Abstract
The coordination chemistry of a new pentadentate bifunctional chelator (BFC), NODA-MPAA 1, containing the 1,4,7-triazacyclononane-1,4-diacetate (NODA) motif with a methyl phenyl acetic acid (MPAA) backbone, and its ability to form stable Al18F-chelates, was investigated. The organofluoroaluminates were easily accessible from the reaction of 1 and AlF3. X-ray diffraction studies revealed aluminum at the center of a slightly distorted octahedron, with fluorine occupying one of the axial positions. The tert-butyl protected prochelator 7, which can be synthesized in one step, is useful for coupling to biomolecules on solid phase or in solution. High yield (55–89%) aqueous 18F-labeling was achieved in 10–15 minutes with a tumor-targeting peptide 4 covalently linked to 1. Defluorination was not observed for at least 4 h in human serum at 37 °C. These results demonstrate the facile application of Al18F chelation using BFC 1 as a versatile labeling method for radiofluorinating other heat-stable peptides for positron emission imaging.
INTRODUCTION
Molecular imaging is the non-invasive visualization of cellular or molecular processes that promises to improve the diagnosis and monitoring of various diseases or conditions.1,2 The most commonly used molecular imaging modalities include fluorescence, bioluminescence, positron-emission tomography (PET), and single-photon emission computed tomography (SPECT).3 PET has emerged as a modality of choice, because it yields images with good spatial resolution and excellent sensitivity.4,5 Among the positron-emitting isotopes (11C, 13N, 15O, 18F, 68Ga, 89Zr, 124I), 18F has the advantages of suitable decay properties (t1/2 = 109.7 min, ~ 97% β+-emission, 635 keV), low β+-trajectory (<2 mm), small atomic size, and is readily produced from a stable heavy atom (18O) precursor.6,7 Currently, one of the most widespread applications of PET/CT is the measurement of glucose metabolism using 18F-fluorodeoxyglucose (18F-FDG).8 Radiofluorination traditionally has involved C-18F bond formation that is performed in anhydrous polar aprotic solvents at 80–160 °C via SN2 or SNAr mechanism.9–11 Although the aryl C-18F bonds are generally stable in vivo, the aliphatic C-18F is often prone to enzymatic cleavage. C-18F bonds have also been introduced by the direct electrophilic addition of molecular fluorine (18F2) across double bonds.12 New radiolabeling chemistries, utilizing the fluorides of phosphorus, silicon, and boron have been explored.13–15
We previously reported a technique that exploits the fluorophilic nature of aluminum to afford direct aqueous 18F-labeling by the formation of stable aluminum-fluoride chelates.16 Initial studies involved two chelation systems, the acyclic BFC diethylenetriaminepentaacetic acid (DTPA) and the macrocyclic BFC 1,4,7-triazacyclononane-1,4,7-triacetic acid (NOTA). While the former had excellent binding kinetics, its complexes were not stable, being prone to hydrolysis and demetalation in serum, whereas the NOTA derivative formed highly stable Al18F complexes.16
The hexadentate p-SCN-Bn-NOTA was used to synthesize IMP449 (Figure 1), a hapten-peptide for in vivo targeting of cancer with a bispecific antibody (bsMAb) pretargeting system, a highly sensitive and specific technique for localizing cancer.17–19 Although the Al18F(IMP449) was stable in vivo and displayed excellent tumor to blood (T/B) ratios, radiochemical yields (RCYs) were low (5–20%).16 This could be due to the participation of all three N-acetate arms of the N3O3 donor set in the complexation, resulting in a restricted coordination site for binding 18F. To address this issue, we synthesized IMP461 (Figure 1), a similar hapten-peptide with NOTA attached via one of its N-acetates, and obtained stable complexes, indicating that it was possible to diminish denticity without significant loss of ligand effectiveness.20,21 This prompted us to focus on pentadentate derivatives of 1,4,7-triazacyclononane (TACN) that could form mononuclear octahedral complexes with aluminum and possess a single coordination site for binding fluorine. Herein, we report the coordination chemistry of a new pentadentate ligand 1, containing the 1,4,7-triazacyclononane-1,4-diacetate motif, and its ability to form physiologically stable Al18F-labeled chelates when covalently linked to a hapten-peptide (IMP485) 4 via an amide bond (Figure 2). These results establish the broad applicability of this chemical strategy in the development of 18F-labeled biomolecules useful for PET imaging.
Figure 1.
Structures of IMP449 and IMP461.
Figure 2.
Structures of NODA-MPAA, IMP485, and their aluminum complexes.
RESULTS
The reaction of 1 (Figure 2) with Al3+ was studied in the presence of 19F-. Reacting equimolar amounts of AlCl3 and 1 at 104 °C for 10 min yields a polar chelate 2 with retention time (tR) 8.98 min (Figure 3A). With equimolar amounts of NaF, AlCl3 and 1, the Al19F complex 3 (13.8%) was obtained, with tR 12.29 min (Figure 3B), along with 2 and unreacted 1. Since AlCl3 and NaF readily react22 to produce AlFn(3-n)+, we investigated the reaction between AlF3 and 1. Although the AlF3 remained as a fine suspension in the 2 mM AlF3 solution, we reacted it with 1 at 105 °C for 10 min and obtained a low yield (4.3%) of 3 (Figure 3C). Suspecting that a hydrophilic organic solvent might better solubilize AlF3, we carried out the reaction in the presence of 50 μL methanol, which increased the yield of 3 to 27.7% (Figure 3D). Subsequent experiments revealed that the presence (50% or more v/v) of other hydrophilic organic solvents (e.g., acetonitrile, ethanol, n-propanol, isopropanol, t-butyl alcohol, acetone, dioxane, tetrahydrofuran, dimethylsulfoxide, or dimethyl formamide) also improved the yield of 3.
Figure 3.
HPLC chromatograms using Method 1: (A) AlCl3 with 1; (B) AlCl3 + NaF with 1; (C) AlF3 with 1; (D) AlF3 with 1 in the presence of MeOH; (E) NaF with 2; (F) NaF with 2 in the presence of EtOH.
The next study determined if fluoride, which is bioisosteric with both the hydrogen atom and the hydroxyl group, could displace the hydroxide from 2. When equimolar amounts of 2 and NaF were reacted at 100 °C for 10 min, a mixture containing 3 (32.3%) along with 1 and 2 (Figure 3E) was obtained, indicating a possible substitution mechanism, accompanied by demetalation. When the above reaction was performed in the presence of ethanol, we obtained a two-fold increase in the yield of 3 (Figure 3F).
To determine their stability and structural features, 2 and 3 were synthesized by reacting 1 with AlCl3 or AlF3, respectively (Scheme 1). When the Al19F chelate 3 was incubated in PBS, pH 7.4, at 37 °C, no detectable loss of aluminum was observed by RP- HPLC, after 24 h. However, the corresponding Al-OH complex 2 was not as stable under the same conditions, undergoing progressive demetalation (e.g., 23% in 3 h, 61% in 12 h; see Supporting Information, Table S1).
Scheme 1.
Reaction of 1 with AlCl3 and AlF3.
The HRMS of 1 (M+H)+ 394.20 and that of 3 with (M+H)+ 438.16 confirms the presence of an aluminum-fluoride complex (see Supporting Information, Figure S1). Mass spectrum of 2 has molecular ion peak (M+H)+ 436.17 of low abundance, the parent peak being 418.16, molecular ion minus hydroxide, indicating that the sixth coordination site is occupied by hydroxide instead of chloride.
In the uncomplexed BFC 1, the N–CH2-Ph and Ph-CH2-COOH protons appear as singlets at δ 4.32 and δ 3.61 (Figure 4A). Upon coordination with aluminum fluoride, the N–CH2-Ph protons are split into a pair of doublets at δ 4.24 and δ 3.74 ppm. The region from δ 2.24 to 3.51 ppm is crowded, with several overlapping multiplets arising due to the splitting of the 12 N-CH2-CH2-N ring and 4 N-CH2-COO- protons, along with dissolved water (δ 3.3 ppm) molecules in DMSO-d6 (Figure 4B). These multiplet patterns arise from ring conformational interconversions, which are not so rapid to hinder isomer detection by NMR spectroscopy.23
Figure 4.
1H NMR of 1 and 3 in DMSO-d6 at room temperature (500 MHz).
Mono-hydrated single crystals of 3 suitable for X-ray crystallography were grown from a concentrated solution of acetonitrile containing traces of water. The intensity data for 3 were measured on a Bruker-Nonius KappaCCD diffractometer (graphite- monochromated Mo Kα radiation, λ = 0.71073 Å, φ-ω scans) at 100 (1) K. The data were corrected for absorption. The single-crystal X-ray crystallography (Figure 5) reveals that 3 crystallizes with approximate dimensions 0.060 × 0.080 × 0.52 mm and were monoclinic with space group P21/c. The final unit-cell constants of 3 were a = 20.463(4) b = 8.495(2), c = 12.692(3) Å, β = 107.52(3)°, V = 2103.9(7) Å3, Z = 4, ρ = 1.438 g cm−1, μ = 0.15 mm−1, formula weight = 1821.66. The structure of 3 was solved with SHELXS-97 and refined by full-matrix least squares on F2 with SHELXL-97. The hydrogen atoms were calculated with the riding model in the structure-factor calculations, but their parameters were not refined. The final discrepancy indices, 2.93 < θ < 27.48°, were R = 0.0601 (calculated on F for 3241 reflections) and Rw = 0.1374 (calculated on F2 for all 4812 reflections) with 281 parameters varied. The major peaks of the final difference map, – 0.39 and + 0.32 e Å3, are near the aluminum atom. For crystallographic data (Table S2) and CIF files, see Supporting Information.
Figure 5.
X-ray crystal structure of Al19F(NODA-MPAA) 3
We next turned to examining radiolabeling yields using 18F− in saline. When a mixture of NODA-MPAA 1 (10 μL, 20 nmol), AlCl3 (5 μL, 10 nmol) and 40 μL 18F− (1.326 mCi) in 0.9% saline were reacted at 101 °C for 15 min, decay corrected RCY post purification by solid-phase extraction (SPE) based on 18F− was 6.9%. With 50 μL ethanol as co-solvent, RCY increased to 52.8%, validating our results with 19F−. When the Al-OH chelate 2 (10 μL, 20 nmol) was reacted with 45 μL 18F− in 0.9% saline at 101 °C for 15 min, in the absence and presence of ethanol, decay corrected RCYs were 29.6% and 63.9%, respectively.
Having established the suitability of NODA-MPAA for radiofluorination, the next studies focused on the radiolabeling of the NODA-MPAA pretargteting hapten-peptide, IMP485 4. Under the conditions shown in Table 1, RCYs increased from ~24% in aqueous conditions, to 75–90% in the presence of ethanol and increasing amounts of Na18F. A two-fold increase in RCYs was also observed in the presence of other hydrophilic organic solvents (Table 2). The Radio-HPLC chromatogram (Method 2) of the SPE-purified peptide Al18F(IMP485) shows a single peak with tR 15.789 min (Figure 6A), while the cold standard Al19F(IMP485) appears at 15.620 min (Figure 6B). There was no evidence of defluorination when Al18F(IMP485) was incubated in human serum at 37 °C for 4 h (see Supporting Information, Figure S3). Efficient radiofluorination (decay corrected RCY 72.3%, 1.28 Ci/μmol) was observed when a solution of AlOH(IMP485) 5 (10 μL, 20 nmol), 110 μL ethanol and 100 μL 18F− (43.3 mCi) in 0.9% saline were reacted at 105 °C for 15 min (Table 3). Under identical conditions, labeling efficiency of IMP485 4, was consistently higher than that observed with the NOTA- derived hapten-peptides IMP449 and IMP461 (see Supporting Information, Table S3–S5).
Table 1.
18F-labeling of 20 nmol IMP485 + 10 nmol Al3+ with varying amounts of Na18F
| Activitya Na18F(mCi) | Aqueous(μL) | Ethanol(μL) | Isolated activityb(mCi) | RCYc(%) |
|---|---|---|---|---|
| 1.02 | 50 | 0 | 0.21 | 23.9 |
| 0.92 | 50 | 50 | 0.60 | 79.9 |
| 3.20 | 50 | 50 | 2.35 | 89.3 |
| 6.20 | 50 | 50 | 4.42 | 86.7 |
| 12.34 | 100 | 100 | 8.35 | 84.4 |
| 34.60 | 115 | 110 | 20.3 | 74.1d |
10 μL IMP485, 5 μL Al3+, 105–110 °C, 15 min.
Isolated activity in (1:1) EtOH/H2O after HLB column purification (SPE).
decay corrected RCY – based on synthesis time of 31 – 37 minutes.
Specific activity: 1.01 Ci/μmol.
Table 2.
Effect of solvent on the 18F-labeling of IMP485
| Activitya Na18F(mCi) | Solvent(50 μL) | Isolated activityb(mCi) | RCYc |
|---|---|---|---|
| 1.416 | − | 0.516 | 43.5 |
| 1.326 | EtOH | 1.006 | 89.4 |
| 1.206 | DMF | 0.886 | 86.5 |
| 1.608 | DMSO | 1.152 | 86.0 |
| 1.246 | CH3CN | 0.966 | 91.3 |
35–50 μL Na F, 10 μL IMP485, 5 μL Al, 105 C, 15 min.
Isolated activity in (1:1) EtOH/H2O after HLB column purification (SPE).
decay corrected RCY – based on synthesis time of 26 – 29 minutes.
Figure 6.
HPLC chromatograms (A) SPE-purified Al18F(IMP485). (B) Al19F(IMP485).
Table 3.
18F-labeling of 20 nmol AlOH(IMP485) with varying amounts of Na18F
| Activitya Na18F(mCi) | Aqueous(μL) | Ethanol(μL) | Isolated activityb(mCi) | RCYc(%) |
|---|---|---|---|---|
| 3.51 | 70 | − | 1.019 | 36.4 |
| 3.34 | 70 | 100 | 2.21 | 83.6 |
| 4.07 | 60 | 75 | 2.44 | 73.8 |
| 43.3 | 110 | 110 | 25.6 | 72.3d |
10 μL AlOH(IMP485), 105 °C, 15 min.
Isolated activity in (1:1) EtOH/H2O after HLB column purification (SPE).
decay corrected RCY – based on synthesis time of 32 – 37 minutes.
Specific activity: 1.28 Ci/μmol.
DISCUSSION
With the availability of compact cyclotrons, automated chemical modules and microreactors, PET imaging using 18F has become the modality of choice for molecular imaging of many small molecules.24 Almost 90% of all clinical PET imaging is performed using 18F-FDG, the other 10% includes agents like 18F-NaF used for bone imaging; 18F-FLT, 18F-choline, and 18F-FET in oncology; and 18F-fallypride, 18F-DOPA, and 18F-altanserin in neurosciences.25 Despite its popularity, FDG cannot be used for many neurological, oncological and cardiological conditions, making it imperative to seek other compounds.26 There are numerous receptor-avid biomolecules, like folate, biotin, bombesin (BBN), gastrin-releasing peptide (GRP), RGD, somatostatin, antibodies, peptides, and proteins, which possess high specificity and affinity.27,28 Development of a facile method of tagging these molecules with 18F would enhance their utility in the detection of pathologically distinct cellular targets. Most 18F-labeling methods are tedious to perform and require the efforts of specialized chemists. Multiple purifications of intermediates are commonly required, resulting in low RCYs. Additionally, 18F-labeling involving C-18F bond formation, which is employed in all clinically-approved 18F-labeled products, requires anhydrous conditions.29 The newer labeling strategies, involving B-18F and Si-18F bond formation, have shown promise, but suffer from the disadvantage of rather high lipophilicity of the building block,30,31 a property that can enhance hepatobiliary clearance in vivo. An ideal method is one in which a peptide could be labeled rapidly in the final step with high specific activity in an aqueous medium.
With experience gained in rapid and quantitative binding of various radiometals with chelates, and with the knowledge that 18F− avidly binds to Al3+, we initiated a project to determine whether a stable Al18F complex could be formed with peptides coupled with an appropriate BFC. Our initial studies were performed with peptides covalently linked to the hexadentate p-SCN-Bn-NOTA and NOTA (Figure 1) via one of its N-acetates.16,21 These peptides yield highly stable Al18F-chelates (in vitro and in vivo), but with low RCYs. We rationalized that with a hexadentate ligand, one of the pendant carboxylates might not participate in the complexation, and the presence of uncoordinated donor atoms potentially could impose steric constraints on the coordinated ligand framework, resulting in isomerism and/or instability. We then set out to synthesize a BFC that could form mononuclear octahedral complexes with aluminum, leaving a vacant site for binding fluorine. Since TACN-derived BFCs are suitable for Al3+ (0.53 Å), we decided to use NO2AtBu with two pendant N-acetate arms as our building block. BFCs with pendant N-acetates are common, since they do not affect the molecular weight significantly and their high hydrophilicity favors rapid renal clearance. The pentadentate NODA-MPAA 1, with its N3O2 donor set, was chosen because the tert-butyl protected prochelator 7 can be synthesized by one simple alkylation (Scheme 2). The methyl phenyl acetic acid (MPAA) backbone acts as a spacer and makes for easy linkage to biomolecules on solid phase or in solution.
Scheme 2.
Synthesis of NODA-MPAA 1 and (tBu)2NODA-MPAA 7.
The X-ray structure (Figure 5) confirms the ability of 1 to form an Al19F monohydrated complex, with the MPAA backbone distanced from the coordination sphere. The complex is neutral with aluminum located at the center of a slightly distorted octahedron, as evidenced by the deviation of the twelve quasi-orthogonal and three linear bond angles, from the ideal angles of 90 and 180° in a regular octahedron. The average of these twelve quasi-orthogonal angles and that of the three linear bond angles are 89.89° and 169.60°, respectively (Table 4). The two TACN ring nitrogens and two oxygens, one from each pendant carboxylate, lie in the equatorial plane. One axial plane is occupied by the third nitrogen of TACN, while the fluorine lies in the opposite axial position, completing the octahedral coordination sphere. The diversity of the bond distances between the aluminum and the hetero atoms N1, N4, N7, O12, O14, and F also contributes to the asymmetry of the aluminum coordination sphere. The Al-N4 bond is significantly shorter (2.036 Å) compared to the Al-N1 (2.074 Å) and Al-N7 (2.080 Å) bonds, while the Al-O14 (1.875 Å) and Al-O12 (1.854 Å) bond lengths are also slightly different. The longer Al-N1 (2.074 Å) bond could be a reflection of the high affinity of the trans fluoro ligand for aluminum. In comparison with the crystallized Al(NOTA), with Al-N and Al-O bond lengths of 2.06 Å and 1.84 Å, respectively,32 the small difference in the average Al-N (2.063 Å) and Al-O (1.865 Å) bond lengths in 3 indicates that aluminum is strongly coordinated by the N3O2 donor set.
Table 4.
Bond lengths (Å) and bond angles (°) for complex 3
| Bond lengths (Å) | Bond angles (°) | ||
|---|---|---|---|
| Al-F | 1.714 (1) | F-Al-N1 | 176.2 (1) |
| Al-O12 | 1.854 (2) | F-Al-N4 | 96.8 (1) |
| Al-O14 | 1.875 (2) | F-Al-N7 | 91.8 (1) |
| Al-N1 | 2.074 (2) | F-Al-O12 | 97.7 (1) |
| Al-N4 | 2.036 (2) | F-Al-O14 | 93.3 (1) |
| Al-N7 | 2.080 (2) | O12-Al-O14 | 95.6 (1) |
| N1-C2 | 1.487 (3) | O12-Al-N1 | 82.6 (1) |
| N1-C9 | 1.506 (3) | O12-Al-N4 | 165.4 (1) |
| N1-C10 | 1.482 (3) | O12-Al-N7 | 95.3 (1) |
| N4-C3 | 1.505 (3) | O14-Al-N1 | 90.4 (1) |
| N4-C5 | 1.492 (3) | O14-Al-N4 | 83.3 (1) |
| N4-C12 | 1.483 (3) | O14-Al-N7 | 167.2 (1) |
| N7-C6 | 1.504 (3) | N1-Al-N4 | 82.9 (1) |
| N7-C8 | 1.496 (3) | N1-Al-N7 | 84.5 (1) |
| N7-C14 | 1.594 (2) | N4-Al-N7 | 84.4 (1) |
When incubated in PBS, pH 7.4, at 37 °C, Al19F(NODA-MPAA) 3 was stable for over 24 h, while the corresponding Al-OH complex 2 underwent demetalation to the extent of 23% in 3 h (see Supporting Information, Table S1). The fact that 3 did not undergo demetalation was gratifying since a stable Al18/19F chelate is our most important prerequisite. It was interesting to note that fluoride, which is bioisosteric with both the hydrogen atom and the hydroxyl group, could displace the hydroxide from 2 (Figure 3E). The driving force for this reaction may be the formation of the strong Al-F bond (580 – 670 kJ/mol) in the product.
We believe that the improved labeling efficiency with peptides covalently linked to NODA-MPAA is due to the presence of a N3O2 donor set that can form stable mononuclear octahedral complexes with aluminum and possess a single coordination site for binding fluorine. The N3O3 donor set (hexadentate) of TACN derivatives like NOTA provide a suitable environment for the formation of highly stable aluminum chelates.23 Low RCYs with IMP449 (N3O3 donor set) may result from the coordination sphere of aluminum being saturated, with limited access for fluorine. A similar situation, although with less restriction, could occur with IMP461, due to the participation of the amido oxygen in the complexation. Under identical conditions, RCYs were higher with IMP485, when compared to IMP449 and IMP461 (see Supporting Information, Tables S3–S5). These results suggest that pentadentate ligands with a free site for binding fluorine are ideal for stable Al18F chelation.
The 18F− produced in the cyclotron comes out as a dilute solution in H218O and is generally contaminated with metals from the target cell. Since our 18F-labeling method involves the use of a BFC, the presence of competing metal ions would hamper effective labeling. 33 Initial labeling studies were performed with 18F− that was bound onto a (anion exchange resin) QMA cartridge, eluted with 0.4 M KHCO3, followed by neutralization of the KHCO3 with acetic acid. With pH being an important parameter for radiofluorination, inconsistencies in the neutralization resulted in varying RCYs. Recently, we found that QMA-bound 18F− could be eluted with 0.9% saline and used without making any pH adjustments.20 Subsequently, we obtained high RCYs with the commercially available USP grade Na18F, an approved bone-imaging agent.34 This was convenient, since it eliminated the concentration, purification and dry-down steps associated with other 18F-labeling methods.
We also demonstrate herein two routes to synthesize the 18F-labeled peptide, either heat the hapten-peptide (IMP485) 4 and Al3+ with commercially available 18F− in saline or synthesize the peptide-aluminum complex 5 and react it with Na18F (see Supporting Information, Scheme S1). The latter approach is attractive since it is based on the reaction a single reactant with Na18F, eliminating any speculation regarding peptide to Al3+ ratios or inadvertently not adding any Al3+. One of the key findings during the development of this methodology was that the presence of an added hydrophilic organic co-solvent (generally 50% or more v/v) doubled the RCYs (Table 2). However, ethanol was the co-solvent of choice for most of our subsequent radiofluorinations, since it is biocompatible and has the added advantage of preventing radiolysis.35 Based on these findings, we explored the possibility of developing a single-vial kit, requiring only the addition of 18F− in saline.36 These studies included an examination of the effects of temperature, reaction time, pH, amount of peptide, peptide to Al3+ ratio, and reaction volume on the radiolabeling efficiency. The development of the lyophilized kit formulation of IMP485, as well as kits for NODA-MPAA derived somatostatin receptor and bombesin binding peptides, will be reported elsewhere.
The theoretical specific activity (SA) of carrier-free 18F− is 1702 Ci/μmol, but varies from 8.5 to 1162 Ci/μmol, due to isotopic dilution from water, reagents and reaction vessels.37 In most clinical applications, 5 to 10 mCi of the 18F-labeled product with specific activities >1 Ci/μmol are required for imaging cellular targets.38 We recently reported the development of a lyophilized kit formulation containing 20 nmol IMP485 that has been radiolabeled to a specific activity of ~4 Ci/μmol.36 This would allow for 4 h from start of radiolabeling to the injection of the patient to retain a SA of 1 Ci/μmol. When the reaction mixture is purified by SPE, one has to bear in mind that the product is a mixture consisting of unlabeled material (IMP485) 4 plus AlOH(IMP485) 5, with Al18/19F-labeled product. As seen in the Supporting Information Figures S4 and S5, excellent separation of Al19F(IMP485) 6 from (IMP485) 4 and AlOH(IMP485) 5 using a C18 column is possible. Therefore, the SA could be increased if the post-labeling purification was by RP-HPLC instead of SPE. However, as mentioned earlier, the precise SA of the 18F-labeled product will also depend on the SA of the 18F−.37 Due to the short half-life of 18F, labeling is frequently performed with high levels of radioactivity, and this in turn imposes a requirement that the radiochemistry be performed in an apparatus that is shielded from the operator and is fully automated.39 Electrochemical methods, currently being explored for the concentration of no-carrier-added 18F− will help in reducing the reaction volume.40 We are confident that 18F-labeling via Al18F chelation, particularly using single-vial kits, can be adapted for use in these automated systems, and would also benefit from microfluidic technology.41
Advancement of an 18F-labeled peptide into the clinic requires that it possess good in vivo stability, excellent targeting, and favorable excretion profile. Animal studies performed using kit formulated IMP485 have shown rapid clearance from the body via the kidneys with negligible presence in the bone, supporting in vivo stability.42 To evaluate the true potential of this novel BFC 1, we have synthesized and 18F-labeled somatostatin receptor-binding and bombesin peptides, and are currently investigating their tumor targeting capabilities.
Finally, while we appreciate that 18F-labeling via Al18F chelation is facile, it has only been described for peptides that are stable at high temperatures. Recently, we reported a maleimide derivative of NODA-MPAA that can be 18F-labeled and then coupled with high efficiency to thiol-containing peptides and proteins.43 This two-step approach opens the possibility for adapting the Al18F chelation to heat-sensitive biomolecules.
CONCLUSIONS
A new pentadentate BFC 1 was synthesized and reacted with AlF3 to yield an organofluoroaluminate 3, which was characterized by RP-HPLC, 1H and 13C NMR, HRMS, and X-ray crystallography. The solid-state X-ray crystal structure of 3 reveals aluminum coordinated by an N3O2 donor set, in slightly distorted octahedron geometry, with fluorine occupying one of the axial positions. The stability of 3 was confirmed when it showed no signs of demetalation over 24 h, when incubated in PBS, pH 7.4, at 37 °C. The partial solubility of AlF3 led us the realization that the addition of a hydrophilic organic co-solvent improves the yield of the organofluoroaluminate. It is interesting to note that the reaction of 1 with AlCl3 and AlF3 yields two distinct chelates, (Al-OH) 2 and (AlF) 3, respectively. High yield aqueous radiofluorination was accomplished in a single step, by heating a mixture of hapten-peptide 4 and Al3+ or the peptide-aluminum complex 5, with commercially available 18F− in saline. The post-labeling purification method (SPE or RP-HPLC) employed was dictated by specific activity requirements. We hope that this robust, user-friendly 18F-labeling methodology will stimulate the rapid development of new PET tracers for the imaging of disease states.
MATERIALS AND METHODS
General Information
All commercially obtained chemicals were analytical grade and used without further purification. AlCl3·6H2O, NaF, AlF3·3H2O, and 4-(bromomethyl)-phenylacetic acid were purchased from Sigma-Aldrich (Milwaukee, WI) and sodium acetate and acetic acid from J. T. Baker (Phillipsburg, NJ). NO2AtBu was purchased from CheMatech (Dijon, France).
Fmoc protected amino acids, seiber amide resin, trifluoroacetic acid were obtained from Creosalus (Louisville, KY). The peptides were synthesized using Protein Technologies, Inc. (Tucson, AZ) peptide synthesizer. The analytical and preparative reverse-phase HPLC (RP-HPLC) columns were purchased from Phenomenex (Torrance, CA) and Waters Corp. (Milford, MA). Two millimolar solutions of AlCl3, NaF, AlF3, and 1 were prepared in 2 mM NaOAc, the pH being maintained in the range of 3.8 to 4.4. Since 1 and 4 displayed a high propensity for complexation with other metals (e.g., copper, gallium and indium) at room temperature, care was taken to ensure that reactants were prepared in metal free containers. Reactions were performed in sealed 2-mL microcentrifuge tubes with cap and O-ring obtained from Fisher Scientific (Agawan, MA). Radiolabeled peptides were purified using Waters Oasis 1 cm3 or 3 cm3 flangeless cartridges. 18F− in saline was purchased from PETNET (Hackensack, NJ). Solid-phase extraction (SPE) cartridges (Sep-Pak light QMA, Sep-Pak Accell plus CM, and Oasis HLB) were obtained from Waters (Milford, MA). 1H and 13C NMR spectra were recorded using Varian Inova NMR spectrometer (Varian, Inc., Palo Alto, CA) at 500 MHz for 1H and 125.7 MHz for 13C at Rutgers University (Newark, NJ). 1H NMR spectra are referenced to the tetramethylsilane peak (δ = 0). Product identification was verified by high-resolution mass spectroscopy (HRMS) using positive mode electrospray ionization with an Agilent Time-of-Flight LC-MS at Immunomedics, Inc. (Morris Plains, NJ). X-ray crystallography data were collected on Bruker-Nonius KappaCCD (Mo Kα radiation) diffractometer at Hunter College of the City University of New York (New York, NY). All calculations were performed using Bruker SHELXS-97 software package, while SHELXL-97 was used for refinement.44
HPLC Methods
Analytical HPLC was performed using a Waters 2695 system equipped with a Phenomenex Gemini C18 reverse-phase column (250 × 4.6 mm, 5 μm, 110 Å).
Method 1
Gradient of 1 min with 100% A (0.1% TFA), then to 90:10 A/B over 5 min, followed by 85:15 A/B (90% acetonitrile, 10% water, 0.1% TFA) over 30 min at a flow rate of 1 mL/min, absorbance was detected at 220 and 254 nm using Waters 2996 photodiode array (PDA) detector.
Method 2
Gradient of 1 min with 100% A (0.1% TFA), then to 90:10 A/B over 5 min, followed by 87:13 A/B (90% acetonitrile, 10% water, 0.1% TFA) over 20 min at a flow rate of 1 mL/min, absorbance was detected at 220 and 254 nm using Waters 2996 photodiode array (PDA) detector. The column effluent was monitored using Perkin Elmer 610TR Radiomatic Flow scintillation analyzer.
Method 3
Linear gradient of 100% C (H2O) to 30% D (50% ethanol) over 20 min at a flow rate of 1 mL/min, absorbance was detected at 220 and 254 nm using Waters 2996 photodiode array (PDA) detector.
Method 4
Products were purified using Waters PrepLC 4000 system with Sunfire Prep C18 OBD reverse-phase column (150 × 30 mm, 5 μm) using a linear gradient of 100% A (0.1% TFA) to 15% B (90% acetonitrile, 10% water, 0.1% TFA) over 80 minutes at a flow rate of 45 mL/min, absorbance was detected at 220 nm using Waters 486 tunable absorbance detector.
Method 5
Product was purified using Waters PrepLC 4000 system with Sunfire Prep C18 OBD reverse-phase column (150 × 30 mm, 5 μm) using a linear gradient of 100% A (0.006 N HCl) to 100% B (acetonitrile) over 60 minutes at a flow rate of 45 mL/min, absorbance was detected at 220 nm using Waters 486 tunable absorbance detector.
Synthesis of the BFCs
2-(4-(carboxymethyl)-7-{[4-(carboxymethyl)phenyl]methyl}-1,4,7-triazacyclononan- 1-yl)acetic acid [H2L] (1). (NODA-MPAA)
To a solution of 4-(bromomethyl)phenylacetic acid (15.7 mg, 0.68 mmol) in anhydrous CH3CN at 0 °C was added dropwise over 20 min a solution of NO2AtBu (26 mg, 0.73 mmol) in CH3CN (5 mL). After 2 h, anhydrous K2CO3 (5 mg) was added to the reaction mixture and allowed to stir at room temperature overnight. Solvent was evaporated and the concentrate was acidified with 2 mL TFA. After 3 h, the reaction mixture was diluted with water and purified by preparative RP-HPLC (Method 4) to yield a white solid (11.8 mg, 43.7%). 1H NMR (500 MHz, DMSO-d6, 25 °C) δ 2.65–3.13 (m, 12 H), 3.32 (d, 2H), 3.47 (d, 2H), 3.61 (s, 2 H), 4.32 (s, 2 H), 7.33 (d, 2H), 7.46 (d, 2H); 13C (125.7 MHz, DMSO-d6) 40.8, 47.2, 49.6, 50.7, 55.2, 58.1, 130.4, 130.5, 130.9, 136.6, 158.4, 158.7, 172.8, 172.9. HRMS (ESI) calculated for C19H27N3O6 (M+H)+ 394.1973; found 394.1979.
2-{4-[(4,7-bis-tert-butoxycarbonylmethyl)-[1,4,7]-triazacyclononan-1- yl)methyl]phenyl}acetic acid (7)
To a solution of 4-(bromomethyl)phenylacetic acid (593 mg, 2.59 mmol) in anhydrous CH3CN (50 mL) at 0 °C were added dropwise over 1 h a solution of NO2AtBu (1008 mg, 2.82 mmol) in CH3CN (50 mL). After 4 h, anhydrous K2CO3 (100.8 mg, 0.729 mmol) was added to the reaction mixture and allowed to stir at room temperature overnight. Solvent was evaporated and the crude was purified by preparative RP-HPLC (Method 5) to yield a white solid (713 mg, 54.5%). 1H NMR (500 MHz, CDCl3, 25 °C, TMS) δ 1.45 (s, 18 H), 2.64–3.13 (m, 16 H), 3.67 (s, 2 H), 4.38 (s, 2 H), 7.31 (d, 2H), 7.46 (d, 2H); 13C (125.7 MHz, CDCl3) δ 28.1, 41.0, 48.4, 50.9, 51.5, 57.0, 59.6, 82.3, 129.0, 130.4, 130.9, 136.8, 170.1, 173.3. HRMS (ESI) calculated for C27H43N3O6 (M+H)+ 506.3225, found 506.3210.
Synthesis of Aluminum chelates
AlOH(NODA-MPAA) (2)
H2L 1 (19.8 mg, 0.05 mmol) was dissolved in 1 mL of 2 mM NaOAc, pH 4.4 and treated with (12 mg, 0.05 mmol) AlCl3·6H2O. The pH was adjusted to 4.5–5.0 and the reaction mixture was refluxed for 15 min. The crude was purified by preparative RP-HPLC (Method 4) to yield a white solid (10.9 mg, 49.8%). HRMS (ESI) calculated for C19H26AlN3O7 (M+H)+ 436.1659; found 436.1664.
AlF(NODA-MPAA) (3)
H2L 1 (40.6 mg, 0.103 mmol) was dissolved in 1 mL of 2 mM NaOAc, pH 4.4, 0.5 mL ethanol and treated with ( 19.5 mg, 0.141mmol) AlF3·3H2O. The pH was adjusted to 4.5–5.0 and the reaction mixture was refluxed for 15 min. On cooling, the pH was once again raised to 4.5–5.0 and the reaction mixture refluxed for 15 min. The crude was purified by preparative RP-HPLC (Method 4) to yield a white solid (22.4 mg, 49.6%). 1H NMR (500 MHz, DMSO-d6, 25 °C) δ 2.24–3.51 (m, 16 H), 3.59 (s, 2 H), 3.74 (d, 1H), 4.24 (d, 1H), 7.29 (d, 2H), 7.40 (d, 2H); 13C (125.7 MHz, DMSO-d6) 40.7, 46.0, 51.9, 52.7, 53.0, 54.5, 59.6, 64.0, 129.9, 131.3, 132.5, 135.8, 172.0, 172.2, 172.9. HRMS (ESI) calculated for C19H25AlFN3O6 (M+H)+ 438.1616; found 438.1627.
Synthesis of the Peptide and its Aluminum chelates
IMP485 (4)
NODA-MPAA-D-Lys(HSG)-D-Tyr-D-Lys(HSG)-NH2. The peptide was synthesized on Sieber amide resin with the amino acids added in the following order: Aloc-D-Lys(Fmoc)-OH, Trt-HSG-OH, Aloc removal, Fmoc-D-Tyr(But)-OH, Aloc-D Lys(Fmoc)-OH, Trt-HSG-OH, Aloc removal, and (tBu)2NODA-MPAA. The peptide was then cleaved and purified by preparative RP-HPLC (Method 4). HRMS (ESI) calculated for C62H89N17O15 (M+H)+ 1312.6797; found 1312.6815.
AlOH(IMP485) (5)
To IMP485 (3) (21.5 mg, 0.016 mmol) was dissolved in 1 mL of 2 mM NaOAc, pH 4.4 and treated with (13.2 mg, 0.055mmol) AlCl3·6H2O. The pH was adjusted to 4.5–5.0 and the reaction mixture was refluxed for 15 min. The crude was purified by preparative RP-HPLC (Method 4) to yield a white solid (11.8 mg). HRMS (ESI) calculated for C62H88AlN17O16 (M+H)+ 1354.6483; found 1354.6431.
Al19F(IMP485) (6)
To IMP485 (3) (16.5 mg, 0.013 mmol) was dissolved in 1 mL of 2 mM NaOAc, pH 4.43, 0.5 mL ethanol and treated with (2.5 mg, 0.018 mmol) AlF3·3H2O. The pH was adjusted to 4.5–5.0 and the reaction mixture was refluxed for 15 min. On cooling the pH was once again raised to 4.5–5.0 and the reaction mixture refluxed for another 15 min. The crude was purified by preparative RP-HPLC (Method 4) to yield a white solid (10.3 mg). HRMS (ESI) calculated for C62H87AlFN17O15 (M+H)+ 1356.6440; found 1356.6458.
18F-labeling of IMP485
10 μL (20 nmol) IMP485, 5 μL AlCl3 (10 nmol), 100 μL Na18F in 0.9% saline (34.6 mCi), and 110 μL ethanol were reacted in 2 mL microcentrifuge tube (sealed) at 105 °C for 15 min. The reaction mixture was cooled, diluted with 2–3 mL deionized water and then transferred into a HLB 3 cm3 cartridge. The solution was eluted under vacuum into an empty 10 mL crimp-sealed vial. The reaction vessel and column were rinsed with 3 × 1 mL portions of water. The HLB cartridge was then transferred to another empty 3 mL vial and the product eluted with 4 × 150 μL 1:1 ethanol/water to yield 20.3 mCi of Al18F(IMP485) (decay corrected RCY 74.1%; SA 1.01 Ci/μmol).
18F-labeling of AlOH(IMP485)
10 μL (20 nmol) AlOH(IMP485), 100 μL Na18F in 0.9% saline (43.3 mCi), and 110 μL ethanol were reacted in 2 mL microcentrifuge tube (sealed) at 105 °C for 15 min. The reaction mixture was cooled, diluted with 2–3 mL deionized water and then transferred into a HLB 3 cm3 cartridge. The solution was eluted under vacuum into an empty 10 mL crimp-sealed vial. The reaction vessel and column were rinsed with 3 × 1 mL portions of water. The HLB cartridge was then transferred to another empty 3 mL vial and the product eluted with 4 × 150 μL 1:1 ethanol/water to yield 25.6 mCi of Al18F(IMP485) (decay corrected RCY 72.3%; SA 1.28 Ci/μmol).
Radio-HPLC
The crude reaction mixture containing unbound 18F− and the product obtained after purification by SPE were analyzed by RP-HPLC (Method 2) using a Perkin Elmer 610TR Radiomatic Flow scintillation analyzer (see Supporting Information, Figure S2), with recoveries of 79.7% and 100.6%, respectively.
Serum Stability
The SPE purified Al18F(IMP485) in 20 μL 1:1 EtOH/H2O (66.3 μCi) was mixed with 200 μL of human serum, placed in the HPLC autosampler heated to 37 °C and injected 1, 2 and 4 μL at time 0, 2 and 4 h, respectively, to account for physical decay (see Supporting Information, Figure S3). No detectable 18F− above background at the void volume was observed up to 4 h. In addition, percent recovery for each run was determined by counting the activity in 6 fractions over the entire 20 min run against a standard prepared from the original mixture. Total recoveries ranged from ~76% to 91%.
Supplementary Material
Acknowledgments
This work was funded in part by NIH grant 5R44RR028018. The facility at which the X-ray crystallography was performed is supported by a “Research Centers in Minority Institutions” award, RR-03037, from the National Center for Research Resources, NIH.
Footnotes
Disclaimer: CAD, WJM and DMG are employed or have financial interests in Immunomedics, Inc. RMS and LJT have disclosed no financial conflicts.
Supporting Information Available. HPLC chromatograms, crystallographic data and CIF files. This information is available free of charge via the Internet at http://pubs.acs.org.
References
- 1.Signore A, Mather SJ, Piaggio G, Malviya G, Dierckx RA. Molecular imaging of inflammation/infection: nuclear medicine and optical imaging agents and methods. Chem Rev. 2010;110:3112–3145. doi: 10.1021/cr900351r. [DOI] [PubMed] [Google Scholar]
- 2.Pysz MA, Gambhir SS, Willmann JK. Molecular imaging: current status and emerging strategies. Clin Radiol. 2010;65:500–516. doi: 10.1016/j.crad.2010.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Wong FC, Kim EE. A review of molecular imaging studies reaching the clinical stage. Eur J Radiol. 2009;70:205–211. doi: 10.1016/j.ejrad.2009.01.049. [DOI] [PubMed] [Google Scholar]
- 4.Ametamey SM, Honer M, Schubiger PA. Molecular imaging with PET. Chem Rev. 2008;108:1501–1516. doi: 10.1021/cr0782426. [DOI] [PubMed] [Google Scholar]
- 5.Chen K, Conti PS. Target-specific delivery of peptide-based probes for PET imaging. Adv Drug Deliv Rev. 2010;62:1005–1022. doi: 10.1016/j.addr.2010.09.004. [DOI] [PubMed] [Google Scholar]
- 6.Miller PW, Long NJ, Vilar R, Gee AD. Synthesis of 11C, 18F, 15O, and 13N radiolabels for positron emission tomography. Angew Chem Int Ed. 2008;47:8998–9033. doi: 10.1002/anie.200800222. [DOI] [PubMed] [Google Scholar]
- 7.Cai L, Lu S, Pike VW. Chemistry with [18F]fluoride ion. Eur J Org Chem. 2008:2853–2873. [Google Scholar]
- 8.Wadsak W, Mitterhauser M. Basic principles of radiopharmaceuticals for PET/CT. Eur J Radiol. 2010;73:461–469. doi: 10.1016/j.ejrad.2009.12.022. [DOI] [PubMed] [Google Scholar]
- 9.Schirrmacher R, Wängler C, Schirrmacher E. Recent developments and trends in 18F-radiochemistry: syntheses and applications. Mini-Rev Org Chem. 2007;4:317–329. [Google Scholar]
- 10.Kim DW, Ahn DK, Oh YH, Lee S, Kil HS, Oh SJ, Lee SJ, Kim JS, Ryu JS, Moon DH, Chi DY. A new class of SN2 reactions catalyzed by protic solvents: facile fluorination for isotopic labeling of diagnostic molecules. J Am Chem Soc. 2006;128:16394–16397. doi: 10.1021/ja0646895. [DOI] [PubMed] [Google Scholar]
- 11.Becaud J, Mu L, Karramkam M, Schubiger PA, Ametamey SM, Graham K, Stellfeld T, Lehman L, Borkowski S, Berndorff D, Dinkelborg L, Srinivasan A, Smits K, Koksch B. Direct one-step 18F-labeling of peptides via nucleophilic aromatic substitution. Bioconjugate Chem. 2009;20:2254–2261. doi: 10.1021/bc900240z. [DOI] [PubMed] [Google Scholar]
- 12.Kachur AV, Dolbier WR, Xu W, Kock CJ. Catalysis of fluorine addition to double bond: an improvement of method for synthesis of 18F PET Agents. Appl Radiat Isot. 2010;68:293–296. doi: 10.1016/j.apradiso.2009.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Studenov AR, Adam MJ, Wilson JS, Ruth TJ. New radiolabelling chemistry: synthesis of phosphorus-[18F]fluorine compounds. J Label Compd Radiopharm. 2005;48:497–500. [Google Scholar]
- 14.Ting R, Adam MJ, Ruth TJ, Perrin DM. Arylfluoroborates and alkylfluorosilicates as potential PET imaging agents: high-yielding aqueous biomolecular 18F-labeling. J Am Chem Soc. 2005;127:13094–13095. doi: 10.1021/ja053293a. [DOI] [PubMed] [Google Scholar]
- 15.Ting R, Lo J, Adam MJ, Ruth TJ, Perrin DM. Capturing aqueous [18F]-fluoride with an arylboronic ester for PET: synthesis and aqueous stability of a fluorescent [18F]-labeled aryltrifluoroborate. J Fluorine Chem. 2008;129:349–358. [Google Scholar]
- 16.McBride WJ, Sharkey RM, Karacay H, D’Souza CA, Rossi EA, Laverman P, Chang CH, Boerman OC, Goldenberg DM. A novel method of 18F radiolabeling for PET imaging. J Nucl Med. 2009;50:991–998. doi: 10.2967/jnumed.108.060418. [DOI] [PubMed] [Google Scholar]
- 17.Schoffelen R, Sharkey RM, Goldenberg DM, Franssen G, McBride WJ, Rossi EA, Chang CH, Laverman P, Disselhorst JA, Eek A, van der Graaf WTA, Oyen WJG, Boerman OC. Pretargeted immunoPET imaging of CEA-expressing tumors with a bispecific antibody and a 68Ga and 18F-labeled hapten peptide in mice with human tumor xenografts. Mol Cancer Ther. 2010;9:1019–1027. doi: 10.1158/1535-7163.MCT-09-0862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Sharkey RM, Karacay H, Cardillo TM, Chang CH, McBride WJ, Rossi EA, Horak ID, Goldenberg DM. Improving the delivery of radionuclides for imaging and therapy of cancer using pretargeting methods. Clin Cancer Res. 2005;11:7109s–7121s. doi: 10.1158/1078-0432.CCR-1004-0009. [DOI] [PubMed] [Google Scholar]
- 19.Rossi EA, Goldenberg DM, Cardillo TM, McBride WJ, Sharkey RM, Chang CH. Stably tethered multifunctional structures of defined composition made by dock and lock method for use in cancer targeting. Proc Natl Acad Sci USA. 2006;103:6841–6846. doi: 10.1073/pnas.0600982103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.McBride WJ, D’Souza CA, Sharkey RM, Karacay H, Rossi EA, Chang CH, Goldenberg DM. Improved 18F labeling of peptides with a fluoride-aluminum-chelate complex. Bioconjugate Chem. 2010;21:1331–1340. doi: 10.1021/bc100137x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Laverman P, McBride WJ, Sharkey RM, Eek A, Joosten L, Oyen WJG, Goldenberg DM, Boerman OC. A novel facile method of labeling octreotide with 18F-fluorine. J Nucl Med. 2010;51:454–461. doi: 10.2967/jnumed.109.066902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Bodor A, Tóth I, Bányai I, Szabó Z, Hefter GT. 19F NMR study of the equilibria and dynamics of the Al3+/F− system. Inorg Chem. 2000;39:2530–2537. doi: 10.1021/ic991248w. [DOI] [PubMed] [Google Scholar]
- 23.de Sa A, Prata MIM, Geraldes CFGC, André JP. Triaza-based amphiphilic chelators: synthetic route, in vitro characterization and in vivo studies of their Ga(III) and Al(III) chelates. J Inorg Biochem. 2010;104:1051–1062. doi: 10.1016/j.jinorgbio.2010.06.002. [DOI] [PubMed] [Google Scholar]
- 24.Elizarov AM. Microreactors for radiopharmaceutical synthesis. Lab Chip. 2009;9:1326–1333. doi: 10.1039/b820299k. [DOI] [PubMed] [Google Scholar]
- 25.Coenen HH, Elsinga PH, Iwata R, Kilbourn MR, Pillai MRA, Rajan MGR, Wagner HN, Zaknum JJ. Fluorine-18 radiopharmaceuticals beyond [18F]FDG in oncology and neurosciences. Nucl Med Biol. 2010;37:727–740. doi: 10.1016/j.nucmedbio.2010.04.185. [DOI] [PubMed] [Google Scholar]
- 26.Nanni C, Fantini L, Nicolini S, Fanti S. Non FDG PET. Clin Radiol. 2010;65:536–548. doi: 10.1016/j.crad.2010.03.012. [DOI] [PubMed] [Google Scholar]
- 27.Lee S, Xie J, Chen X. Peptide-based probes for targeted molecular imaging. Biochemistry. 2010;49:1364–1376. doi: 10.1021/bi901135x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Schottelius M, Wester HJ. Molecular imaging targeting peptide receptors. Methods. 2009;48:161–177. doi: 10.1016/j.ymeth.2009.03.012. [DOI] [PubMed] [Google Scholar]
- 29.Flavell RR, Kothari P, Bar-Dagan M, Synam M, Vallabhajosula S, Friedman JM, Muir TW, Ceccarini G. Site-specific 18F-labeling of the protein hormone leptin using a general two-step ligation procedure. J Am Chem Soc. 2008;130:9106–9112. doi: 10.1021/ja801666z. [DOI] [PubMed] [Google Scholar]
- 30.Ting R, Aguilera TA, Crisp JL, Hall DJ, Eckelman WC, Vera DR, Tsein RY. Fast 18F labeling of a near-infrared fluorophore enables positron emission tomography and optical imaging of sentinal lymph nodes. Bioconjugate Chem. 2010;21:1811–1819. doi: 10.1021/bc1001328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Wängler C, Waser B, Alke A, Iovkova L, Buchholz HG, Niedermoser S, Jurkschat K, Fottner C, Schirrmacher R, Reubi JC, Wester HJ, Wängler B. One-step 18F labeling of carbohydrate-conjugated octreotate-derivatives containing a silicon-fluoride-acceptor (SiFA): in vitro and in vivo evaluation as tumor imaging agents for positron emission tomography (PET) Bioconjugate Chem. 2010;21:2289–2296. doi: 10.1021/bc100316c. [DOI] [PubMed] [Google Scholar]
- 32.Jyo A, Kohno T, Terazono Y, Kawano S. Crystal structure of the aluminum(III) complex of 1,4,7-triazacyclononane-N,N′,N″-triacetate. Anal Sci. 1990;6:629–630. [Google Scholar]
- 33.Svadberg A, Clarke A, Dyrstad K, Martinsen I, Hjelstuen OK. A critical study on borosilicate glassware and silica-based QMA’s in nucleophilic substitution with [18F]fluoride: influence of aluminum, boron and silicon on the reactivity of [18F]fluoride. Appl Radiat Isot. 2011;69:289–294. doi: 10.1016/j.apradiso.2010.09.013. [DOI] [PubMed] [Google Scholar]
- 34.Hockley BG, Scott PJH. An automated method for preparation of [18F]sodium fluoride for injection, USP to address the technetium-99m isotope shortage. Appl Radiat Isot. 2010;68:117–119. doi: 10.1016/j.apradiso.2009.08.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Jacobson MS, Dankwart HR, Mahoney DW. Radiolysis of 2-[18F]fluoro-2-deoxy-D-glucose and the role of ethanol and radioactive concentration. Appl Radiat Isot. 2009;67:990–995. doi: 10.1016/j.apradiso.2009.01.005. [DOI] [PubMed] [Google Scholar]
- 36.McBride WJ, D’Souza CA, Cardillo TM, Goldenberg DM. A facile kit for rapid radiofluorination of peptides for PET. J Nucl Med. 2011;52(Supplement 1):313. [Google Scholar]
- 37.Füchtner F, Preusche S, Mäding P, Zessin J, Steinbach J. Factors affecting the specific activity of [18F]fluoride from a [18O]water target. Nuklearmedizin. 2008;47:116–119. [PubMed] [Google Scholar]
- 38.Lasne MC, Perrio C, Rouden J, Barre L, Roeda D, Dolle F, Crouzel C. Chemistry of β+-emitting compounds based on fluorine-18. Top Curr Chem. 2002;222:201–258. [Google Scholar]
- 39.Yao C-H, Lin K-J, Weng C-C, Hsiao I-T, Ting Y-S, Yen T-C, Jan T-R, Skovronsky D, Kung M-P, Wey S-P. GMP-compliant automated synthesis of [18F]AV-45 (Florbetapir F-18) for imaging β-amyloid plaques in human brain. Appl Radiat Isot. 2010;68:2293–2297. doi: 10.1016/j.apradiso.2010.07.001. [DOI] [PubMed] [Google Scholar]
- 40.Saiki H, Iwata R, Nakanishi H, Wong R, Ishikawa Y, Furumoto S, Yamahara R, Sakamoto K, Ozeki E. Electrochemical concentration of no-carrier-added [18F]fluoride from [18O]water in a disposable microfluidic cell for radiosynthesis of 18F-labeled radiopharmaceuticals. Appl Radiat Isot. 2010;68:1703–1708. doi: 10.1016/j.apradiso.2010.02.005. [DOI] [PubMed] [Google Scholar]
- 41.Pascali G, Mazzone G, Saccomanni G, Manera C, Salvadori PA. Microfluidic approach for fast labeling optimization and dose-on-demand implementation. Nucl Med Biol. 2010;37:547–555. doi: 10.1016/j.nucmedbio.2010.03.006. [DOI] [PubMed] [Google Scholar]
- 42.D’Souza CA, McBride WJ, Todaro LJ, Sharkey RM, Karacay H, Rossi EA, Chang C-H, Goldenberg DM. Evaluation of 1,4,7-triazacyclononane-1,4-diacetate (NODA) ligand for development of PET imaging agents. J Nucl Med. 2011;52(Supplement 1):171. [Google Scholar]
- 43.McBride WJ, D’Souza CA, Sharkey RM, Goldenberg DM. A rapid method for F-18 labeling of proteins for PET. J Nucl Med. 2011;52(Supplement 1):170. [Google Scholar]
- 44.Sheldrick GM. SHELXS-97 and SHELXL-97. University of Göttingen; Göttingen, Germany: 1997. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.








