cockrell GM, Coord. Chem. Rev, 2007, 251, 1678–1689. [Google Scholar]
- 48.Hancock RD, J. Chem. Educ, 1992, 69, 615. [Google Scholar]
- 49.Burai L, Fábián I, Király R, Szilágyi E and Brücher E, J. Chem. Soc., Dalton Trans, 1998, 2, 243–248. [Google Scholar]
- 50.McMurry TJ, Pippin CG, Wu C, Deal KA, Brechbiel MW, and Mirzadeh S and O. A. Gansow, J. Med. Chem, 1998, 41, 3546–3549. [DOI] [PubMed] [Google Scholar]
- 51.Price EW, Cawthray JF, Bailey GA, Ferreira CL, Boros E, Adam MJ and Orvig C, J. Am. Chem. Soc, 2012, 134, 8670–8683. [DOI] [PubMed] [Google Scholar]
- 52.Gans P, Sabatini A and Vacca A, Talanta, 1996, 43, 1739–1753. [DOI] [PubMed] [Google Scholar]
- 53.Gans P, Sabatini A and Vacca A, Ann. Chim, 1999, 89, 45–49. [Google Scholar]
- 54.Baur B, Solbach C, Andreolli E, Winter G, Machulla H-J and Reske S, Pharmaceuticals, 2014, 7, 517–529. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Breeman WAP, Jong M, Visser TJ, Erion JL and Krenning EP, Eur. J. Nucl. Med. Mol. Imaging, 2003, 30, 917–920. [DOI] [PubMed] [Google Scholar]
- 56.Pruszyński M, Majkowska-Pilip A, Loktionova NS, Eppard E and Roesch F, Appl. Radiat. Isot, 2012, 70, 974–979. [DOI] [PubMed] [Google Scholar]
- 57.Graves SA, Hernandez R, Fonslet J, England CG, Valdovinos HF, Ellison PA, Barnhart TE, Elema DR, Theuer CP, Cai W, Nickles RJ and Severin GW, Bioconjugate Chem, 2015, 26, 2118–2124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Ehlerding EB, Ferreira CA, Aluicio-Sarduy E, Jiang D, Lee HJ, Theuer CP, Engle JW and Cai W, Mol. Pharmaceutics, 2018, 15, 2606–2613. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Dearling JLJ, Voss SD, Dunning P, Snay E, Fahey F, Smith SV, Huston JS, Meares CF, Treves ST and Packard AB, Nucl. Med. Biol, 2011, 38, 29–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Carroll V, Demoin DW, Hoffman TJ and Jurisson SS, Radiochim. Acta, 2012, 100, 653–667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Price EW and Orvig C, in The Chemistry of Molecular Imaging, John Wiley & Sons, Inc, Hoboken, NJ, 2014, pp. 105–135. [Google Scholar]
- 62.Hillier SM, Maresca KP, Femia FJ, Marquis JC, Foss CA, Nguyen N, Zimmerman CN, Barrett JA, Eckelman WC, Pomper MG, Joyal JL and Babich JW, Cancer Res, 2009, 69, 6932–6940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Lee C, Yang W and Parr RG, Phys. Rev. B, 1988, 37, 785–789. [DOI] [PubMed] [Google Scholar]
- 64.Becke AD, Phys. Rev. A, 1988, 38, 3098–3100. [DOI] [PubMed] [Google Scholar]
- 65.Alliot C, Kerdjoudj R, Michel N, Haddad F and Huclier-Markai S, Nucl. Med. Biol, 2015, 42, 524–529. [DOI] [PubMed] [Google Scholar]
- 66.Hay PJ and Wadt WR, J. Chem. Phys, 1985, 82, 270–283. [Google Scholar]
- 67.Becke AD, J. Chem. Phys, 1993, 98, 5648–5652. [Google Scholar]
- 68.Cao X and Dolg M, J. Chem. Phys, 2001, 115, 7348–7355. [Google Scholar]
- 69.Gran G, Analyst, 1952, 77, 661–671. [Google Scholar]
- 70.Paul MA and Long FA, Chem. Rev, 1957, 57, 1–45. [Google Scholar]
- 71.Vickery RC, J. Chem. Soc, 1952, 0, 1895–1898. [Google Scholar]
- 72.Baes CF and Mesmer RS, The Hydrolysis of Cations, John Wiley & Sons, Robert E. Krieger Publishing Company, Malabor Florida, USA, 1986. [Google Scholar]
- 73.Harris WR, Carrano CJ and Raymond KN, J. Am. Chem. Soc, 1979, 101, 2213–2214. [Google Scholar]
- 74.Aluicio-Sarduy E, Hernandez R, Valdovinos HF, Kutyreff CJ, Ellison PA, Barnhart TE, Nickles RJ and Engle JW, Appl. Radiat. Isot, 2018, 142, 28–31. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
