Skip to main content

State of the art procedures towards reactive [18F]fluoride in PET tracer synthesis

Abstract

Background

Positron emission tomography (PET) is a powerful, non-invasive preclinical and clinical nuclear imaging technique used in disease diagnosis and therapy assessment. Fluorine-18 is the predominant radionuclide used for PET tracer synthesis. An impressive variety of new ‘late-stage’ radiolabeling methodologies for the preparation of 18F-labeled tracers has appeared in order to improve the efficiency of the labeling reaction.

Main body

Despite these developments, one outstanding challenge into the early key steps of the process remains: the preparation of reactive [18F]fluoride from oxygen-18 enriched water ([18O]H2O). In the last decade, significant changes into the trapping, elution and drying stages have been introduced. This review provides an overview of the strategies and recent developments in the production of reactive [18F]fluoride and its use for radiolabeling.

Conclusion

Improved, modified or even completely new fluorine-18 work-up procedures have been developed in the last decade with widespread use in base-sensitive nucleophilic 18F-fluorination reactions. The many promising developments may lead to a few standardized drying methodologies for the routine production of a broad scale of PET tracers.

Background

Molecular imaging is the in vivo visualization of biological processes at the molecular level via the interaction of a probe with its target (Gu et al. 2011). Within the field of molecular imaging, positron emission tomography (PET) is a powerful, non-invasive preclinical and clinical nuclear imaging modality. The use of PET requires administration of a tracer radiolabeled with usually short-lived positron-emitting nuclides such as 15O, 13 N, 11 C, 18F, 64Cu, 68Ga or 89Zr. Although more positron emitting radionuclides find their way into PET applications, fluorine-18 remains the most frequently used radionuclide due to its favorable chemical and physical properties. A convenient half-life of 109.7 min, a clean decay profile (97% positron emission) and a low positron energy (maximum 0.635 MeV) resulting in high-resolution PET images (Van Der Born et al. 2017).

Several radiosynthetic methods are currently available for incorporation of fluorine-18 into biologically active compounds. In recent years, the unique and distinctive radiochemistry has been excellently reviewed (Van Der Born et al. 2017; Deng et al. 2019; Preshlock et al. 2016b). Although many different techniques have been implemented to improve the efficiency of the labeling reaction, only a few have dealt with the early key steps of the process, which consist of the removal of [18F]fluoride from its oxygen-18 enriched water ([18O]H2O) matrix and the generation of its reactive form. The landmark paper by Hamacher et al., describing a high yielding synthesis of 2-[18F]fluoro-2-deoxy-D-glucose ([18F]FDG), more or less defined the conventional route of producing reactive [18F]fluoride (Hamacher et al. 1986). This method starts with the production of fluorine-18 by proton irradiation of [18O]H2O in a cyclotron via the 18O(p,n)18F nuclear reaction. The radionuclide is obtained as strongly hydrated [18F]fluoride ions typically ranging from 50 to 200 GBq and with high molar activity (up to 1 TBq/µmol) (Aerts et al. 2010; Brichard and Aigbirhio 2014; Pees et al. 2018; Wessmann et al. 2012). The [18F]fluoride is trapped on an anion-exchange resin, typically a Waters Sep-Pak QMA, allowing the removal and recovery of expensive target [18O]H2O. The radioactivity is then eluted with a ~ 10% aqueous acetonitrile (MeCN) solution containing an inorganic base and cryptand, typically K2CO3/K222 (20 : 40 µmol). Subsequent drying of the eluate by repeated azeotropic distillation with MeCN provides the [18F]fluoride starting material. Finally, the residual anhydrous [18F]KF/K222 complex is taken up in a polar aprotic solvent suitable for the subsequent labeling reaction.

Although optimized, the overall procedure is time-consuming in the context of 18F-radiochemistry (around 10–15 min) and can lead to losses of significant amounts of radioactivity (up to 30%) due to unspecific adsorption on the reactor surface during azeotropic drying (Brichard and Aigbirhio 2014; Pees et al. 2018; Wessmann et al. 2012). This procedure is also complex to automate and miniaturize, e.g. for the use in microfluidic devices, and it is hard to control the precise hydration state of fluoride leading to irreproducible results. Most importantly, substantial amounts of base are needed to isolate [18F]fluoride from the cartridge thereby limiting its synthetic utility in the subsequent radiofluorination. The principles of the conventional work-up still remain in modern radiofluorination methods, but significant changes into especially the elution step, but also the trapping and drying stages have been introduced in the last decade (Fig. 1). In this review, we provide an overview of the strategies and recent developments in the production of reactive [18F]fluoride and their applications in PET tracer synthesis in the last decade. To demonstrate that these isolation and activation protocols are indeed valuable assets, the technical ease, efficiency (radiochemical yield (RCY), activity yield (AY), radiochemical conversion (RCC), molar activity (Am) and time will be discussed in each case (Coenen et al. 2017; Herth et al. 2021). Where applicable, the proof-of-concept application and routine production of representative and clinically relevant 18F-labelled radiopharmaceuticals under conditions of good manufacturing practice (GMP) are described.

Fig. 1
figure 1

Overview of the conventional route for the purification and drying of aqueous [18F]fluoride with an anion-exchange cartridge

Main text

Trapping on anion-exchange cartridges

During trapping, the negatively-charged [18F]fluoride ions are attracted to a positively-charged resin and exchange with the negatively charged ions to adsorb to the surface while the [18O]water continues to the outlet. Depending on the functional groups bound to the resin, the cartridges can be strongly or weakly basic. In addition, the negatively charged ions used in the preconditioning step are a major contributor to the basicity of the reaction mixture after elution. Hence, the type and size of anion-exchange resin as well as the preconditioning agent can substantially influence the work-up procedure of fluorine-18 with regard to azeotropic drying, losses of significant amounts of [18F]F on the reactor surfaces or the amount of base and phase transfer catalyst needed to elute [18F]F.

Solid support of the anion-exchange cartridge

Generally, [18F]F is extracted from the [18O]water on a polymer resin impregnated with quaternary methylammonium salts (QMA). Due to their strong anion exchange capacity, several weaker quaternary and tertiary ammonium salts have been developed to reduce the amount of base and cryptand required to elute the [18F]F and to perform on-resin radiolabeling reactions. Alternatively, main group elements, such as boranes and phosphazenes, have been studied in the design of cartridge resins for the preparation of dry [18F]fluoride in the absence of an azeotropic drying step.

Quaternary ammonium salts

Resins based on quaternary ammonium salts were investigated with a custom-synthesized 4-(N,N-dialkylamino)-pyridinium functionalized anion-exchange resin for the collection, drying and on-column 18F-radiolabeling reaction (Mulholland et al. 1989; Toorongian et al. 1990). Problems such as solubility of counter anions and removal of phase transfer catalyst reagents were eliminated, resulting in the radiosynthesis of [18F]FDG with an average AY of 41 ± 15.6% (n = 104), which is comparable to the classical method (44 ± 4% AY, n = 7). In addition, Aerts et al. described the use of long alkyl chain quaternary ammonium salts as resin to quantitatively trap [18F]fluoride and recover > 85% of the trapped activity with 1 mL of a non-protic solvent (Aerts et al. 2010). However, the co-eluted concentration of water is higher than that obtained after classical azeotropic drying of a K2CO3/K222 solution, which might cause problems in water-sensitive reactions such as Cu-catalyzed radiofluorinations. Furthermore, the quantitative recovery of [18F]F (98.5 ± 1.0%, n = 23) with just 2 µmol of K222/KHCO3 in MeOH has been described using a commercially available, mixed-mode quaternary ammonium salt resin (MAX, Oasis) (Fig. 2A) (Iwata et al. 2017). The amount of K2CO3/K222 could be even further reduced by a subsequent cation-exchange cartridge (Oasis MCX) (Iwata et al. 2018). This procedure enabled the microscale radiosynthesis several clinically relevant precursors in RCC’s comparable to those reported in the literature. Importantly, since the solvent scale is small, poor reproducibility of the radiosyntheses was obtained.

Tertiary ammonium salts

Instead of quaternary amines, a commercially available piperazine based anion-exchange resin (Oasis WAX) has been studied for trapping of the aqueous [18F]F and its release by elution with a weak anion-exchanger, in this case pyridinium sulfonate (25 µmol) in DMA (78% elution efficiency) (Fig. 2B) (Antuganov et al. 2019). The reactive [18F]fluoride was then directly used in a Cu-catalyzed radiofluorination reaction obtaining high RCCs for a series of simple arylBPin substrates and a RCY of 35–38% (n = 2) for 4-[18F]FPhe within 90 min.

Phosphonium borane salts

[18F]Fluoride has been prepared according to a simple two-step sequence including trapping of aqueous [18F]F on a cartridge pre-loaded with the phosphonium borane [(Ph2MeP)C6H4(BMes2)]+, followed by quantitative release of [18F]F from the cartridge using a solution of tetrabutylammonium cyanide (TBACN) (one equivalent with respect to borane) in anhydrous MeCN (Fig. 2C) (Perrio et al. 2017). Loading of the cartridge with borane salt and using equimolar amounts of TBACN to replace [18F]F on the cartridge as well as to convert the excess borane salt to borane-CN are a prerequisite. Subsequent radiofluorination was successfully applied to model compounds and to the synthesis of [18F]setoperone (25% RCC) (Crouzel et al. 1988).

Phosphazene bases

Based on the work by Lemaire et al. on the radiofluorination with phosphazene bases, polymer supported phosphazene bases (P2tBu and P2PEG) have been used for the efficient extraction of fluorine-18 from [18O]H2O. Treatment of the resin with [18O]H2O forms the P2tBuH+OH ion-pair and [18F]F is trapped via the [18F]F/OH anion-exchange to form the P2tBuH+[18F]F adduct (Fig. 2D) (Lemaire et al. 2010; Mathiessen and Zhuravlev 2013). The resin trapped > 97% of [18F]F after which the solution of substrate was passed through the column effecting on-column radiofluorination for several model compounds. An advantage of this method compared to the ones mentioned above, is that the column can be reused without any loss in [18F]F trapping efficiency. However, repeated use is limited to 3 runs. [18F]FDG was produced with a RCY of 41% (n = 1) on a 120 GBq scale with an Am >11 GBq/µmol, which is comparable to the benchmark production of [18F]FDG (44 ± 4% AY, n = 7).

Fig. 2
figure 2

 A protected [18F]FDG with a mixed-mode quaternary ammonium salt resin (MAX, Oasis). B Radiosynthesis of 4-[18F]FPhe from with a tertiary ammonium based anion-exchange resin (Oasis WAX). C Radiosynthesis of [18F]Setoperone with a phosphonium borane [(Ph2MeP)C6H4(BMes2)]+ anion-exchange resin. D Radiosynthesis of [18F]FDG with a phosphazene base (P2tBu) supported anion-exchange resin; n.m.: not mentioned

Size of the anion-exchange cartridge

Reducing the size of the anion-exchange cartridge has also been considered as an alternative for reducing water and/or base content. In macro-scale reactions, this method has only been applied once in the radiosynthesis of [18F]fluoroazomycin arabinosid ([18F]FAZA) (Hayashi et al. 2011). A small home-made QMA cartridge containing 25 mg of the QMA resin, compared to the conventional 130 mg, resulted in a high RCY of 52.4 ± 5.3% (n = 8) with just 5.0/6.0 µmol of K2CO3/K222 within 50.5 ± 1.5 min.

Miniature anion-exchange cartridges

One of the first reports demonstrated concentration of [18F]fluoride (up to 10 GBq) in 1–5 mL of [18O]H2O with a custom-build miniature cartridge (~ 5 µL bed volume) packed with QMA resin (Bejot et al. 2011). Trapping and release of [18F]F with only 2.1 µmol K2CO3 in H2O followed by 4.8 µmol K222 in MeCN resulted in 95–97% recovery of [18F]F with a total processing time of ~ 11 min. Since then, commercial micro-cartridges (OptiLynx, 5 µL bed volume) have enabled the concentration of [18F]fluoride (up to 110 GBq) to a final volume of 5–45 µL (Chao et al. 2018; Elizarov et al. 2010; Lebedev et al. 2013). On average, > 99% of [18F]F is trapped and > 92% is subsequently released with a typical K2CO3/K222 mixture (0.50/0.60 µmol) in ~ 3 min. Noteworthy, the OptiLynx cartridge filled with Chromabond PS-HCO3 resin has been successfully used in the preparation of human doses of 5-[18F]fluorouracil (Hoover et al. 2016). Efficient elution of [18F]F (81%) was obtained with a MeCN/H2O mixture of K3PO4/18-crown-6 (0.5/3.0 µmol). However, a major drawback of these procedures is the inconvenient manual packing of resin in the cartridge and the lack of reproducibility that this operation implies.

Miniature anion-exchange particles

For the integration of anion-exchange resins on the microfluidic device, several techniques have been reported. First, solution suspended anion-exchange beads have been reported by Lee et al. (2005). After trapping of fluorine-18 on a chip (26 MBq in 1 µL [18O]H2O), 18 MBq was recovered by elution with 5 nmol of aqueous K2CO3 solution and an azeotropic drying step (69% efficiency). The method provided 4 MBq of [18F]FDG with a RCY of 38% within 14 min. In addition, Leonardis et al. manually loaded recyclable anion-exchange resin particles (Chromabond PS-HCO3) into on-chip cavities (Leonardis et al. 2011). The trapping efficiency of 5–7 GBq [18F]fluoride in 4 mL [18O]H2O was > 90% and ≥ 95% of the trapped radioactivity was eluted with 9 µmol K2CO3 in 250 µL MeCN/H2O within 6 min. Furthermore, concentration of [18F]fluoride has been demonstrated with 10–15 mg embedded QMA resin (Sep-Pak Accell Plus cartridge) (Salvador et al. 2017). For activities of 19 GBq, trapping of 2 mL of [18F]fluoride in [18O]H2O was achieved with 98% efficiency. The trapped activity could then be eluted into a volume of 20 µL K2CO3 (7.2 µmol) with > 87% recovery. Instead of filling on-chip cavities, an insert element pre-filled with QMA resin (Sep-Pak® Light QMA) has been integrated on-chip (Rensch et al. 2014). 243–394 MBq in 100–500 µL [18O]H2O was concentrated with ~ 90% efficiency, which is comparable to conventional cartridge performance.

Anion-exchange monolith

Integration of the cartridges or resin particles in the chip is typically a challenge. Therefore Ismail et al. used a functionalized polymer monolith (polystyrene imidazolium chloride) instead of packed resin particles (Ismail et al. 2014). Trapping of [18F]fluoride solutions (1.5–7.4 MBq) had an efficiency of 97 ± 4% (n = 39) and it was shown theoretically that higher activities could be trapped by extending the length of the monolith. Various eluents were tested, recovering at least > 85% of the activity in a volume of 200 µL.

Preconditioning of the anion-exchange cartridge

Preconditioning of the QMA cartridge has been shown to have a significant effect on the [18F]F recovery in certain cases. For example, when using an eluent of iodonium salt precursor in MeOH, the presence of HCO3 anions on the anion-exchange resin are key. The rapid and efficient exchange of the diaryl iodonium salts counteranions with hydrogen carbonate results in an elution efficiency of > 90%. In contrast, other resin anions such as Cl, BF4, CF3SO3 and CF3CO2 lowered the elution yields to 14–49% (Maisonial-besset et al. 2018; Pauton et al. 2019).

Oxalate and triflate counteranions (C2O4 2−, OTf)

For Cu-mediated 18F-radiolabeling reactions, it has been reported that a combination of appropriate preconditioning and concentration of base in the eluent can improve [18F]F recovery when eluting with weak bases (pKa < 7) as illustrated by [18F]Cabozantinib (Lien et al. 2018; Mossine et al. 2017). In a full-scale experiment using 15 GBq up to 40% of [18F]F was lost due to sticking to the reaction vessel or on the anion-exchange column resulting in only 0.7% RCY. Modifying the preconditioning (oxalate solution) and the elution (1.6/6.5/15 µmol of K2CO3/K2C2O4/K222) resulted in only 13% loss of activity, with the RCY increasing to 2.8 ± 0.05% (n = 4) and Am of 17 ± 8 GBq/µmol (Fig. 3A). In general, when using weakly basic eluents such as K2CO3/K2C2O4/K222 and K2CO3/KOTf, the anion-exchange cartridge should be preconditioned with the corresponding weak counteranions, i.e. oxalate and triflate, respectively (Cai et al. 2020; Guibbal et al. 2019; Li et al. 2019a; McCammant et al. 2017; Preshlock et al. 2016a; Wu et al. 2019). Also when using copper complexes as elution agents, preconditioning with aqueous LiOTf solution can result in higher [18F]F recovery, but strongly depends on the type of QMA cartridge used (Lahdenpohja et al. 2019).

Mesylate and phosphate counteranions (OMs, HPO4 , PO4 2−)

In aliphatic radiofluorination reactions, controlling the base amount has been achieved by changing the bicarbonate counteranions on the resin to the inert mesylate anion, using 0.2 M of KOMs solution in the preconditioning step (Lee et al. 2011b, 2012). [18F]F can then be eluted with only 20 µmol KOMs resulting an eluate without any basic components. The exact amount of base or wanted precursor/base ratio can then be directly added to the reaction vessel. For example, 18F-incorporation yields for 3′-deoxy-3′-[18F]fluorothymidine ([18F]FLT) synthesis were 46.3 ± 5.5% (n = 3) and 17.79% (n = 1) for KOMs and K2CO3 preconditioned resins, respectively (Lee et al. 2012). Remarkably, in the synthesis of N-[18F]fluoropropylcarbomethoxyiodophenyl-nortropane ([18F]FP-CIT) the amount of base during nucleophilic 18F-fluorination was finely controlled by adding 6 µmol tetrabutylammonium hydroxide (TBAOH) to the reactor directly (Lee et al. 2011b). Over 192 preparations of [18F]FP-CIT have been performed with a mean RCY of 42.5 ± 10.9% and Am of 64.4 ± 4.5 GBq/µmol (Fig. 3B). Furthermore, di- and tribasic phosphates have also been reported as counteranions in aliphatic fluorination reactions, but application to a pharmaceutically relevant compound has not been reported (Seo et al. 2011).

Fig. 3
figure 3

Radiosynthesis of [18F]Cabozantinib and [18F]FP-CIT after preconditioning the anion-exchange resin with an aqueous potassium oxalate (A) and potassium mesylate (B) solution, respectively

Elution of anion-exchange cartridges with non-conventional bases, cryptands, solvents and additives

After trapping, the [18F]fluoride is eluted from the cartridge. Over the years several methods have been examined to optimize, but especially simplify this process. Multiple mixtures of anions and cryptands of both high and low basicity have successfully been applied in the elution of [18F]F from an anion-exchange cartridge. In addition, solvents ranging from aqueous polar aprotic to polar protic and even charged ones have been used to significantly reduce the time of the subsequent azeotropic drying step. Moreover, instead of bases and cryptands, the elution has been performed with precursors serving as the cation. Finally, ways of making the radiolabeling reactions tolerant to water have been given attention in order to omit a drying step.

Elution with weakly basic potassium salts and cryptand in MeCN/H2O

Typically [18F]fluoride is extracted from QMA resins by elution with a ~ 10% aqueous MeCN solution containing relatively large amounts of potassium salts and cryptand, e.g. K2CO3/K222 (20:40 µmol). However, also weak bases and neutral phase transfer catalysts can be used for this purpose. So far, these eluents have been tailored to meet the specific needs of a subsequent reaction of interest.

Potassium carbonate/Kryptofix-222 (K2CO3/K222)

A very straightforward method for decreasing the amount of K2CO3/K222 is to omit the trapping/elution step of [18F]F and instead to directly evaporate the target enriched water. In the preparation of (S)-N-[(1-allyl-2-pyrrolidinyl)methyl]-5-(3-[18F]fluoropropyl)-2,3-dimethoxybenzamide ([18F]Fallypride) the amount of K2CO3/K222 was only 2.5–5 µmol by evaporating the [18O]H2O together with K222 and K2CO3 (Mukherjee et al. 1995). However, evaporation of a large volume of [18O]H2O is time-consuming and expensive. As an alternative, decreasing the amount of K2CO3 but maintaining the amount of K222 compared to the classical conditions, sometimes combined with using a larger amount of water, has proven to be successful in the radiolabeling of 2-nitropurine-based nucleosides, e.g. 2-[18F]Fludarabine (Guillouet et al. 2014; Marchand et al. 2010, 2016). Similarly, the high-base labeling conditions can be adapted by eluting [18F]F with a small aliquot of the conventional K2CO3/K222 solution (Zlatopolskiy et al. 2015). For example, a solution of K2CO3 (0.87 µmol) and K222 (1.5 µmol) led to efficient elution of [18F]F from the QMA cartridge (82 ± 2%, n = 27), which resulted in the production of [18F]FAMTO in a RCY of 18 ± 2% with an Am of 105 ± 39 GBq/µmol from EOB within 120 min (Fig. 5A) (Bongarzone et al. 2019). In contrast, the formation of [18F]FAMTO under the classical conditions (K2CO3/K222 12/20 µmol) was not observed.

Potassium carbonate/potassium oxalate/Kryptofix-222 (K2CO3/K2C2O4/K222)

Soon after the landmark paper by Hamacher et al., a system with potassium oxalate (K2C2O4) as base and minimal amounts of K2CO3 (≤ 0.36 µmol) was reported for the radiofluorination of butyrophenone neuroleptics (Hamacher and Hamkens 1995; Katsifis et al. 1993). The conventional strong basic K2CO3/K222 system led to degradation of the tosylated precursors, but with the alternative system 18F-incorporation was facilitated. This moderately basic eluent has been used in the nucleophilic substitution of several aromatic scaffolds including those of clinically relevant radiotracers such as [18F]-3-fluoro-5-[(pyridin-3-yl)ethynyl]benzonitrile ([18F]FPEB), [18F]Flumazenil ([18F]FMZ), meta-[18F]fluorobenzylguanidine ([18F]mFBG) and L-3-[18F]fluoro-α-methyltyrosine ([18F]FMT) (Hamill et al. 2005, 2011; Hostetler et al. 2011a, b; Labas et al. 2011). Following the K2C2O4/K2CO3 elution protocol, Preshlock et al. reported the successful Cu-mediated nucleophilic 18F-fluorination of clinical PET tracers from their arylboronic ester precursors. Using this protocol, degradation of the base-sensitive copper catalyst was prevented resulting in isolated RCY’s up to 39% and Am >100 GBq/µmol (Preshlock et al. 2016a). Exemplary is the clinical production of 6-[18F]fluoro-L-dihydroxy-phenylalanine ([18F]FDOPA). By using K2CO3/K2C2O4/K222 (0.72/6.0/17 µmol), a dose of 2.18 GBq of [18F]FDOPA could be isolated corresponding to 40 ± 4% RCY (n = 5), while classical elution of the QMA cartridge with K2CO3 (14 µmol) and K222 (24 µmol) gave access to a clinical dose of [18F]FDOPA in 12% RCY (n = 1) (Fig. 4) (Tredwell et al. 2014). Since then, several examples of the radiofluorination of aryl boronic ester substrates with K2CO3/K2C2O4/K222 as eluent for the QMA cartridge have been reported (Ahmed et al. 2019; Bratteby et al. 2020; Clemente et al. 2019; Guibbal et al. 2020; Haider et al. 2019; Prause et al. 2018). Besides preventing degradation of base-sensitive precursors and catalysts, K2C2O4 has proven to be of aid in reducing racemization in substrates with a chiral center. The initial radiosynthesis of L-S-(3-[18F]fluoropropyl)homocysteine ([18F]-L-FPHCys) under conventional radiolabeling conditions resulted in a racemization of 25% towards [18F]-D-FPHCys due to proton abstraction of the R-hydrogen (Bourdier et al. 2011b). Under identical reaction conditions except for the use of non-nucleophilic K2CO3/K2CO4 (0.36/16 µmol) instead of K2CO3 (14 µmol), [18F]-L-FPHCys was obtained from the tosylate precursor without racemization in a AY of 20 ± 5% and radiochemical and enantiomeric purity of > 98% in 65 min.

Potassium carbonate/potassium triflate (K2CO3/KOTf)

To limit the quantities of strongly basic K2CO3 in Cu-mediated radiofluorination reactions of boron precursors, Mossine et al. examined the elution of [18F]F with a weakly basic combination of KOTf and K2CO3 (27/0.36 µmol) (Mossine et al. 2015). With this new eluent, 80% of the [18F]F was recovered from the QMA and no loss of radioactivity during the azeotropic drying step was observed. Noteworthy, these weakly basic [18F]F elution conditions have been used in the preparation of PET tracers that could not be obtained with conventional methods on a preparative scale from both boron and stannane precursors (Elie et al. 2019; Makaravage et al. 2016; Wu et al. 2019). For example, radiotracers for the in vivo imaging of TrkA/B/C receptors and oncogenic Trk fusion proteins 2019; Li et al. 2019a).

Fig. 4
figure 4

Radiosynthesis of [18F]FDOPA after elution of [18F]fluoride with a moderately basic K2CO3/K2C2O4/K222 eluent (A) and the conventional K2CO3/K222 solution (B)

Potassium dihydrogen phosphate/Kryptofix-222 (KH2PO4/K222)

Although not common, a solution of KH2PO4 (29 µmol) with K222 (40 µmol) has been used for elution of [18F]F in the automated synthesis of [18F]FMZ from its boronic ester precursor (Fig. 5C) (Preshlock et al. 2016a). [18F]FMZ was isolated in 16 ± 4% RCY (n = 10) comparable to the 19% RCY (n = 1) obtained by using K2C2O4/K2CO3 (0.72/6.0 µmol) and K222 (17 µmol) as eluent.

Potassium bicarbonate/18-crown-6 (KHCO3/18-Crown-6)

In specific cases, not only changing to a mild basic agent is necessary, but also to a neutral phase transfer catalyst. In the radiolabeling reaction of 4-[18F]fluoroglutamine all four possible isomers and a significant amount of elimination byproducts were formed in the presence of K222 and K2CO3 (Qu et al. 2011). The less basic KHCO3/18-Crown-6 (13/28 µmol) catalyst system led to the successful automated preparation of the desired L-glutamine isomer (2 S,4R)-4-[18F]fluoroglutamine (4-[18F]FGln) in 8.4 ± 3.4% AY (n = 10) and high stereoisomeric purity (> 91 ± 8%) (Fig. 5D) (Lieberman et al. 2011).

Elution with tetraalkylammonium salts as base and cryptand in MeCN/H2O

Tetraalkylammonium salts have an enhanced solubility in organic solvents and are therefore widely used as an alternative to the K2CO3/K222 system. Generally, an ion exchange between the [18F]F and the counter-anion of the ammonium salt occurs during elution of the activity from the QMA resin with a tetraalkylammonium salts solution. The tetraalkylammonium/[18F]F complex serves then as the phase transfer catalyst. The different alternatives will hereafter be discussed.

Tetrabutylammonium hydroxide (TBAOH)

Early on, Parent et al. reported the labeling of N-(3-fluoro-4-nitronaphthyl)-cis-5-norbornene-endo-2,3-dicarboxylic imide (3-[18F]F-NNDI) through an extremely mild, SNAr displacement reaction of an o-nitro-activated arene trimethylammonium salt with fluorine-18 by using TBAOH (2 µL) (Parent et al. 2006). This procedure allowed the preparation of 3-[18F]F-NNDI in 70–81% RCY.

Fig. 5
figure 5

 A Radiosynthesis of [18F]FAMTO after elution of [18F]fluoride with an aliquot of the conventional K2CO3/K222 solution. B Radiosynthesis of [18F]SDM-8 after elution of [18F]fluoride with the weakly basic K2CO3/KOTf eluent. C Radiosynthesis of [18F]FMZ after elution of [18F]fluoride with the moderately basic KH2PO4/K222 eluent. D Radiosynthesis of [18F]FGln after elution of [18F]fluoride with the moderately basic KHCO3/18-Crown-6 eluent; n.m.: not mentioned

Tetrabutylammonium bicarbonate (TBAHCO3)

Currently, TBAHCO3 is the most commonly used ammonium salt for the preparation of [18F]TBAF. It has proven its value in both base-sensitive aliphatic and aromatic nucleophilic substitution reactions. For example, TBAHCO3 (8.6 µmol) increased the RCY of 16α-[18F]fluoroestradiol ([18F]FES) from the cyclic sulfonate precursor up to 32 ± 8% compared to a RCY of 15 ± 2% with K2CO3/K222 (13/26 µmol) as phase transfer catalyst (Fig. 6) (Fedorova et al. 2018; Knott et al. 2011; Wang et al. 2020b). Furthermore, Chang et al. showed that the synthesis of 3-[18F]fluoro-1-(2′-nitro-1′-imidazolyl)-2-propanol ([18F]FMISO) using [18F]TBAF instead of [18F]KF/K222 in the substitution of the tosyl moiety increased the RCY from 15 ± 3% to 30 ± 5% (AY, n = 75) (Chang et al. 2007; Cherif et al. 1994). In addition, TBAHCO3 (30 µmol) provided equally good RCYs as the classical K2CO3/K222 system (51/58 µmol) (48 ± 3% and 54.4 ± 2.9% RCY (n = 3), respectively) in the [18F]FMISO synthesis with the THHP protecting group (Nandy et al. 2010; Oh et al. 2005). In the case of small peptidic radiotracers, the use of [18F]TBAF results in clean reaction profiles with high yields, while classical bases such as K2CO3 and KOAc typically result in lower yields. Interestingly, Dornan et al. reported the efficient production of 2-(3‐{1‐carboxy‐5‐[(6‐[18F]fluoro‐pyridine‐3‐carbonyl)‐amino]‐pentyl}‐ureido)‐pentanedioic acid ([18F]DCFPyL) using a direct, single‐step radiofluorination reaction without the use of protecting groups in the presence of [18F]TBAF (Fig. 7A) (Dornan et al. 2018). This simplified approach enabled automated preparation of [18F]DCFPyL in 26 ± 6% AY (n = 7) and with high Am (1403 ± 153 GBq/µmol) within 28 min. Presently, [18F]TBAF has allowed the preparation of clinically-relevant radiotracers, via fully-automated procedures including O-(2-[18F]fluoroethyl)-L-tyrosine ([18F]FET), 16β-[18F]fluoro-5α-dihydrotestosteron ([18F]FDHT), 2′-deoxy-2′-[18F]fluoro-5-methyl-1-β-D-arabinofuranosyluracil ([18F]FMAU), prostate‐specific membrane antigen ([18F]PSMA-1007) and [18F]DCFPyl (Bouvet et al. 2016; Cardinale et al. 2017; Lakshminarayanan et al. 2017; Lazari et al. 2015; Li et al. 2011; Ravert et al. 2016; Shamni et al. 2019).

Fig. 6
figure 6

Radiosynthesis of [18F]FES after elution of [18F]fluoride with a TBAHCO3 solution (A) and the conventional K2CO3/K222 solution (B)

Tetrabutylammonium methanesulfonate (TBAOMs)

The use of TBAOMs (36–40 µmol) as very mild base for the generation of [18F]TBAF has significantly increased in recent years due to the development of iodonium containing precursors i.e. diaryliodonium salts and spirocyclic iodonium ylides (SCIDY) (Vavere et al. 2018). These precursors are generally unstable under basic conditions and undergo radical-induced decomposition under the conventional radiolabeling conditions with [18F]KF/K222 (Lee et al. 2011a; Shi et al. 2016). For example, in the radiolabeling of the TSPO PET ligand N,N-diethyl-2-(2-(4-[18F]fluorophenyl)-5,7-dimethylpyrazolo[1,5- a]pyrimidine-3-yl)-acetamide ([18F]FDPA) only trace amounts of product were detected in the presence of K2CO3/K222. In contrast, TBAOMs (36 µmol) as eluent provided > 95% [18F]F elution efficiency and 4.2% AY (n = 3) and Am of 529 ± 23 GBq/µmol within 60 min (Fig. 7B) (Wang et al. 2017, 2020a). TBAOMs as phase transfer catalyst has enabled the efficient and automated production of several PET radiotracers from their SCIDY precursors, e.g. N-(2,5-dimethoxybenzyl)-N-(5-[18F]fluoro-2- phenoxyphenyl)acetamide ([18F]DAA1106) (Fujinaga et al. 2018; Kumata et al. 2018).

Tetrabutylammonium triflate (TBAOTf)

The [18F]KF generated from the weakly basic K2CO3/KOTf eluent causes side reactions e.g. protodeborylation and/or hydroxydeborylation in Cu- mediated fluorination reactions of arylboronic pinacol esters due to poor dissolution (Mossine et al. 2018). To facilitate efficient elution from QMA resins and rapid dissolution of [18F]F, the weak nitrogenous base TBAOTf (13–19 µmol) has been used (Zhang et al. 2017a, b). Noteworthy, the only validated production for human use of [18F]FDOPA from [18F]F uses an aqueous eluent consisting of TBAOTf and Cs2CO3 (19/0.30 µmol) as a replacement for KOTf and K2CO3, respectively (Fig. 7C). Although, the AY and Am (6 ± 1% and 141 ± 77 GBq/µmol) were comparable to the RCY and Am obtained under conventional conditions (12% and 112 GBq/µmol), the chemical purity of [18F]FDOPA was improved, qualifying it for clinical use (Mossine et al. 2019).

Tetraethylammonium bicarbonate (TEAB)

As [18F]TBAF, [18F]TEAF has been discovered as an alternative to the [18F]KF/K222 system in the radiolabeling of diaryliodonium salts and spirocyclic iodonium ylides. For example, incorporation of [18F]fluoride into the 6,10-dioxaspiro[4.5]decane-7,9-dione (SPI5) spirocyclic iodonium(III) ylide precursor of [18F]LY2459989 was significantly more efficient with TEAB (10 µmol, 43% RCC) as eluent than with K2CO3/K222 (2.5/9.3 µmol, 14% RCC) (Cai et al. 2017) The simple combination of dried [18F]fluoride with TEAB (10–42 µmol) and the SPI5 iodonium ylide precursor in N,N-dimethylformamide (DMF) has been used to prepare several clinically relevant PET radiopharmaceuticals such as [18F]UCB-J and [18F]Lorlatinib (Collier et al. 2017; Li et al. 2019b; Nkepang et al. 2016). Noteworthy, radiolabeling with dried [18F]TEAF generated for the first time 5-[18F]fluorouracil from nucleophilic [18F]fluoride in 11 ± 4% AY (n = 3) with Am of 15 GBq/µmol as well as a ready to inject solution of [18F]FPEB in 20 ± 5% AY (n = 3) and with high Am (666 ± 52 GBq/µmol) (Fig. 7D) (Liang et al. 2019; Rotstein et al. 2014; Stephenson et al. 2015). In addition, radiofluorination with TEAB in DMF has also been successfully applied with the spiroadamantyl-1,3-dioxane-4,6-dione (SPIAd) precursor as illustrated by the radiolabeling of [18F]FMT (11 ± 1% AY, n = 3), [18F]mFBG (14 ± 1% AY, n = 3) and 4-[18F]fluoro‐m‐hydroxy-phenethylguanidine ([18F]4F-MHPG, 7.8 ± 1.4% AY, n = 8), from which the latter two could not be directly radiolabeled using conventional nucleophilic aromatic substitution with [18F]KF/K222 (Jung et al. 2019; Rotstein et al. 2016).

Fig. 7
figure 7

 A Radiosynthesis of [18F]DCFPyL after elution of [18F]fluoride with a TBAHCO3 solution. B Radiosynthesis of [18F]FDPA after elution of [18F]fluoride with a very mild basic TBAOMs solution. C Radiosynthesis of [18F]FDOPA after elution of [18F]fluoride with a weakly basic TBAOTf solution. D Radiosynthesis of [18F]Uracil after elution of [18F]fluoride with a TEAB solution

Weakly basic eluents by using polar protic solvents (R-OH)

Limited amounts of base are useful in the radiolabeling of base-sensitive precursors, but they do not always efficiently elute [18F]F from the anion-exchange cartridge. For example, 3.6/15 µmol K2CO3/K222 recovers only 18% of the radioactivity and 26 µmol TBAHCO3 elutes 74% of the [18F]F from the QMA resin (Moon et al. 2010). In order to find a proper balance between the selection of base and [18F]F elution efficiency, polar protic solvents have been studied extensively.

Potassium salts in methanol (MeOH)

As shown by Moon et al., a MeOH/H2O (5:1) solution with 5.8 µmol K2CO3 and 29 µmol K222 is able to quantitatively extract [18F]fluoride from an anion-exchange cartridge (Moon et al. 2010). This eluent has enabled the automatic production of [18F]Fallypride from its base-sensitive tosylate precursor in high RCY (71 ± 1.9%, n = 3) compared to the classical eluent with 11/15 µmol K2CO3/K222 in MeCN/H2O (26 ± 4.9% RCY, n = 3) (Fig. 8). Similarly, [18F]FMISO could be prepared in 25.1 ± 5.0% AY (n = 3) using only 14 µmol of the weak base KHCO3 in MeOH (Lee et al. 2013). Polar protic elution systems have also been applied in Cu-mediated radiofluorination reactions of boron pinacol esters and (mesityl)diaryl iodonium salts and in the radiolabeling of diaryliodonium salts (Tang et al. 2017; Zlatopolskiy et al. 2015). In the automated preparation of [18F]FMZ, starting from the base labile diaryliodonium salt precursor optimal yields (80.4 ± 3.2% RCC, n = 3) were achieved with only 4.9 µmol K2CO3 and 15 µmol K222, while increasing the amount of base to 8.1 µmol K2CO3 already resulted in a significant drop in RCC (42.2 ± 4.4%, n = 3) (Moon et al. 2011). In order to access [18F]FMZ suitable for human studies, the QMA was eluted with the required low amounts of K2CO3/K222 in MeOH, leading to a high RCY of 63.5 ± 3.2% (n = 26) and high Am of 370–450 GBq/µmol within 60.6 ± 1.1 min (Fig. 9A). In addition, automation of the radiosynthesis of 4-[18F]FGln, where the combination of KHCO3 and 18-Crown-6 (13/28 µmol) was proven to be important for the enantiomeric purity, resulted in superior elution efficiency (> 95%) when MeOH was used instead of MeCN/H2O (Qu et al. 2011; Zhang et al. 2016). This process was highly reproducible with regard to the elution efficiency and optical purity of the final product.

Fig. 8
figure 8

Radiosynthesis of [18F]Fallypride after elution of [18F]fluoride with a methanolic K2CO3/K222 solution (A) and the conventional K2CO3/K222 solution (B); n.m.: not mentioned

Tetrabutylammonium salts in MeOH

As with the K2CO3/K222 system in MeOH, limited amounts of TBAHCO3 (13 µmol) in MeOH are able to extract > 95% of the [18F]F from the QMA resin (Moon et al. 2010). This elution system also afforded [18F]Fallypride in a higher RCY of 71 ± 1.2% (n = 3) compared to the classical eluent with 39 µmol TBAHCO3 in MeCN/H2O (44 ± 4.1% RCY, n = 3). In addition, the formation of radiolabeled byproducts in the automated synthesis of [18F]FET from the Ni (II) complex of (S)-tyrosine Schiff base was observed in the presence of high amounts of TBAHCO3 (38 µmol), which was required for the quantitative elution with MeCN/H2O as solvent (Lakshminarayanan et al. 2017). When using a methanolic, or even ethanolic, solution of the neutral phase transfer catalyst TBAOTs (9.7 µmol), high elution efficiencies of [18F]F from the QMA cartridge were obtained (Fig. 9B). As a result, [18F]FET was prepared with Am of 9.6–38.4 GBq/µmol and AY of 40.6 ± 3.4% (n = 5) compared to 24.6 ± 2% AY (n = 5) in the presence of TBAHCO3 (Orlovskaya et al. 2019a). In addition, very mild conditions including preconditioning of the QMA with K3PO4 (5 mmol) followed by elution with TBAOMs or TBAOTf (20 µmol) in MeOH has made the aliphatic and Cu-mediated aromatic 18F-labeling of highly base sensitive monosubstituted tetrazines possible (Andersen et al. 2022; Bratteby et al. 2021). As such, 2-fluoro-4-(1,2,4,5-tetrazin-3-yl)-2-[18F]fluoroethyl-benzoate could be obtained in 5.2 ± 2.8% RCY and with Am of 96.3 ± 5.8 GBq/µmol after only 5 minutes at 100°C in tBuOH/MeCN (5:1). Moreover, the clinical lead compound 2,2′-(3-[18F]fluoro-5-(1,2,4,5-tetrazin-3-yl)benzyl)azanediyl)diacetic acid was produced in a RCY of 14 ± 3% (n = 3) and with Am of 201 ± 30 GBq/µmol from the stannane precursor.

Tetraethylammonium salts in MeOH

TEAB (2.4–4.2 µmol) in MeOH has often been used in the Cu-mediated radiofluorination reaction of pinacol esters (Bernard-Gauthier et al. 2018; Drerup et al. 2016; Singleton et al. 2020). In the case of 6-[18F]fluoro-L-tryptophan ([18F]FTrp), 1 µmol TEAB resulted in a RCC of 32 ± 8%. Since such a small amount of TEAB is inappropriate for large-scale productions, methanolic TEAB (4 µmol) has been used for eluting the [18F]F from the cartridge during automation. Following this procedure, 6-[18F]Trp could be synthesized with an overall AY of 15.8 ± 4% (n = 4) and a Am of 95–280 GBq/µmol within 110 min (Fig. 9C) (Schäfer et al. 2016; Zlatopolskiy et al. 2018). Furthermore, tetraethyl ammonium triflate (TEAOTf, 2.5 µmol) in MeOH has been successfully applied in radiofluorodestannylation reactions with the base-sensitive copper catalyst Cu(py)4(OTf)2. This eluent resulted in almost complete radioactivity recovery (95–97%) and provided several aromatic amino acids, e.g. [18F]FMT and [18F]FDOPA, in high AYs (32–54%) (Zarrad et al. 2017).

Tetraethylammonium salts in n-butanol (nBuOH)

To obviate an azeotropic drying or solvent-exchange step, the elution of [18F]F from the QMA cartridge using alcohols other than MeOH or ethanol (EtOH) have been investigated (Zischler et al. 2017). The radiolabeling of boronic and stannyl substrates often use a solution of TEAB (0.7–9.4 µmol) in nBuOH to elute [18F]F from the anion-exchange resin into a vial containing a solution of the precursor and copper catalyst in the desired solvent. Subsequently, the resulting solution is directly heated without a drying step. This general procedure has enabled the preparation of clinically relevant PET tracers including [18F]TRACK, [18F]Triacoxib, [18F]FDOPA and 6-[18F]fluorodopamine ([18F]FDA) (Bailey et al. 2019; Litchfield et al. 2020; Zischler et al. 2017). In the case of [18F]FDOPA, the unoptimized RCY (40 ± 4%, n = 1) was as high as the isolated RCY using the K2CO3/K2C2O4/K222 (0.72/6.0/17 µmol) eluent with an azeotropic drying step, showing the potential of this protocol (Fig. 9D) (Preshlock et al. 2016a; Zischler et al. 2017). In contrast, high Am [18F]TRACK (188 GBq/µmol) was obtained using TEAB in nBuOH, but in a AY of only 4% (n = 1), while a AY of 8% (n = 1) was achieved with a solution of TEAB in MeOH (Bailey et al. 2019; Bernard-Gauthier et al. 2018). In accordance with several studies, nBuOH enables the most efficient 18F-incorporation from the alcohols investigated, however typically radioactivity recovery is around 80% resulting in a decrease in RCY (Zarrad et al. 2017; Zischler et al. 2017; Zlatopolskiy et al. 2018). The addition of more base to the eluent could increase [18F]F recovery to > 80%, but this can also cause a drop in RCY.

Fig. 9
figure 9

 A Radiosynthesis of [18F]FMZ after elution of [18F]fluoride with a methanolic K2CO3/K222 solution. B Radiosynthesis of [18F]FET after elution of [18F]fluoride with the neutral phase transfer catalyst TBAOTs in ethanol. C Radiosynthesis of 6-[18F]Trp after elution of [18F]fluoride with alcoholic TEAB solutions. D Radiosynthesis of [18F]FDOPA after elution of [18F]fluoride with TEAB in nBuOH.

Elution with positively charged precursor molecules in alcoholic solvents: minimalist approach

Besides potassium and tetraalkylammonium salts, labeling precursors bearing a positively charged nitrogen, iodine, sulfur or oxygen functionality (onium salts) have been investigated for the quantitative recovery of [18F]F from an anion-exchange resin (Feni et al. 2017; Richarz et al. 2014). Solutions of onium salts in especially MeOH can directly extract [18F]F from a QMA resin in excellent yields of > 95%. Low-boiling methanol can subsequently be removed completely at 70–80 °C within 2–3 min without the need for azeotropic drying. The resulting onium [18F]fluoride salts are directly converted into 18F-labeled compounds without addition of a base or any other additives under so-called minimalist conditions.

Onium salts in MeOH

The first synthesis of 2-amino-5-[18F]fluoropyridines via palladium-catalyzed reaction of 2-bromo-5-[18F]fluoropyridine with several amines has been described using this approach (Pauton et al. 2019). The radiofluorination step of 2-bromo-5- [18F]fluoropyridine was performed under minimalist conditions to prevent interference of small amounts of base in the subsequent amination reaction. Especially, the application of the minimalist approach to Cu-mediated radiofluorination reactions has allowed the efficient preparation of several clinically relevant PET-tracers on a preparative scale regardless of their electronic properties, e.g. L-4-[18F]fluorophenylalanine ([18F]FPhe), [18F]DAA1106 and [18F]FDA (Zischler et al. 2016; Zlatopolskiy et al. 2015). The synthesis of [18F]FDA using the (4-anisyl)(aryl)iodonium precursor and [18F]KF/K222 in toluene was reported in 16.4 ± 3.0% AY (n = 4) with Am >9.25 GBq/µmol in 53 ± 4 min (Vavere et al. 2014). The minimalist conditions using the same iodonium salt precursor afforded a dose of 550 MBq of [18F]FDA (46% RCY, n = 1) as a ready-to-use solution within 130 min. Similarly, Zlatopolskiy et al. reported on the (MeCN)4CuOTf mediated radiofluorination of the (4-anisyl)(aryl)iodonium salt precursor under minimalist conditions resulting in [18F]FPhe in 53–66% RCY (n = 3) in two steps with an Am of 109 GBq/µmol compared to 20 ± 2% RCY (n = 2) with > 37 GBq/µmol Am using the conventional method (Neumann et al. 2013; Zlatopolskiy et al. 2015). Importantly, complete removal of MeOH is a prerequisite to obtain high RCYs. Therefore, extension of the evaporation time to 5 min instead of the usual 1–2 min is advised in Cu-mediated radiofluorination reactions of aryliodonium precursors (Zischler et al. 2016). The fully automated radiofluorination of [18F]FDOPA from its diaryliodonium salt precursor has been improved using the minimalist approach. Under conventional conditions (12 µmol K2CO3 and 25 µmol K222), [18F]FDOPA could be produced with an Am of 35 ± 4 GBq/µmol and a RCY of 14 ± 4% (n = 1) in 117 ± 4 min (Fig. 10) (Kuik et al. 2015). In contrast, quantitative elution of QMA-loaded [18F]F with the iodonium precursor (13 µmol) in MeOH (97 ± 0.03%), evaporation of MeOH and the addition of toluene yielded [18F]FDOPA in 27–38% RCY (n = 5) and high Am of 170–230 GBq/µmol in 64 min (Maisonial-besset et al. 2018). As with anilinium and iodonium salts, almost complete [18F]F recovery from the anion-exchange cartridge has been achieved with aryl sulfonium salts in MeOH (Richarz et al. 2014; Xu et al. 2020). Importantly, the elution efficiency of onium salts is correlated with the pKa value of the counteranion of the onium precursor; the stronger the basicity of the counteranion, the worse the elution of [18F]F (Maisonial-besset et al. 2018).

Fig. 10
figure 10

Radiosynthesis of [18F]FDOPA after elution of [18F]fluoride with the diaryliodonium precursor in an alcoholic solution (A) and the conventional K2CO3/K222 solution (B)

Onium salts in alcoholic polar aprotic solvent

The minimalist procedure has also been successfully adapted to the radiofluorination of onium salts in the absence of a solvent exchange step. In the case of uronium salts, the [18F]fluoride work-up procedure includes elution of [18F]F from the QMA with an uronium salt (11–21 µmol) in a mixture of EtOH and DMSO (Neumann et al. 2016). The elution procedure obviates the need for a solvent exchange step and subsequent heating of the resulting solution directly provides the product. However, the elution efficiency can be as low as 50% due to the low amounts of base and EtOH used. Addition of weak bases such as TBAOTf and TEAB to the eluent increased the elution efficiency, but the gain was offset by a decrease in RCY (Beyzavi et al. 2017). Under these adapted minimalist conditions, [18F]Bavarostat was produced under cGMP in a RCY of 14 ± 4% (n = 23) and an average Am of 204 ± 75 GBq/µmol (Fig. 11A) (Celen et al. 2020; Strebl et al. 2017). Recently, [18F]Atorvastatin has been prepared using the uronium precursor in MeOH/DMSO (1/3 v/v), which resulted in 88 ± 5% elution efficiency, a RCY of 19 ± 6% (n = 10) and an Am of 65 ± 32 GBq/µmol (Clemente et al. 2020). In addition, a balance between [18F]F recovery and 18F-incorporation has been achieved in the alcohol-enhanced Cu-mediated radiofluorination of (aryl)(mesityl)iodonium salts with a solution of the radiolabeling precursor (21 µmol) in 20% MeOH in DMF (Fig. 11B) (Orlovskaya et al. 2019b). A further increase of the MeOH content was detrimental and caused a rapid decrease of the 18F-incorporation yield due to the formation of 18F(MeOH)n clusters which substantially decrease the nucleophilicity of fluoride. Finally, 18F-labeled pyridines with electron‐withdrawing substituent(s) has been prepared by slow elution of [18F]F from the QMA cartridge with a solution of a quaternary ammonium triflate precursor in MeCN/tert-butanol (tBuOH) (1/4 v/v) (Basuli et al. 2018).

Fig. 11
figure 11

 A Radiosynthesis of [18F]Bavarostat after elution of [18F]fluoride with the uronium precursor in an alcoholic solution. B Radiosynthesis of 4-[18F]FPhe after elution of [18F]fluoride with the iodonium precursor in alcoholic polar aprotic solution

Elution with base in aqueous polar aprotic solvents to form in situ OH ions

Wessmann et al. and describes the use of hydroxide ions (OH) in anhydrous MeCN as the eluent in order to exchange with the [18F]F trapped on the resin (Wessmann et al. 2012). The major advantage of this method is omission of the time-consuming azeotropic evaporation of water. Although high 18F-incorporation yields (70–90%) were observed with fractions of the eluate of [K222]OH, using the entire eluate resulted in only incorporation (10–20%) caused by degradation of the precursors in the presence of large amounts of OH. This could be compensated by increasing the amount of precursor, but several alternative procedures have been developed using the OH/[18F]F exchange principle. These procedures use basic ligands or other non-ionic weak bases with a high conjugate pKa in the presence of trace amounts of water to generate OH in situ (Mossine et al. 2017).

Strong bases

The first report on the in situ OH production described the strong organic base P2Et (45 µmol, pKa=32.9) in aqueous MeCN, which resulted in nearly quantitative elution of [18F]F (Fig. 12A) (Lemaire et al. 2010). The elution volume was smaller with eluents containing larger amounts of water or bases with higher pKa values. Importantly, pKa values of around 30 were required for quantitative elution of the [18F]F from the cartridge. It still needs to be clarified whether a base strength of this magnitude is compatible with common precursors for nucleophilic 18F-fluorination.

Weak bases

Lindner et al. used the weak basic and nucleophilic K2HPO4 in aqueous protic tBuOH (1% of water) as eluent in the synthesis of [18F]FDG (Lindner et al. 2016). The use of tBuOH ensured high radioactivity recovery (86 ± 5%) and enabled nearly quantitative conversion (83 ± 8%). The excellent recovery and incorporation of [18F]F provided [18F]FDG in up to 59% RCY (n = 1), which is comparable to the AY of 44 ± 4% (n = 7) obtained by Hamacher et al. using conventional fluorination conditions, but significantly faster (20 and ~ 50 min, respectively) (Fig. 12B) (Hamacher et al. 1986; Lindner et al. 2016). Finally, in the Cu-mediated radiofluorination of arylboron reagents, solutions of 4-dimethylaminopyridinium trifluoromethanesulfonate (DMAP·OTf) in aqueous DMF or N,N-dimethylacetamide (DMA) have been discovered as useful eluents (Mossine et al. 2017; Zhang et al. 2019b). This approach does not require azeotropic drying and the absence of base increases the stability of both precursor and the copper reagent leading to better RCYs. This approach was successfully applied in the radiosynthesis of [18F]FMZ (Zhang et al. 2019b). Using a solution of DMAP·OTf (82 µmol) in DMA as eluent, [18F]FMZ was obtained in 26–47% RCY (n = 3) with an Am of 100–126 GBq/µmol (Fig. 12C).

Metal catalysts

[18F]Fluoride can also be eluted from the anion-exchange cartridge using an organic, aqueous solution of metal catalysts. In case of Cu-mediated radiofluorination reactions, Cu(OTf)2 (48 µmol), and to a lesser extent Cu(OTf)2(py)4, in DMA have been shown to be a suitable elution agent for [18F]F (79% elution efficiency) (Lahdenpohja et al. 2019; Mossine et al. 2017). Even though higher amounts of copper complex slightly improved the [18F]F recovery, large amounts of copper in the reaction solution complicated the purification of the final products on preparative scale. An example of this method has been given by Preshlock et al. in the automated production of [18F]FMZ with a AY of 16% (n = 1) using Cu(OTf)2(py)4 (40 µmol) in DMA as eluent (Fig. 12D) (Preshlock et al. 2016a). Also a solution of manganese catalysts (4–30 µmol) in pure acetone or a mixture with MeCN can directly elute [18F]F from an anion-exchange cartridge with > 90% efficiency (Huang et al. 2014; Liu et al. 2018). The eluate has successfully been used in the late-stage 18F-labeling of several unactivated aliphatic C − H bonds.

Fig. 12
figure 12

Radiosynthesis of [18F]FDG after elution of [18F]fluoride with the strong basic P2Et in aqueous MeCN (A) and the weak basic K2HPO4 in aqueous tBuOH (b); radiosynthesis of [18F]FMZ after elution of [18F]fluoride with the weak basic DMAP·OTf in aqueous DMF (C) and the metal catalyst Cu(OTf)2(py)4 in aqueous MeCN (D); n.m.: not mentioned

Direct addition of [18F]fluoride in [18O]H2O to reaction medium with additives

An alternative to the radiofluorination on a solid-phase without a need of azeotropic drying is the use of additives in the reaction medium. These additives allow the radiolabeling reaction to take place in the presence of a significant amount of water. To this end, highly viscous solvents, metal catalysts and enzymes have been studied as additives.

Ionic liquids

The rapid and convenient synthesis of 18F-fluoroalkanes has been enabled by the use of ionic liquids as reaction media, allowing the direct addition of the [18O]H2O target water containing [18F]fluoride and obviating the time-consuming azeotropic drying step (Kim et al. 2003, 2004). Generally, ionic liquids contain a lipophilic cation moiety based on imidazolium salts e.g. [bmim] and a counter anion that does not possess an exchangeable fluorine to prevent exchange with the [18F]fluoride ion. However, large amounts of starting activity are not possible, because excessive amounts of H2O in the reaction mixture reduce the yield of the 18F-fluorinated product. Besides a successful application in the synthesis of [18F]FDG, RCY of 59.1 ± 5.1% (dc, n = 3) compared to 44 ± 4% AY (n = 7) with the conventional method, no clinically relevant compounds or drug scaffolds have been shown (Fig. 13A) (Hamacher et al. 1986; Kim et al. 2003, 2004). A severe disadvantage of this approach is the high viscosity of these solvents, which limits their wider application.

Catalysts

Sergeev et al. reported the radiofluorination in up to 25 vol% of water by using titanium dioxide (TiO2) nanoparticles as catalyst (Sergeev et al. 2015). The TiO2 catalyst coordinates both aqueous [18F]fluoride and tosyl precursors to its surface via hydrogen bonding and thereby facilitates the reaction. In the presence of TBAHCO3 (0.36 µmol) as a phase- transfer agent, efficient radiolabeling of aromatic, aliphatic, and cycloaliphatic tosylated precursors has been performed without the need for a drying step. For example, the RCY and Am for TiO2 catalyzed synthesis of [18F]Fallypride were found to be 71.0 ± 2.3% (n = 3) and 185 ± 74 GBq/µmol, respectively (Fig. 13B). This is similar to those typically obtained with conventional procedures, e.g. 66 ± 8% RCY (n = 8) and Am of 15–78 GBq/µmol with 15 µmol TBAHCO3 and azeotropic drying (Lazari et al. 2014). However, increasing the amount of radioactivity in the reaction also increases the volume of water, which consequently requires a [18F]fluoride concentration step. In addition, triflate, nosylate and precursors with additional oxo moieties are not compatible with this method. Finally, enzymatic approaches offer a unique, mild, and selective method for the incorporation of fluorine-18 into substrates modified both on the adenine base and on the ribose ring by direct addition of [18F]F in [18O]H2O. The fluorinase enzyme from Streptomyces Cattleya has been successfully applied as catalyst in the 18F-labeling of nucleosides, e.g. [18F]-5′-fluoro-5′-deoxyadenosine ([18F]-5′-FDA), [18F]-5′-fluoro-5′-deoxyinosine and [18F]-5-fluoro-5-deoxyribose (Deng et al. 2006; Martarello et al. 2003; Winkler et al. 2008). However, the fluorinase is a relatively specific enzyme, it can never be a general fluorination reagent, and large amounts of the enzyme are required which makes purification of the product rather difficult.

Fig. 13
figure 13

 A Radiosynthesis of [18F]FDG with the ionic liquid [bmin][OTf] as additive. B Radiosynthesis of [18F]Fallypride with TiO2 as catalyst; n.m.: not mentioned

Drying of reactive [18F]fluoride ion from [18O]H2O

A complete methodological alternative to avoid a solvent evaporation step or the drying on a stationary phase is the conversion of aqueous [18F]fluoride into gaseous 18F-synthons or deposing it in an electric field. In both methods the radionuclide can be easily dried, either via distillation and trapping in a solvent or on an activated resin or via releasing under a low electric field. Both methods make an azeotropic drying process obsolete.

Synthesis of reactive [18F]fluoride species via 18F-gaseous synthons – [18F]fluoride relay

A relatively new concept for the synthesis of reactive [18F]fluoride species is [18F]fluoride relay. In [18F]fluoride relay, fluorine-18 is first transferred from aqueous [18F]F to an easily purified 18F-gaseous reagent. An anhydrous [18F]fluoride species is subsequently released from this gas by a simple chemical or physical interaction.

Phosphazenium [18F]hydrofluoride (P2Et-[18F]HF)

Mathiessen et al. reported the use of P2Et-[18F]HF obtained by the acidification of [18O]H2O with 98% H2SO4, transfer of the formed [18F]HF by an argon flow and subsequent trapping on solid bound phosphazene base P2Et (Fig. 14A) (Mathiessen et al. 2011). Under optimal conditions, the [18F]HF transfer yield was 16–82%. Mannose triflate (52 µmol) was converted into [18F]FDG at 120 °C for 20 min without any loss of stereochemical integrity and a RCY of 82 ± 5% (n = 3). However, this method appears technologically demanding for automation.

[18F]Acetyl fluoride ([18F]AcF)

[18F]AcF as novel synthon for the transfer of [18F]fluoride through gas lines in an anhydrous form was produced by trapping [18F]F on an anion-exchange column (MP-64) at 60 °C followed by the reaction with acetic anhydride (Fig. 14B) (Jiang et al. 2015). The resultant gaseous [18F]AcF was efficiently trapped in anhydrous apolar protic solvents after purification through a column containing Porapak Q resin and sodium sulfate at 30 °C in 87 ± 3% RCY (n = 10) in 15 min. The purified [18F]AcF has also been trapped on solid-phase extraction cartridges such as Oasis WAX, but limited information is available to date. Following dissolution of [18F]AcF in MeCN containing TEAB (52 µmol), 18F-fluorination of the mannose triflate provided [18F]FDG in > 90% incorporation rate. Except for a few model radiofluorination reactions, the scope of [18F]AcF has not been investigated to date.

[18F]Triflyl fluoride

Recently the preparation of [18F]triflyl fluoride has been discovered (Pees et al. 2018). The reaction of N,N-bis(trifluoromethylsulfonyl)aniline with aqueous [18F]F yields volatile [18F]triflyl fluoride at 40 °C (Fig. 14C). After distillation over a sicapent column, dry [18F]triflyl fluoride is released in the presence of base and K222 in efficiencies > 90% within 5 min. The type and amount of base are variable and thus can be adapted with regard to the subsequent radiofluorination reaction. This method was used to efficiently radiolabel aromatic and aliphatic compounds in moderated to high RCYs as well as the PET tracers [18F]FES and [18F]FET in good RCYs (17–57% and 10–55%, respectively, n = 2) and Am of up to 123 GBq/µmol. A comparable range in molar activity and RCYs was observed using the standard elution and drying procedure (Bourdier et al. 2011a; Fedorova et al. 2018). As an extension of this method, Dahl et al. showed the in-loop radiofluorination with [18F]triflyl fluoride as the labeling agent (Dahl et al. 2019). N-acetyl-N-(2-[18F]fluoroethoxybenzyl)-2-phenoxy-5-pyridinamine ([18F]FEPPA) and 7-(6-[18F]fluoropyridin-3-yl)-5 H-pyrido[4,3-β]indole ([18F]T807) were synthesized with AYs of 29% (n = 1) and 27% (n = 1), respectively, and Am >350 GBq/µmol for both radiotracers. The in‐loop radiofluorination methodology provided equal or superior yields compared with conventional reactions, e.g. the syntheses of [18F]T807 and [18F]FEPPA have previously been reported in AYs of 14 ± 3% (n = 3) and 38 ± 3% (n = 15), respectively (Berroteran-Infante et al. 2018; Shoup et al. 2013).

[18F]ethenesulfonyl fluoride ([18F]ESF)

[18F]ESF forms a simple way of activating [18F]fluoride and can conveniently be shipped for off-site use (Zhang et al. 2019a). [18F]ESF is produced from [18F]fluoride and 2,4,6-trichlorophenylethenesulfonate in the presence of TEAB (60 µmol) and subsequently distilled onto a silica-plus cartridge with an average RCY of 57 ± 11% (Fig. 14D). [18F]ESF was converted into a reactive [18F]fluoride species by elution with TEAB (12 µmol) in a chosen solvent. The RCCs using [18F]ESF method were comparable to the use of traditionally dried [18F]TEAF complex in the case of clinically used aliphatic and aromatic compounds. An exception were boron and stannane precursors in Cu-mediated 18F-labeling reactions where the RCYs obtained using [18F]ESF were lower than the ones obtained using [18F]TEAF. This might be due to the use of [18F]TEAF instead of [18F]KF/K222 complexes reported previously, to different automated drying sequences or to other currently uncontrolled parameters. It was also observed that the standard deviations for the radiofluorinations were generally lower when using [18F]ESF, which could be linked to the consistent quality of this labeling reagent compared to the [18F]TEAF coming from a drying process that might introduce varying grades of moisture.

Fig. 14
figure 14

 A Radiosynthesis of [18F]FDG via the gaseous [18F]HF synthon. B Radiosynthesis of 1-(tert-butyl)-4-[18F]fluorobenzene via the gaseous [18F]acetyl fluoride synthon. C Radiosynthesis of [18F]FEPPA via the gaseous [18F]triflyl fluoride synthon. D Radiosynthesis of protected [18F]mFBG via the gaseous [18F]ethenesulfonyl fluoride synthon; n.m.: not mentioned

Synthesis of anhydrous [18F]fluoride via electrochemical deposition and magnetic droplets

Electrochemical deposition

Instead of anion-exchange resins, concentration of [18F]fluoride from the target water has also been performed using electrodeposition (Cardinale et al. 2014; Hamacher et al. 2002; Hamacher and Coenen 2006; Kronenberg et al. 2007; Melean et al. 2011; Meleán et al. 2015; Reischl et al. 2002; Sadeghi et al. 2013; Saito et al. 2007; Wagner et al. 2009). Basically, [18F]fluoride from the water target is deposited on the anodic site of an electrochemical cell after which it is recovered under reversed voltage in aprotic solvent containing a phase transfer catalyst. Although this method has proven to be not very efficient, its convenience has led to several applications, especially in microfluidic devices. For example, Saiki et al. reported a microfluidic cell that could trap [18F]fluoride from 1 to 2 mL of [18O]H2O by applying a potential and then release the activity into 275 µL of eluent with an overall efficiency of ~ 60% (Saiki et al. 2010; Wong et al. 2012). An alternative platform makes use electrowetting-on-dielectric (EWOD) (Chen et al. 2014; Keng et al. 2012). An EWOD microfluidic chip typically consists of two parallel plates, one patterned with electrodes and one coated with a conductor to act as a ground electrode. An electrical potential is then applied to the desired electrodes to achieve operations such as droplet transport or heating. For example, a 200 µL droplet of [18F]fluoride in [18O]H2O with base and phase transfer catalyst is loaded and its volume reduced by heating the nearby electrodes. An electrical potential is then applied to the electrodes to move the droplet to a heating site. Here, the target water is further removed by evaporation resulting in a residue of 5 µL in 10 min. [18F]FLT and [18F]Fallypride have been successfully synthesized on an EWOD device in 63 ± 5% RCY (n = 5) and 65 ± 6% RCY (n = 7), respectively (Javed et al. 2014a, b). These yields are comparable to, or greater than, those obtained by conventional approaches e.g. 40% RCY (n = 1) for [18F]FLT and 66 ± 8% RCY (n = 8) for [18F]Fallypride (Lazari et al. 2014; Suehiro et al. 2007). In addition, the Am of [18F]Fallypride (730 GBq/µmol) was two times higher on the EWOD device than synthesized in a macroscale radiosynthesizer (330 GBq/µmol), despite starting with significantly less radioactivity (270 MBq and 7400 MBq, respectively). Although this approach is suitable for modest starting volumes, impractically large chips and many time-consuming sequential 200 µL evaporations are required to handle volumes of ~ 2 mL of [18F]fluoride in enriched water that are generated in most cyclotrons.

Magnetic droplet microfluidics (MDM)

Finally, Fiel et al. described the use of magnetic particles to manipulate [18O]H2O droplets in the [18F]fluoride preconcentration step (Fiel et al. 2015). A magnet is moved to manipulate the magnetic particles which hold the small target water droplet by surface tension, but also possess an anion-exchange property that is capable of capturing and releasing the [18F]F. The capture efficiency was 79% using low initial activities (~ 74 MBq) and 25 mg of magnetic particles. After evaporation of the solvent, [18F]F is released with 50 µL of K2CO3 (7.3 µmol) in 91–95% efficiency. The preconcentration step took approximately 5 min and the [18F]fluoride solution was preconcentrated by 15-fold. However, it was found that when the initial 18F-activity was increased to ~ 300 MBq, the capture efficiency already decreased to ca. 59%, which makes clinical application questionable.

Conclusions

The large number of publications on the preparation of reactive [18F]fluoride from oxygen-18 enriched water over the last decade reflects the growing importance of the complex activation step in 18F-fluorination chemistry. Success and progress have been achieved in base-sensitive nucleophilic 18F-fluorination reactions, providing improved, modified or even completely new fluorine-18 work-up procedures. The currently most used elution system still comprises the conventional K2CO3/K222 eluent, but the mild base tetraethylammonium bicarbonate (TEAB) as phase-transfer catalysts has been shown to be a useful alternative in the 18F-labeling of iodonium(III) containing precursors, e.g. [18F]UCB-J. In addition, the moderately basic K2CO3/K2C2O4/K222 eluent as well as positively charged precursors in alcoholic medium, i.e. the minimalist approach, in especially Cu-mediated nucleophilic 18F-fluorination reactions have found widespread use.

It should be mentioned that many developments listed in this review only comprise small adaptations compared to the conventional methodology and are focused on a specific reaction or substrate. Hence, several of these methods have already proven their advantage and suitability for 18F-labeled tracer synthesis, but more comparative evaluations are desirable. In contrast, new approaches have been reported and hold promise, but often only concern specific reactions with primarily simple substrates as application. From these methods, the recent [18F]triflyl fluoride procedure has to be mentioned as a successful example of a new [18F]fluoride work-up methodology because of its use in clinical PET tracer production.

Despite the large variety of new activation procedures, none has completely replaced the conventional K2CO3/K222 elution method and further expansion of miniaturization and kit-like procedures is an area of research with room for improvement. Automated cassettes and technologies like microfluidic devices to produce reactive [18F]fluoride will increasingly be used in PET chemistry by providing safe, standardized and automated GMP production tools. Considering the many promising developments going on, there is great confidence in further inventions and the establishment of a few standardized drying methodologies for the routine production of a broad scale of PET tracers.

Availability of data and materials

Not applicable.

Abbreviations

[18F]-5′-FDA:

[18F]-5′-fluoro-5′-deoxyadenosine

[18F]-L-FPHCys:

L-S-(3-[18F]fluoropropyl)homocysteine

[18F]4 F-MHPG:

4-[18F]fluoro‐m‐hydroxy-phenethylguanidine

[18F]AcF:

[18F]acetyl fluoride

[18F]DAA1106:

N-(2,5-dimethoxybenzyl)-N-(5-[18F]fluoro-2- phenoxyphenyl)acetamide

[18F]DCFPyL:

2-(3‐{1‐Carboxy‐5‐[(6‐[18F]fluoro‐pyridine‐3‐carbonyl)‐amino]‐pentyl}‐ureido)‐pentanedioic acid

[18F]ESF:

[18F]ethenesulfonyl fluoride

[18F]Fallypride:

(S)-N-[(1-allyl-2-pyrrolidinyl)methyl]-5-(3-[18F]fluoropropyl)-2,3-dimethoxybenzamide

[18F]FAZA:

[18F]fluoroazomycin arabinosid

[18F]FDG:

2-[18F]fluoro-2-deoxy-D-glucose

[18F]FDHT:

16β-[18F]fluoro-5α-dihydrotestosteron

[18F]FDOPA:

6-[18F]fluoro-L-dihydroxy-phenylalanine

[18F]FDPA:

N,N-diethyl-2-(2-(4-[18F]fluorophenyl)-5,7-dimethylpyrazolo[1,5- a]pyrimidine-3-yl)-acetamide

[18F]FEPPA:

N-acetyl-N-(2-[18F]fluoroethoxybenzyl)-2-phenoxy-5-pyridinamine

[18F]FES:

16α-[18F]fluoroestradiol

[18F]FET:

O-(2-[18F]fluoroethyl)-L-tyrosine

[18F]FLT:

3′-Deoxy-3′-[18F]fluorothymidine

[18F]FMAU:

2′-Deoxy-2′-[18F]fluoro-5-methyl-1-β-D-arabinofuranosyluracil

[18F]FMISO:

3-[18F]fluoro-1-(2′-nitro-1′-imidazolyl)-2-propanol

[18F]FMT:

L-3-[18F]fluoro-α-methyltyrosine

[18F]FMZ:

[18F]Flumazenil

[18F]FP-CIT:

N-[18F]fluoropropylcarbomethoxyiodophenyl-nortropane

[18F]FPEB:

[18F]-3-fluoro-5-[(pyridin-3-yl)ethynyl]benzonitrile

[18F]FPhe:

L-4-[18F]fluorophenylalanine

[18F]FTrp:

6-[18F]fluoro-L-tryptophan

[18F]mFBG:

meta-[18F]fluorobenzylguanidine

[18F]T807:

7-(6-[18F]fluoropyridin-3-yl)-5 H-pyrido[4,3-β]indole

[18O]H2O:

Oxygen-18 enriched water

3-[18F]F-NNDI:

N-(3-fluoro-4-nitronaphthyl)-cis-5-norbornene-endo-2,3-dicarboxylic imide

4-[18F]FGln:

(2 S,4R)-4-[18F]fluoroglutamine

Am :

Molar activity

AY:

Activity yield

DMA:

N,N-dimethylacetamide

DMAP·OTf:

4-Dimethylaminopyridinium trifluoromethanesulfonate

DMF:

N,N-dimethylformamide

EtOH:

Ethanol

EWOD:

Electrowetting-on-dielectric

GMP:

Good manufacturing practice

K222 :

kryptofix-222

MeCN:

Acetonitrile

MeOH:

Methanol

nBuOH:

n-butanol

P2Et-[18F]HF:

Phosphazenium [18F]hydrofluoride

PET:

Positron emission tomography

RCC:

Radiochemical conversion

RCY:

Radiochemical yield

SCIDY:

Spirocyclic iodonium ylides

SPI5:

6,10-Dioxaspiro[4.5]decane-7,9-dione

SPIAd:

Spiroadamantyl-1,3-dioxane-4,6-dione

TBACN:

Tetrabutylammonium cyanide

TBAHCO3 :

Tetrabutylammonium bicarbonate

TBAOH:

Tetrabutylammonium hydroxide

TBAOMs:

Tetrabutylammonium methanesulfonate

TBAOTf:

Tetrabutylammonium triflate

tBuOH:

tert-butanol

TEAB:

Tetraethylammonium bicarbonate

TEAOTf:

Tetraethyl ammonium triflate

References

  • Aerts J, Voccia S, Lemaire C, Giacomelli F, Goblet D, Thonon D, et al. Fast production of highly concentrated reactive [18F]fluoride for aliphatic and aromatic nucleophilic radiolabelling. Tetrahedron Lett. 2010;51(1):64–6.

    CAS  Google Scholar 

  • Ahmed H, Haider A, Varisco J, Stankovic M, Wallimann R, Gruber S, et al. Structure–affinity relationships of 2,3,4,5-tetrahydro–1H–3-benzazepine and 6,7,8,9-tetrahydro–5H–benzo[7]annulen-7-amine analogues and the discovery of a radiofluorinated 2,3,4,5-tetrahydro–1H–3-benzazepine congener for imaging GluN2B subunit-containing. J Med Chem. 2019;62:9450–70.

    CAS  PubMed  Google Scholar 

  • Andersen IV, García-vázquez R, Battisti UM, Herth MM. Optimization of direct aromatic 18F‐Labeling of tetrazines. Molecules. 2022;27(13):1–11.

    Google Scholar 

  • Antuganov D, Zykov M, Timofeev V, Timofeeva K, Antuganova Y, Orlovskaya V, et al. Copper-mediated radiofluorination of aryl pinacolboronate esters: a straightforward protocol by using pyridinium sulfonates. Eur J Org Chem. 2019;5:918–22.

    Google Scholar 

  • Bailey JJ, Kaiser L, Lindner S, Wu M, Thiel A, Soucy J, et al. First-in-human brain imaging of [18F]TRACK, a PET tracer for tropomyosin receptor kinases. ACS Chem Neurosci. 2019;10:1697–2702.

    Google Scholar 

  • Basuli F, Zhang X, Jagoda EM, Choyke PL, Swenson RE. Rapid synthesis of maleimide functionalized fluorine-18 labeled prosthetic group using “radio‐fluorination on the Sep‐Pak” method. J Label Compd Radiopharm. 2018;61:599–605.

    CAS  Google Scholar 

  • Bejot R, Elizarov AM, Ball E, Zhang J, Miraghaie R, Kolb HC, et al. Batch-mode microfluidic radiosynthesis of N-succinimidyl-4-[18F]fluorobenzoate for protein labelling. J Label Compd Radiopharm. 2011;54:117–22.

    CAS  Google Scholar 

  • Bernard-Gauthier V, Mossine AV, Mahringer A, Aliaga A, Bailey JJ, Shao X, et al. Identification of [18F]TRACK, a fluorine-18-Labeled tropomyosin receptor kinase (trk) inhibitor for PET imaging. J Med Chem. 2018;61(4):1737–43.

    CAS  PubMed  Google Scholar 

  • Berroteran-Infante N, Balber T, Furlinger P, Bergmann M, Lanzenberger R, Hacker M, et al. [18F]FEPPA: improved automated radiosynthesis, binding affinity, and preliminary in vitro evaluation in colorectal cancer. ACS Med Chem Lett. 2018;9:177–81.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Beyzavi MH, Mandal D, Strebl MG, Neumann CN, D’Amato EM, Chen J, et al. 18F–Deoxyfluorination of phenols via Ru π-complexes. ACS Cent Sci. 2017;3:944–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Bongarzone S, Basagni F, Sementa T, Singh N, Gakpetor C, Faugeras V, et al. Development of [18F]FAMTO: a novel fluorine-18 labelled positron emission tomography (PET) radiotracer for imaging CYP11B1 and CYP11B2 enzymes in adrenal glands. Nucl Med Biol. 2019;69:14–21.

    Google Scholar 

  • Bourdier T, Greguric I, Roselt P, Jackson T, Faragalla J, Katsifis A. Fully automated one-pot radiosynthesis of O-(2-[18F]fluoroethyl)-L-tyrosine on the TracerLab FX FN module. Nucl Med Biol. 2011a;38(5):645–51.

    CAS  PubMed  Google Scholar 

  • Bourdier T, Shepherd R, Berghofer P, Jackson T, Fookes CJR, Denoyer D, et al. Radiosynthesis and biological evaluation of L- and D-S-(3-[18F]fluoropropyl)homocysteine for tumor imaging using positron emission tomography. J Med Chem. 2011b;54:1860–70.

    CAS  PubMed  Google Scholar 

  • Bouvet V, Wuest M, Jans H, Janzen N, Genady AR, Valliant JF, et al. Automated synthesis of [18F]DCFPyL via direct radiofluorination and validation in preclinical prostate cancer models. EJNMMI Res. 2016;6(40):1–15.

    Google Scholar 

  • Bratteby K, Torkelsson E, L’Estrada ET, Peterson K, Shalgunov V, Xiong M, et al. In vivo veritas: 18F–radiolabeled glycomimetics allow insights into the pharmacological fate of galectin–3 inhibitors. J Med Chem. 2020;63:747–55.

    CAS  PubMed  Google Scholar 

  • Bratteby K, Shalgunov V, Battisti UM, Petersen IN, Van Den Broek SL, Ohlsson T, et al. Insights into elution of anion exchange cartridges: opening the path toward aliphatic 18F-radiolabeling of base-sensitive tracers. ACS Pharmacol Transl Sci. 2021;4(5):1556–66.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Brichard L, Aigbirhio FI. An efficient method for enhancing the reactivity and flexibility of [18F]fluoride towards nucleophilic substitution using tetraethylammonium bicarbonate. Eur J Org Chem. 2014;28:6145–9.

    Google Scholar 

  • Cai Z, Li S, Pracitto R, Navarro A, Shirali A, Ropchan J, et al. Fluorine-18-labeled antagonist for PET imaging of kappa opioid receptors. ACS Chem Neurosci. 2017;8:12–6.

    CAS  PubMed  Google Scholar 

  • Cai Z, Li S, Zhang W, Pracitto R, Wu X, Baum E, et al. Synthesis and preclinical evaluation of an 18F–labeled synaptic vesicle glycoprotein 2A PET imaging probe: [18F]SynVesT–2. ACS Chem Neurosci. 2020;11:592–603.

    CAS  PubMed  Google Scholar 

  • Cardinale J, Ermert J, Humpert S, Coenen HH. Iodonium ylides for one-step, no-carrier-added radiofluorination of electron rich arenes, exemplified with 4-(([18F]fluorophenoxy)phenylmethyl)piperidine NET and SERT ligands. RSC Adv. 2014;4:17293–9.

    CAS  Google Scholar 

  • Cardinale J, Martin R, Remde Y, Schäfer M, Hienzsch A, Hübner S, et al. Procedures for the GMP-compliant production and quality control of [18F]PSMA-1007: a next generation radiofluorinated tracer for the detection of prostate cancer. Pharmaceuticals. 2017;10(77):1–18.

    Google Scholar 

  • Celen S, Rokka J, Gilbert TM, Koole M, Vermeulen I, Serdons K, et al. Translation of HDAC6 PET imaging using [18F]EKZ-001-cGMP production and measurement of HDAC6 target occupancy in nonhuman primates. ACS Chem Neurosci. 2020;11(7):1093–101.

    CAS  PubMed  Google Scholar 

  • Chang CW, Chou TK, Liu RS, Wang SJ, Lin WJ, Chen CH, et al. A robotic synthesis of [18F]fluoromisonidazole ([18F]FMISO). Appl Radiat Isot. 2007;65:682–6.

    CAS  PubMed  Google Scholar 

  • Chao PH, Lazari M, Hanet S, Narayanam MK, Murphy JM, Van Dam RM. Automated concentration of [18F]fluoride into microliter volumes. Appl Radiat Isot. 2018;141:138–48.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Chen S, Javed MR, Kim H, Lei J, Lazari M, Shah GJ, et al. Lab on a chip tomography (PET) tracers using a single digital microfluidic reactor chip. Lab Chip. 2014;14:902–10.

    CAS  PubMed  Google Scholar 

  • Cherif A, Yang DJ, Tansey W, Kim EE, Wallace S. Rapid synthesis of 3-[18F]fluoro-1-(2′-nitro-1′-imidazolyl)-2-propanol ([18F]fluoromisonidazole). Pharm Res. 1994;11(3):466–9.

    CAS  PubMed  Google Scholar 

  • Clemente GS, Zarganes-tzitzikas T, Dömling A, Elsinga PH. Late-stage copper-catalyzed radiofluorination of an arylboronic ester derivative of atorvastatin. Molecules. 2019;24(4210):1–9.

    Google Scholar 

  • Clemente GS, Rickmeier J, Antunes IF, Zarganes-tzitzikas T, Dömling A, Ritter T, et al. [18F]atorvastatin: synthesis of a potential molecular imaging tool for the assessment of statin-related mechanisms of action. EJNMMI Res. 2020;10(34):1–13.

    Google Scholar 

  • Coenen HH, Gee AD, Adam M, Antoni G, Cutler CS, Fujibayashi Y, et al. Consensus nomenclature rules for radiopharmaceutical chemistry—setting the record straight. Nucl Med Biol. 2017;55:v–xi.

    CAS  PubMed  Google Scholar 

  • Collier TL, Normandin MD, Stephenson NA, Livni E, Liang SH, Wooten DW, et al. Synthesis and preliminary PET imaging of 11 C and 18F isotopologues of the ROS1/ALK inhibitor lorlatinib. Nat Commun. 2017;8:1–12.

    Google Scholar 

  • Crouzel C, Venet M, Irie T, Sanz G, Boullais C. Labeling of a serotoninergic ligand with 18F: [18F]setoperone. J Label Compd Radiopharm. 1988;25(4):403–14.

    CAS  Google Scholar 

  • Dahl K, Garcia A, Stephenson NA, Vasdev N. In-loop” 18F‐fluorination: a proof‐of‐concept study. J Label Compd Radiopharm. 2019;62:292–7.

    CAS  Google Scholar 

  • Deng H, Cobb SL, Gee AD, Lockhart A, Martarello L, Mcglinchey RP, et al. Fluorinase mediated C–18F bond formation, an enzymatic tool for PET labelling. Chem Commun. 2006;6:652–4.

    Google Scholar 

  • Deng X, Rong J, Wang L, Vasdev N, Zhang L, Josephson L, et al. Chemistry for positron emission tomography: recent advances in 11 C-, 18F-, 13 N-, and 15O-labeling reactions. Angew Chemie Int Ed. 2019;58(9):2580–605.

    CAS  Google Scholar 

  • Dornan MH, Simard J-M, Leblond A, Juneau D, Delouya G, Saad F, et al. Simplified and robust one-step radiosynthesis of [18F]DCFPyL via direct radiofluorination and cartridge‐based purification. J Label Compd Radiopharm. 2018;61:757–63.

    CAS  Google Scholar 

  • Drerup C, Ermert J, Coenen HH. Synthesis of a potent aminopyridine-based nNOS-inhibitor by two recent no-carrier-added 18F-labelling methods. Molecules. 2016;21(1160):1–21.

    Google Scholar 

  • Elie J, Vercouillie J, Arlicot N, Lemaire L, Bidault R, Bodard S, et al. Design of selective COX-2 inhibitors in the (aza)indazole series. Chemistry, in vitro studies, radiochemistry and evaluations in rats of a [18F]PET tracer. J Enzyme Inhib Med Chem. 2019;34(1):1–7.

    CAS  PubMed  Google Scholar 

  • Elizarov AM, Van Dam RM, Shin YS, Kolb HC, Padgett HC, Stout D, et al. Design and optimization of coin-shaped microreactor chips for PET radiopharmaceutical synthesis. J Nucl Med. 2010;51(2):282–8.

    CAS  PubMed  Google Scholar 

  • Fedorova O, Nikolaeva V, Krasikova R. Automated SPE-based synthesis of 16α-[18F]fluoroestradiol without HPLC purification step. Appl Radiat Isot. 2018;141:57–63.

    CAS  PubMed  Google Scholar 

  • Feni L, Omrane MA, Fischer M, Zlatopolskiy BD, Neumaier B, Neundorf I. Convenient preparation of 18F-labeled peptide probes for potential claudin-4 PET imaging. Pharmaceuticals. 2017;10(99):1–15.

    Google Scholar 

  • Fiel SA, Yang H, Scha P, Weng S, Inkster JAH, Wong MCK, et al. Magnetic droplet microfluidics as a platform for the concentration of [18F]fluoride and radiosynthesis of sulfonyl [18F]fluoride. ACS Appl Mater. 2015;7:12923–9.

    CAS  Google Scholar 

  • Fujinaga M, Kumata K, Zhang Y, Hatori A, Yamasaki T, Mori W, et al. Synthesis of two novel [18F]fluorobenzene-containing radiotracers via spirocyclic iodonium ylides and positron emission tomography imaging of translocator protein (18 kDa) in ischemic brain. Org Biomol Chem R Soc Chem. 2018;16(37):8325–35.

    Google Scholar 

  • Gu Y, Huang D, Liu Z, Huang J, Zeng W. Labeling strategies with F-18 for positron emission tomography imaging. Med Chem (Los Angeles). 2011;7(5):334–44.

    CAS  Google Scholar 

  • Guibbal F, Meneyrol V, Ait-Arsa I, Diotel N, Patché J, Veeren B, et al. Synthesis and automated labeling of [18F]Darapladib, a Lp-PLA2 ligand, as potential PET imaging tool of atherosclerosis. ACS Med Chem Lett. 2019;10(5):743–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Guibbal F, Isenegger PG, Wilson TC, Pacelli A, Mahaut D, Sap JBI, et al. Manual and automated Cu-mediated radiosynthesis of the PARP inhibitor [18F]olaparib. Nat Protoc. 2020;15(April):1525–41.

    CAS  PubMed  Google Scholar 

  • Guillouet S, Patin D, Tirel O, Delamare J, Gourand F, Deloye JB, et al. Fully automated radiosynthesis of 2-[18F]fludarabine for PET imaging of low-grade lymphoma. Mol Imaging Biol. 2014;16:28–35.

    PubMed  Google Scholar 

  • Haider A, Iten I, Ahmed H, Adrienne M, Gruber S, Kr SD, et al. Identification and preclinical evaluation of a radiofluorinated benzazepine derivative for imaging the GluN2B subunit of the ionotropic NMDA receptor. J Nucl Med. 2019;60(2):259–67.

    CAS  PubMed Central  Google Scholar 

  • Hamacher K, Coenen HH. No-carrier-added nucleophilic 18F-labelling in an electrochemical cell exemplified by the routine production of [18F]altanserin. Appl Radiat Isot. 2006;64:989–94.

    CAS  PubMed  Google Scholar 

  • Hamacher K, Hamkens W. Remote controlled cne-step production of 18F labeled butyrophenone neuroleptics exemplified by the synthesis of n.c.a. [18F]N-methylspiperone. Appl Radiat Isot. 1995;46(9):911–6.

    CAS  Google Scholar 

  • Hamacher K, Coenen HH, Stocklin G. Efficient stereospecific synthesis of no-carrier-added 2-[18F]Fluoro-2-Deoxy-D-Glucose using aminopolyether supported nucleophilic substitution. J Nucl Med. 1986;27:235–8.

    CAS  PubMed  Google Scholar 

  • Hamacher K, Hirschfelder T, Coenen HH. Electrochemical cell for separation of [18F]fluoride from irradiated 18O-water and subsequent no carrier added nucleophilic fluorination. Appl Radiat Isot. 2002;56:519–23.

    CAS  PubMed  Google Scholar 

  • Hamill TG, Krause S, Ryan C, Bonnefous C, Govek S, Seiders TJ, et al. Synthesis, characterization, and first successful monkey imaging studies of metabotropic glutamate receptor subtype 5 (mGluR5) PET radiotracers. Synapse. 2005;56:205–16.

    CAS  PubMed  Google Scholar 

  • Hamill TG, Eng W, Jennings A, Lewis R, Thomas S, Wood S, et al. The synthesis and preclinical evaluation in rhesus monkey of [18F]MK-6577 and [11 C]CMPyPB glycine transporter 1 positron emission tomography radiotracers. Synapse. 2011;65:261–70.

    CAS  PubMed  Google Scholar 

  • Hayashi K, Furutsuka K, Takei M, Muto M, Nakao R, Aki H, et al. High-yield automated synthesis of [18F]fluoroazomycin arabinoside ([18F]FAZA) for hypoxia-specific tumor imaging. Appl Radiat Isot. 2011;69(7):1007–13.

    CAS  PubMed  Google Scholar 

  • Herth MM, Ametamey S, Antuganov D, Bauman A, Berndt M, Brooks AF, et al. On the consensus nomenclature rules for radiopharmaceutical chemistry: reconsideration of radiochemical conversion. Nucl Med Biol. 2021;93:19–21.

    CAS  PubMed  Google Scholar 

  • Hoover AJ, Lazari M, Ren H, Narayanam MK, Murphy JM, Van Dam RM, et al. A transmetalation reaction enables the synthesis of [18F]5-fluorouracil from [18F]fluoride for human PET imaging. Organometallics. 2016;35:1008–14.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Hostetler ED, Eng W, Joshi AD, Sanabria-bohórquez S, Kawamoto H, Ito S, et al. Synthesis, characterization, and monkey PET studies of [18F]MK-1312, a PET tracer for quantification of mGluR1 receptor occupancy by MK-5435. Synapse. 2011a;65:125–35.

    CAS  PubMed  Google Scholar 

  • Hostetler ED, Sanabria-bohórquez S, Fan H, Zeng Z, Gammage L, Miller P, et al. [18F]fluoroazabenzoxazoles as potential amyloid plaque PET tracers: synthesis and in vivo evaluation in rhesus monkey. Nucl Med Biol. 2011b;38(8):1193–203.

    CAS  PubMed  Google Scholar 

  • Huang X, Liu W, Ren H, Neelamegam R, Hooker JM, Groves JT. Late stage benzylic C–H fluorination with [18F]fluoride for PET imaging. J Am Chem Soc. 2014;136(19):6842–5.

    CAS  PubMed  Google Scholar 

  • Ismail R, Irribaren J, Javed R, Machness A, Van Dam RM, Keng PY. Cationic imidazolium polymer monoliths for efficient solvent exchange, activation and fluorination on a continuous flow system. RSC Adv. 2014;4:25348–56.

    CAS  Google Scholar 

  • Iwata R, Pascali C, Terasaki K, Ishikawa Y, Furumoto S, Yanai K. Minimization of the amount of Kryptofix 222 - KHCO3 for applications to microscale 18F-radiolabeling. Appl Radiat Isot. 2017;125:113–8.

    CAS  PubMed  Google Scholar 

  • Iwata R, Pascali C, Terasaki K, Ishikawa Y, Furumoto S, Yanai K. Practical microscale one-pot radiosynthesis of 18F‐labeled probes. J Label Compd Radiopharm. 2018;61:540–9.

    CAS  Google Scholar 

  • Javed MR, Chen S, Kim H, Wei L, Czernin J, Kim C, et al. Efficient radiosynthesis of 3′-deoxy-3′-18F-fluorothymidine using electrowetting-on-dielectric digital microfluidic chip. J Nucl Med. 2014a;55:321–9.

    CAS  PubMed  Google Scholar 

  • Javed MR, Chen S, Lei J, Collins J, Sergeev M, Kim H, et al. High yield and high specific activity synthesis [18F]fallypride in a batch microfluidic reactor for micro-PET imaging. Chem Commun. 2014b;50:1192–4.

    CAS  Google Scholar 

  • Jiang H, Dimagno SG, Degrado TR. Production and transport of gaseous 18F-synthons: 18F-acyl fluoride. J Fluor Chem. 2015;180:181–5.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Jung YW, Gu G, Raffel DM. Improved synthesis of 4-[18F]fluoro-m-hydroxyphenethylguanidine using an iodonium ylide precursor. J Label Compd Radiopharm. 2019;62(12):835–42.

    CAS  Google Scholar 

  • Katsifis A, Hamachert K, Schnitter J, Stocklin G. Optimization studies concerning the direct nucleophilic fluorination of butyrophenone neuroleptics. Appl Radiat Isot. 1993;44(7):1015–20.

    CAS  Google Scholar 

  • Keng PY, Chen S, Ding H, Sadeghi S, Shah GJ, Dooraghi A et al. Micro-chemical synthesis of molecular probes on an electronic microfluidic device. Proc Natl Acad Sci. 2012;109(3):690–5.

    CAS  PubMed  Google Scholar 

  • Kim DW, Choe YS, Chi DY. A new nucleophilic fluorine-18 labeling method for aliphatic mesylates: reaction in ionic liquids shows tolerance for water. Nucl Med Biol. 2003;30:345–50.

    CAS  PubMed  Google Scholar 

  • Kim HW, Jeong JM, Lee Y-S, Chi DY, Chung K-H, Lee DS, et al. Rapid synthesis of [18F]FDG without an evaporation step using an ionic liquid. Appl Radiat Isot. 2004;61:1241–6.

    CAS  PubMed  Google Scholar 

  • Knott KE, Grätz D, Hübner S, Jüttler S, Zankl C, Müller M. Simplified and automatic one-pot synthesis of 16a-[18F]fluoroestradiol without high-performance liquid chromatography purification. J Label Compd Radiopharm. 2011;54:749–53.

    CAS  Google Scholar 

  • Kronenberg UB, Drewes B, Sihver W, Coenen HH. N-2-(4-N-(4-[18F]fluorobenzamido)phenyl)-propyl-2- propanesulphonamide: synthesis and radiofluorinatio of a putative AMPA receptor ligand. J Label Compd Radiopharm. 2007;50:1169–75.

    CAS  Google Scholar 

  • Kuik W, Kema IP, Brouwers AH, Zijlma R, Neumann KD, Dierckx RAJO, et al. In vivo biodistribution of no-carrier-added 6-18F-fluoro-3,4-dihydroxy-L-phenylalanine (18F-DOPA), produced by a new nucleophilic substitution approach, compared with carrier-added 18F-DOPA, prepared by conventional electrophilic substitution. J Nucl Med. 2015;56(1):106–12.

    CAS  PubMed  Google Scholar 

  • Kumata K, Zhang Y, Fujinaga M, Ohkubo T, Mori W, Yamasaki T, et al. [18F]DAA1106: automated radiosynthesis using spirocyclic iodonium ylide and preclinical evaluation for positron emission tomography imaging of translocator protein (18 kDa). Bioorg Med Chem. 2018;26(17):4817–22.

    CAS  PubMed  Google Scholar 

  • Labas R, Gilbert G, Nicole O, Dhilly M, Abbas A, Tirel O, et al. Synthesis, evaluation and metabolic studies of radiotracers containing a 4-(4-[18F]-fluorobenzyl)piperidin-1-yl moiety for the PET imaging of NR2B NMDA receptors. Eur J Med Chem. 2011;46(6):2295–309.

    CAS  PubMed  Google Scholar 

  • Lahdenpohja SO, Rajala NA, Rajander J, Kirjavainen AK. Fast and efficient copper-mediated 18F-fluorination of arylstannanes, aryl boronic acids, and aryl boronic esters without azeotropic drying. EJNMMI Radiopharm Chem. 2019;4(28):1–11.

    Google Scholar 

  • Lakshminarayanan N, Kumar A, Roy S, Pawar Y, Chaudhari P. Improved method for preparing Ni(II) complex of (S)-tyrosine Schiff base and its use in the automated synthesis of O-(2′-[18F]fluoroethyl)-L-tyrosine using solid-phase extraction purification. Appl Radiat Isot. 2017;127:122–9.

    CAS  PubMed  Google Scholar 

  • Lazari M, Collins J, Shen B, Farhoud M, Yeh D, Maraglia B, et al. Fully automated production of diverse 18F-labeled PET tracers on the ELIXYS multireactor radiosynthesizer without hardware modification. J Nucl Med. 2014;42(3):203–10.

    Google Scholar 

  • Lazari M, Lyashchenko SK, Burnazi EM, Lewis JS, Van Dam RM, Murphy JM. Fully-automated synthesis of 16β-18F-fluoro-5α-dihydrotestosteron (FDHT) on the ELIXYS radiosynthesizer. Appl Radiat Isot. 2015;103:9–14.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Lebedev A, Miraghaie R, Kotta K, Ball CE, Zhang J, Buchsbaum MS, et al. Batch-reactor microfluidic device: first human use of a microfluidically produced PET radiotracer. Lab Chip. 2013;13:136–45.

    CAS  PubMed  Google Scholar 

  • Lee C, Sui G, Elizarov A, Shu CJ, Shin Y, Dooley AN, et al. Multistep synthesis of a radiolabeled imaging probe using integrated microfluidics. Science. 2005;310:1793–7.

    CAS  PubMed  Google Scholar 

  • Lee BC, Kim JS, Kim BS, Son JY, Hong SK, Park HS, et al. Aromatic radiofluorination and biological evaluation of 2-aryl-6-[18F]fluorobenzothiazoles as a potential positron emission tomography imaging probe for b-amyloid plaques. Bioorg Med Chem. 2011a;19(9):2980–90.

    CAS  PubMed  Google Scholar 

  • Lee SJ, Oh SJ, Moon WY, Choi MS, Kim JS, Chi DY, et al. New automated synthesis of [18F]FP-CIT with base amount control affording high and stable radiochemical yield: a 1.5-year production report. Nucl Med Biol. 2011b;38(4):593–7.

    CAS  PubMed  Google Scholar 

  • Lee SJ, Oh SJ, Chi DY, Moon DH, Ryu JS. High yielding [18F]fluorination method by fine control of the base. Bull Korean Chem Soc. 2012;33(7):2177–80.

    CAS  Google Scholar 

  • Lee SJ, Hyun JS, Oh SJ, Yu KH, Ryu JS. Development of a new precursor-minimizing base control method and its application for the automated synthesis and SPE purification of [18F]fluoromisonidazole ([18F]FMISO). J Label Compd Radiopharm. 2013;56:731–5.

    CAS  Google Scholar 

  • Lemaire CF, Aerts JJ, Voccia S, Libert LC, Mercier F, Goblet D, et al. Fast production of highly reactive no-carrier-added [18F]fluoride for the labeling of radiopharmaceuticals. Angew Chemie Int Ed. 2010;49:3161–4.

    CAS  Google Scholar 

  • Leonardis F, De, Pascali G, Salvadori PA, Watts P, Pamme N. On-chip pre-concentration and complexation of [18F]fluoride ions via regenerable anion exchange particles for radiochemical synthesis of positron emission tomography tracers. J Chromatogr A. 2011;1218(29):4714–9.

    PubMed  Google Scholar 

  • Li Z, Cai H, Conti PS. Automated synthesis of 2′-deoxy-2′-[18F]fluoro-5-methyl-1-β-D-arabinofuranosyluracil ([18F]-FMAU) using a one reactor radiosynthesis module. Nucl Med Biol. 2011;38(2):201–6.

    CAS  PubMed  Google Scholar 

  • Li S, Cai Z, Wu X, Holden D, Pracitto R, Kapinos M, et al. Synthesis and in vivo evaluation of a novel PET radiotracer for imaging of synaptic vesicle glycoprotein 2A (SV2A) in nonhuman primates. ACS Chem Neurosci. 2019a;10:1544–54.

    CAS  PubMed  Google Scholar 

  • Li S, Cai Z, Zhang W, Holden D, Lin S, Finnema SJ, et al. Synthesis and in vivo evaluation of [18F]UCB-J for PET imaging of synaptic vesicle glycoprotein 2A (SV2A). Eur J Nucl Med Mol Imaging. 2019b;46:1952–65.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Liang SH, Wang L, Stephenson NA, Rotstein BH, Vasdev N. Facile 18F labeling of non-activated arenes via a spirocyclic iodonium(III) ylide method and its application in the synthesis of the mGluR 5 PET radiopharmaceutical [18F]FPEB. Nat Protoc. 2019;14(5):1530–45.

    CAS  PubMed  Google Scholar 

  • Lieberman BP, Ploessl K, Wang L, Qu W, Zha Z, Wise DR, et al. PET imaging of glutaminolysis in tumors by 18F-(2S,4R)4-Fluoroglutamine. J Nucl Med. 2011;52(12):1947–55.

    CAS  PubMed  Google Scholar 

  • Lien VT, Klaveness J, Olberg DE. One-step synthesis of [18F]cabozantinib for use in positron emission tomography imaging of c-Met. J Label Compd Radiopharm. 2018;61:11–7.

    CAS  Google Scholar 

  • Lindner S, Rensch C, Neubaur S, Neumeier M, Salvamoser R, Samper V, et al. Azeotropic drying free [18F]FDG synthesis and its application to a lab-on-chip platform. Chem Commun R Soc Chem. 2016;52:729–32.

    CAS  Google Scholar 

  • Litchfield M, Wuest M, Glubrecht D, Wuest F. Radiosynthesis and biological evaluation of [18F]triacoxib: a new radiotracer for PET imaging of COX–2. Mol. Pharm. 2020;17:251–61.

    CAS  Google Scholar 

  • Liu W, Huang X, Placzek MS, Krska SW, Mcquade P, Hooker JM, et al. Site-selective 18F fluorination of unactivated C–H bonds mediated by a manganese porphyrin. Chem Sci R Soc Chem. 2018;9:1168–72.

    CAS  Google Scholar 

  • Maisonial-besset A, Serre A, Ouadi A, Schmitt S, Canitrot D, Leal F, et al. Base/cryptand/metal-free automated nucleophilic radiofluorination of [18F]FDOPA from iodonium salts: importance of hydrogen carbonate counterion. Eur J Org Chem. 2018;48:7058–65.

    Google Scholar 

  • Makaravage KJ, Brooks AF, Mossine AV, Sanford MS, Scott PJH. Copper-mediated radiofluorination of arylstannanes with [18F]KF. Org Lett. 2016;18:5440–3.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Marchand P, Lorilleux C, Gourand F, Sobrio F, Peyronnet D, Dhilly M, et al. Efficient radiosynthesis of 2-[18F]fluoroadenosine: a new route to 2-[18F]fluoropurine nucleosides. ACS Med Chem Lett. 2010;1:240–3.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Marchand P, Ouadi A, Pellicioli M, Schuler J, Laquerriere P, Boisson F, et al. Automated and efficient radiosynthesis of [18F]FLT using a low amount of precursor. Nucl Med Biol. 2016;43(8):520–7.

    CAS  PubMed  Google Scholar 

  • Martarello L, Schaffrath C, Deng H, Gee AD, Lockhart A, Hagan DO. The first enzymatic method for C–18F bond formation: the synthesis of 5′-[18F]-fluoro-5′-deoxyadenosine for imaging with PET. J Label Compd Radiopharm. 2003;46:1181–9.

    CAS  Google Scholar 

  • Mathiessen B, Zhuravlev F. Automated solid-phase radiofluorination using polymer-supported phosphazenes. Molecules. 2013;18:10531–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Mathiessen B, Jensen ATI, Zhuravlev F. Homogeneous nucleophilic radiofluorination and fluorination with phosphazene hydrofluorides. Chem - A Eur J. 2011;17:7796–805.

    CAS  Google Scholar 

  • McCammant MS, Thompson S, Brooks AF, Krska SW, Scott PJH, Sanford MS. Cu-mediated C-H 18F-fluorination of electron-rich (hetero)arenes. Org Lett. 2017;19(14):3939–42.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Melean JC, Ermert J, Coenen HH. Enantiospecific synthesis of 2-[18F]fluoro-L-phenylalanine and 2-[18F]fluoro-L-tyrosine by isotopic exchange. Org Biomol Chem. 2011;9:765–9.

    Google Scholar 

  • Meleán JC, Ermert J, Coenen HH. A three-step radiosynthesis of 6-[18F]fluoro-L-meta-tyrosine starting with [18F]fluoride. J Label Compd Radiopharm. 2015;58:133–40.

    Google Scholar 

  • Moon BS, Park JH, Lee HJ, Kim JS, Kil HS, Lee BS, et al. Highly efficient production of [18F]fallypride using small amounts of base concentration. Appl Radiat Isot. 2010;68(12):2279–84.

    CAS  PubMed  Google Scholar 

  • Moon BS, Kil HS, Park JH, Kim JS, Park J, Chi DY, et al. Facile aromatic radiofluorination of [18F]flumazenil from diaryliodonium salts with evaluation of their stability and selectivity. Org Biomol Chem. 2011;9:8346–55.

    CAS  PubMed  Google Scholar 

  • Mossine AV, Brooks AF, Makaravage KJ, Miller JM, Ichiishi N, Sanford MS, et al. Synthesis of [18F]Arenes via the copper-mediated [18F]fluorination of Boronic acids. Org Lett. 2015;17(23):5780–3.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Mossine AV, Brooks AF, Ichiishi N, Makaravage KJ, Sanford MS, Scott PJH. Development of customized [18F]fluoride elution techniques for the enhancement of copper-mediated late-stage radiofluorination. Sci Rep. 2017;7(233):1–9.

    CAS  Google Scholar 

  • Mossine AV, Brooks AF, Bernard-Gauthier V, Bailey JJ, Ichiishi N, Schirrmacher R, et al. Automated synthesis of PET radiotracers by copper-mediated 18F‐fluorination of organoborons: importance of the order of addition and competing protodeborylation. J Label Compd Radiopharm. 2018;61:228–36.

    CAS  Google Scholar 

  • Mossine AV, Tanzey SS, Brooks AF, Makaravage KJ, Ichiishi N, Miller JM, et al. One-pot synthesis of high molar activity 6-[18F]fluoro-l-DOPA by Cu-mediated fluorination of a BPin precursor. Org Biomol Chem R Soc Chem. 2019;17(38):8701–5.

    CAS  Google Scholar 

  • Mukherjee J, Yang Z, Das MK, Brown T. Fluorinated benzamide neuroleptics-III. Development of (S)-N-[(1 -allyl-2-pyrrolidinyl)methyl]-5-(3-[18F]fluoropropyl)-2,3-dimethoxybenzamide as an improved dopamine D-2 receptor tracer. Nucl Med Biol. 1995;22(3):283–96.

    CAS  PubMed  Google Scholar 

  • Mulholland GK, Mangner TJ, Jewett DM, Kilbourn MR. Polymer-supported nucleophilic radiolabeling reactions with [18F]fluoride and [11 C]cyanide ion on quaternary ammonium resins. J Label Compd Radiopharm. 1989;1–12:378–80.

    Google Scholar 

  • Nandy S, Rajan MGR, Korde A, Krishnamurthy NV. The possibility of a fully automated procedure for radiosynthesis of fluorine-18-labeled fluoromisonidazole using a simplified single, neutral alumina column purification procedure. Appl Radiat Isot. 2010;68(10):1937–43.

    CAS  PubMed  Google Scholar 

  • Neumann K, Glaspy K, Vavere A, Snyder S, DiMagno S. Radiosynthesis of 4-fluorophenylalanine from a diaryliodonium salt precursor. J Nucl Med. 2013;54(Supplement 2):1069.

    Google Scholar 

  • Neumann CN, Hooker JM, Ritter T. Concerted nucleophilic aromatic substitution with 19F– and 18F–. Nature. 2016;534:369–73.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Nkepang GN, Hedrick AF, Awasthi V, Gali H. Facile synthesis of para-[18F]fluorohippurate via iodonium ylide-mediated radiofluorination for PET renography. Bioorg Med Chem Lett. 2016;26(2):479–83.

    CAS  PubMed  Google Scholar 

  • Oh SJ, Chi DY, Mosdzianowski C, Kim JY, Gil HS, Kang SH, et al. Fully automated synthesis of [18F]fluoromisonidazole using a conventional [18F]FDG module. Nucl Med Biol. 2005;32:899–905.

    CAS  PubMed  Google Scholar 

  • Orlovskaya V, Fedorova O, Nadporojskii M, Krasikova R. A fully automated azeotropic drying free synthesis of O-(2-[18F]fluoroethyl)-L-tyrosine ([18F]FET) using tetrabutylammonium tosylate. Appl Radiat Isot. 2019a;152:135–9.

    CAS  PubMed  Google Scholar 

  • Orlovskaya VV, Modemann DJ, Kuznetsova OF, Fedorova OS, Urusova EA, Kolks N, et al. Alcohol-supported Cu-mediated 18F-fluorination of iodonium salts under “minimalist” conditions. Molecules. 2019b;24(3197):1–17.

    Google Scholar 

  • Parent EE, Dence CS, Sharp TL, Welch MJ, Katzenellenbogen JA. Synthesis and biological evaluation of a fluorine-18-labeled nonsteroidal androgen receptor antagonist, N-(3-[18F]fluoro-4-nitronaphthyl)-cis-5-norbornene-endo-2,3-dicarboxylic imide. Nucl Med Biol. 2006;33:615–24.

    CAS  PubMed  Google Scholar 

  • Pauton M, Gillet R, Aubert C, Bluet G, Gruss-Leleu F, Roy S, et al. The first radiosynthesis of 2-amino-5-[18F]fluoro-pyridines via a “minimalist” radiofluorination/palladium-catalyzed amination sequence from anisyl(2-bromopyridinyl)iodonium triflate. Org Biomol Chem R Soc Chem. 2019;17:6359–63.

    CAS  Google Scholar 

  • Pees A, Sewing C, Vosjan MJWD, Tadino V, Herscheid JDM, Windhorst AD, et al. Fast and reliable generation of [18F]triflyl fluoride, a gaseous [18F]fluoride source. Chem Commun R Soc Chemi. 2018;54(72):10179–82.

    CAS  Google Scholar 

  • Perrio C, Schmitt S, Pla D, Gabbei FP, Chansaenpak K, Mestre-Voegtle B, et al. [18F]-Fluoride capture and release: azeotropic drying free nucleophilic aromatic radiofluorination assisted by a phosphonium borane. Chem Commun. 2017;53:340–3.

    CAS  Google Scholar 

  • Prause M, Niedermoser S, Schirrmacher R, Wängler C, Wängler B. Synthetic approaches towards [18F]fluoro-DOG1, a potential radiotracer for the imaging of gastrointestinal stromal tumors. Tetrahedron Lett. 2018;59(35):3332–5.

    CAS  Google Scholar 

  • Preshlock S, Calderwood S, Verhoog S, Tredwell M, Huiban M, Hienzsch A, et al. Enhanced copper-mediated 18F-fluorination of aryl boronic esters provides eight radiotracers for PET applications. Chem Commun. 2016a;52(54):8361–4.

    CAS  Google Scholar 

  • Preshlock S, Tredwell M, Gouverneur V. 18F-labeling of arenes and heteroarenes for applications in positron emission tomography. Chem Rev. 2016b;116(2):719–66.

    CAS  PubMed  Google Scholar 

  • Qu W, Zha Z, Ploessl K, Lieberman BP, Zhu L, Wise DR, et al. Synthesis of optically pure 4-fluoro-glutamines as potential metabolic imaging agents for tumors. J Am Chem Soc. 2011;133(4):1122–33.

    CAS  PubMed  Google Scholar 

  • Ravert HT, Holt DP, Chen Y, Mease RC, Fan H, Pomper MG, et al. An improved synthesis of the radiolabeled prostate-specific membrane antigen inhibitor, [18F]DCFPyL. J Label Compd Radiopharm. 2016;59:439–50.

    CAS  Google Scholar 

  • Reischl G, Kienzle GJ, Machulla H. Electrochemical radiofluorination: labeling of benzene with [18F]fluoride by nucleophilic substitution. J Radioanal Nucl Chem. 2002;254(2):409–11.

    CAS  Google Scholar 

  • Rensch C, Lindner S, Salvamoser R, Leidner S, Böld C, Samper V, et al. A solvent resistant lab-on-chip platform for radiochemistry applications. Lab Chip. 2014;14:2556–64.

    CAS  PubMed  Google Scholar 

  • Richarz R, Krapf P, Zarrad F, Urusova EA, Neumaier B, Zlatopolskiy BD. Neither azeotropic drying, nor base nor other additives: a minimalist approach to 18F-labeling. Org Biomol Chem R Soc Chem. 2014;12:8094–9.

    CAS  Google Scholar 

  • Rotstein BH, Stephenson NA, Vasdev N, Liang SH. Spirocyclic hypervalent iodine(III)-mediated radiofluorination of non-activated and hindered aromatics. Nat Commun Nat Publ Group. 2014;5(4365):1–7.

    Google Scholar 

  • Rotstein BH, Wang L, Liu RY, Patteson J, Kwan EE, Vasdev N, et al. Mechanistic studies and radiofluorination of structurally diverse pharmaceuticals with spirocyclic iodonium(III) ylides. Chem Sci R Soc Chem. 2016;7:4407–17.

    CAS  Google Scholar 

  • Sadeghi S, Liang V, Cheung S, Woo S, Wu C, Ly J, et al. Reusable electrochemical cell for rapid separation of [18F]fluoride from [18O]water for flow-through synthesis of 18F-labeled tracers. Appl Radiat Isot. 2013;75:85–94.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Saiki H, Iwata R, Nakanishi H, Wong R, Ishikawa Y, Furumoto S, et al. 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(9):1703–8.

    CAS  PubMed  Google Scholar 

  • Saito F, Nagashima Y, Goto A, Iwaki M, Takahashi N, Oka T, et al. Electrochemical transfer of 18F from 18O water to aprotic polar solvent. Appl Radiat Isot. 2007;65:524–7.

    CAS  PubMed  Google Scholar 

  • Salvador B, Luque A, Fernandez-maza L, Corral A, Orta D, Fernández I, et al. Disposable PDMS chip with integrated [18F]fluoride pre-concentration cartridge for radiopharmaceuticals. J Microelectromech Syst. 2017;26(6):1442–8.

    CAS  Google Scholar 

  • Schäfer D, Weiß P, Ermert J, Meleán JC, Zarrad F, Neumaier B. Preparation of no-carrier-added 6-[18F]fluoro-L-tryptophan via Cu-mediated radiofluorination. Eur J Org Chem. 2016;27:4621–8.

    Google Scholar 

  • Seo JW, Lee BS, Lee SJ, Oh SJ, Chi DY. Fast and easy drying method for the preparation of activated [18F]fluoride using polymer cartridge. Bull Korean Chem Soc. 2011;32(1):71–6.

    CAS  Google Scholar 

  • Sergeev ME, Morgia F, Lazari M, Wang CJ, Van Dam RM. Titania-catalyzed radiofluorination of tosylated precursors in highly aqueous medium. J Am Chem Soc. 2015;137:5686–94.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Shamni O, Nebeling B, Grievink H, Mishani E. Fine-tuning of the automated [18F]PSMA‐1007 radiosynthesis. J Label Compd Radiopharm. 2019;62:252–8.

    CAS  Google Scholar 

  • Shi H, Braun A, Wang L, Liang SH, Vasdev N, Ritter T. Synthesis of 18F-difluoromethylarenes from aryl (pseudo) halides. Angew. Chemie Int Ed. 2016;55(36):10786–90.

    CAS  Google Scholar 

  • Shoup TM, Yokell DL, Rice PA, Jackson RN, Livni E, Johnson KA, et al. A concise radiosynthesis of the tau radiopharmaceutical, [18F]T807. J Label Compd Radiopharm. 2013;56:736–40.

    CAS  Google Scholar 

  • Singleton TA, Bdair H, Bailey JJ, Choi S, Aliaga A, Rosa-Neto P, et al. Efficient radiosynthesis and preclinical evaluation of [18F]FOMPyD as a positron emission tomography tracer candidate for TrkB/C receptor imaging. J Label Compd Radiopharm. 2020;63(3):144–50.

    CAS  Google Scholar 

  • Stephenson NA, Holland JP, Kassenbrock A, Yokell DL, Livni E, Liang SH, et al. Iodonium ylide-mediated radiofluorination of 18F-FPEB and validation for human use. J Nucl Med. 2015;56(3):489–92.

    CAS  PubMed  Google Scholar 

  • Strebl MG, Campbell AJ, Zhao W-N, Schroeder FA, Riley MM, Chindavong PS, et al. HDAC6 brain mapping with [18F]Bavarostat enabled by a Ru-mediated deoxyfluorination. ACS Cent Sci. 2017;3:1006–14.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Suehiro M, Vallabhajosula S, Goldsmith SJ, Ballon DJ. Investigation of the role of the base in the synthesis of [18F]FLT. Appl Radiat Isot. 2007;65:1350–8.

    CAS  PubMed  Google Scholar 

  • Tang T, Gill HS, Ogasawara A, Tinianow JN, Vanderbilt AN, Williams S, et al. Preparation and evaluation of L- and D-5-[18F]fluorotryptophan as PET imaging probes for indoleamine and tryptophan 2,3-dioxygenases. Nucl Med Biol. 2017;51:10–7.

    CAS  PubMed  Google Scholar 

  • Toorongian SA, Mulholland GK, Jewett DM, Bachelor MA, Kilbourn MR. Routine production of 2-deoxy-2-[18F]fluoro-D-glucose by direct nucleophilic exchange on a quaternary 4-aminopyridinium resin. Nucl Med Biol. 1990;17(3):273–9.

    CAS  Google Scholar 

  • Tredwell M, Preshlock SM, Taylor NJ, Gruber S, Huiban M, Passchier J, et al. A general copper-mediated nucleophilic 18F fluorination of arenes. Angew Chemie Int Ed. 2014;53(30):7751–5.

    CAS  Google Scholar 

  • Van Der Born D, Pees A, Poot AJ, Orru RVA, Windhorst AD, Vugts DJ. Fluorine-18 labelled building blocks for PET tracer synthesis. Chem Sci Rev R Soc Chem. 2017;46(15):4709–73.

    Google Scholar 

  • Vavere A, Butch E, Shulkin B, Snyder S. Synthesis and preliminary validation of 6-[18F]-Fluorodopamine ([18F]FDA) as a PET tracer for neuroblastoma. J Nucl Med. 2014;55(Supplement 1):1410.

    Google Scholar 

  • Vavere AL, Neumann KD, Butch ER, Hu B, Dimagno SG, Snyder SE. Improved, one-pot synthesis of 6‐[18F]fluorodopamine and quality control testing for use in patients with neuroblastoma. J Label Compd Radiopharm. 2018;61:1069–80.

    CAS  Google Scholar 

  • Wagner FM, Ermert J, Coenen HH. Three-step, ‘“one-pot”’ radiosynthesis of 6-fluoro-3,4-dihydroxy-L-phenylalanine by isotopic exchange. J Nucl Med. 2009;50(10):1724–30.

    CAS  PubMed  Google Scholar 

  • Wang L, Cheng R, Fujinaga M, Yang J, Zhang Y, Hatori A, et al. A facile Radiolabeling of [18F]FDPA via Spirocyclic Iodonium Ylides: preliminary PET Imaging Studies in Preclinical Models of Neuroinflammation. J Med Chem. 2017;60(12):5222–7.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Wang L, Yao S, Tang R, Zhu H, Zhang L, Gong J, et al. A concisely automated synthesis of TSPO radiotracer [18F]FDPA based on spirocyclic iodonium ylide method and validation for human use. J Label Compd Radiopharm. 2020a;63(3):119–28.

    CAS  Google Scholar 

  • Wang M, Glick-Wilson BE, Zheng QH. Fully automated radiosynthesis and quality control of estrogen receptor targeting radiopharmaceutical 16α-[18F]fluoroestradiol ([18F]FES) for human breast cancer imaging. Appl Radiat Isot. 2020b;160:109109.

    CAS  PubMed  Google Scholar 

  • Wessmann SH, Henriksen G, Wester HJ. Cryptate mediated nucleophilic 18F-fluorination without azeotropic drying. NuklearMedizin. 2012;51(1):1–8.

    CAS  PubMed  Google Scholar 

  • Winkler M, Domarkas J, Schweiger LF, Hagan DO. Fluorinase-coupled base swaps: synthesis of [18F]-5′-deoxy-5′-fluorouridines. Angew Chemie Int Ed. 2008;47:10141–3.

    CAS  Google Scholar 

  • Wong R, Iwata R, Saiki H, Furumoto S, Ishikawa Y, Ozeki E. Reactivity of electrochemically concentrated anhydrous [18F]fluoride for microfluidic radiosynthesis of 18F-labeled compounds. Appl Radiat Isot. 2012;70(1):193–9.

    CAS  PubMed  Google Scholar 

  • Wu C, Chen Y, Lin J, Li J, Chen J, Hsieh T, et al. Development of a novel radioligand for imaging 18-kD translocator protein (TSPO) in a rat model of Parkinson’s disease. BMC Med Imaging. 2019;19(78):1–9.

    Google Scholar 

  • Xu P, Zhao D, Berger F, Hamad A, Rickmeier J, Petzold R, et al. Site-selective late-stage aromatic [18F]fluorination via aryl sulfonium salts. Angew Chemie Int Ed. 2020;59(5):1956–60.

    CAS  Google Scholar 

  • Zarrad F, Zlatopolskiy BD, Krapf P, Zischler J, Neumaier B. A practical method for the preparation of 18F-labeled aromatic amino acids from nucleophilic [18F]fluoride and stannyl precursors for electrophilic radiohalogenation. Molecules. 2017;22(2231):1–25.

    Google Scholar 

  • Zhang X, Basuli F, Shi Z, Xu B, Blackman B, Choyke PL, et al. Automated synthesis of [18F](2S,4R)-4-fluoroglutamine on a GE TRACERlab TM FX-N pro module. Appl Radiat Isot. 2016;112:110–4.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Zhang Z, Lau J, Kuo H, Zhang C, Colpo N, Bénard F, et al. Synthesis and evaluation of 18F-labeled CJ-042794 for imaging prostanoid EP4 receptor expression in cancer with positron emission tomography. Bioorg Med Chem Lett. 2017a;27(10):2094–8.

    CAS  PubMed  Google Scholar 

  • Zhang Z, Lau J, Zhang C, Colpo N, Nocentini A, Supuran CT, et al. Design, synthesis and evaluation of 18F-labeled cationic carbonic anhydrase IX inhibitors for PET imaging. J Enzyme Inhib Med Chem. 2017b;32(1):722–30.

    CAS  PubMed  PubMed Central  Google Scholar 

  • Zhang B, Fraser BH, Klenner MA, Chen Z, Liang SH, Massi M, et al. [18F]ethenesulfonyl fluoride as a practical radiofluoride relay reagent. Chem A Eur J. 2019a;25:7613–7.

    CAS  Google Scholar 

  • Zhang X, Basuli F, Swenson RE. An azeotropic drying-free approach for copper‐mediated radiofluorination without addition of base. J Label Compd Radiopharm. 2019b;62:139–45.

    CAS  Google Scholar 

  • Zischler J, Krapf P, Richarz R, Zlatopolskiy BD, Neumaier B. Automated synthesis of 4-[18F]fluoroanisole, [18F]DAA1106 and 4-[18F]FPhe using Cu-mediated radio fluorination under “minimalist” conditions. Appl Radiat Isot. 2016;115:133–7.

    CAS  PubMed  Google Scholar 

  • Zischler J, Kolks N, Modemann D, Neumaier B, Zlatopolskiy BD. Alcohol-enhanced Cu-mediated radiofluorination. Chem A Eur J. 2017;23(14):3251–6.

    CAS  Google Scholar 

  • Zlatopolskiy BD, Zischler J, Krapf P, Zarrad F, Urusova EA, Kordys E, et al. Copper-mediated aromatic radiofluorination revisited: efficient production of PET tracers on a preparative scale. Chem A Eur J. 2015;21(15):5972–9.

    CAS  Google Scholar 

  • Zlatopolskiy BD, Zischler J, Schafer D, Urusova EA, Guliyev M, Bannykh O, et al. Discovery of 7–[18F]fluorotryptophan as a novel positron emission tomography (PET) probe for the visualization of tryptophan metabolism in vivo. J Med Chem. 2018;61:189–206.

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

No funding was received.

Author information

Authors and Affiliations

Authors

Contributions

L.H. wrote the manuscript. D.V. and A.W. read and approved the final manuscript.

Corresponding author

Correspondence to Albert D. Windhorst.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Haveman, L.Y.F., Vugts, D.J. & Windhorst, A.D. State of the art procedures towards reactive [18F]fluoride in PET tracer synthesis. EJNMMI radiopharm. chem. 8, 28 (2023). https://doi.org/10.1186/s41181-023-00203-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s41181-023-00203-5

Keywords