Next Article in Journal
Optimal Tests for Combining p-Values
Previous Article in Journal
Novel Techniques for Microbiological Safety in Meat and Fish Industries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phase Transfer Catalysts and Role of Reaction Environment in Nucleophilc Radiofluorinations in Automated Synthesizers

by
Raisa N. Krasikova
* and
Viktoriya V. Orlovskaya
N.P. Bechtereva Institute of the Human Brain, Russian Academy of Sciences, 197376 Saint-Petersburg, Russia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(1), 321; https://doi.org/10.3390/app12010321
Submission received: 18 November 2021 / Revised: 28 December 2021 / Accepted: 28 December 2021 / Published: 29 December 2021
(This article belongs to the Special Issue Next Generation Radiopharmaceuticals for PET-Based Diagnostics)

Abstract

:
Incorporation of [18F]fluorine into PET radiotracer structure has traditionally been accomplished via nucleophilic pathways. The [18F]fluoride is generated in an aqueous solution via proton irradiation of oxygen-18 enriched water and must to be introduced into water-free organic solutions in order to generate reactive species. Thus nucleophilic 18F-fluorination traditionally included steps for [18F]fluoride concentration on the anion exchange resin, followed by removal of residual water via azeotropic distillation with MeCN, a time-consuming process associated with radioactivity losses and difficult automation. To circumvent this, several adsorption/elution protocols were developed based on the minimization of water content in traditional kryptofix-based [18F]fluoride eluents. The use of pre-dried KOH/kryptofix solutions, tertiary alcohols, and strong organic bases was found to be effective. Advances in transition metal-mediated SNAr approaches for radiolabeling of non-activated aromatic substrates have prompted development of alternative techniques for reactive [18F]fluoride species generation, such as organic solutions of non-basic alkyl ammonium and pyridinium sulfonates, etc. For radiofluorinations of iodonium salts precursors, a “minimalist” approach was introduced, avoiding the majority of pitfalls common to more complex methods. These innovations allowed the development of new time-efficient and convenient work-up procedures that are easily implementable in modern automated synthesizers. They will be the subject of this review.

1. Introduction

Positron Emission Tomography (PET) is an important molecular imaging modality that has long been used for clinical diagnostics, particularly in the field of oncology. It allows quantification of physiological and pathological processes in living subjects on a molecular level through detection of radiotracer—biologically active or drug-like molecule labeled with positron-emitting radionuclides—biodistribution. Among cyclotron-produced radionuclides, 11C, 13N, and 15O, the 18F possesses the longest half-life (109.8 min) and favorable physicochemical and nuclear properties. Low positron energy of 0.635 MeV allows for highest theoretical spatial resolution due to the shortest maximum range of positrons in tissue (2.4 mm in water). Over 300 GBq per batch of the radionuclide can be easily produced in a low-to-medium energy cyclotron via the 18O(p,n)18F reaction occurring during proton irradiation of 18O-enriched water in a high pressure target [1]. With this production approach, starting [18F]fluoride and corresponding labeled tracers are available in high molar activity (Am) allowing for microdosing approach in receptors-binding studies. Ca. 2-h half-life permits complex synthetic approaches, providing access to a large variety of 18F-labeled radiopharmaceuticals (RPs) and also allows shipping of multiple clinical doses from a centralized production site to hospitals without access to cyclotron and radiochemistry/radiopharmacy facilities.
[18F]fluoride is currently the most practical and generally available source of this radionuclide, and in this form, it is used in the vast majority of the 18F-labeling procedures. Nucleophilic substitution of an appropriate leaving group in a precursor molecule with [18F]fluoride is the most prominent method for the production of a variety of clinically relevant radiotracers [2,3,4,5]. [18F]fluoride generated in a water target as strongly hydrated anion is generally unreactive in nucleophilic substitution scenarios. Therefore, the pre-processing steps are required to isolate [18F]F from an aqueous environment and transform it into a reactive nucleophile species.
Early work on the nucleophilic 18F-fluorinations included the removal of the bulk target water (1–3 mL volumes) by thermal distillation in the presence of K2CO3, added to prevent the formation of gaseous [18F]HF. Following it up with azeotropic distillation with acetonitrile allowed for complete removal of residual water, however, the whole procedure was lengthy and generally impractical. Alternatively, separation of the [18F]fluoride from [18O]H2O (allowing for recycling of enriched water) on a strong anion-exchange cartridge (AEC) was introduced [6] and was found to be ideally suited for processing 18F with implementation in the commercial automated synthesizers following shortly [7,8,9]. The recovery process is composed of two steps: the trapping of radionuclide on the resin and elution of adsorbed radioactivity with an organic–aqueous solution containing a base and a phase-transfer catalyst (PTC), most commonly the combination of K2CO3 and a cryptand (Kryptofix, K2.2.2, Sigma-Aldrich, St. Louis, Missouri, United States). The addition of cryptand enables solubilization of the [18F]fluoride in a polar aprotic solvent and enhances nucleophilicity of the fluoride ion. The first application of K2.2.2. as cryptand in a 18F-radiolabeling was reported in 1986 for aliphatic nucleophilic synthesis of a fatty acid [10]; soon after, it was employed in the synthesis of 2-deoxy-2-[18F]fluoro-D-glucose ([18F]FDG) [11], the most commonly used fluorine-18 radiotracer. Following this formative work, a radiolabeling method using azeotropically dried [18F]KF/K2.2.2./K2CO3 was used in the preparation of a great variety of radiotracers for clinical and research applications in both aliphatic and aromatic nucleophilic substitution scenarios [2,3,4,5,12]. As an alternative to K2.2.2./K2CO3(KHCO3), tetrabutyl ammonium bicarbonate (Bu4NHCO3) was introduced at about the same time [13] for solubilization of [18F]fluoride, and, due to its lower basicity, was found to be effective in a number of nucleophilic fluorinations.
Since the presence of water severely reduces the reactivity of the fluoride ion in nucleophilic substitution reactions, trace water from the eluent mixture need to be carefully removed. Customarily this is achieved through a multi-step azeotropic drying process with the addition of several aliquots of acetonitrile. Although this procedure was implemented in most of the currently available commercial synthesis modules, the efficiency of drying depends on the reactor vessel design, heat transfer, vacuum/drying gas flow, and time. Incomplete drying strongly (and negatively) affects the reactivity of the [18F]fluoride, drastically reducing yields. As a whole, the azeotropic drying step is time-consuming (6–15 min) and is usually associated with significant losses of radioactivity due to decay and absorption of the radionuclide on the inner surfaces of the reaction vessel.
Another limitation of this approach is the use of the substantial amounts of the base to facilitate elution of [18F]fluoride retained on the standard large-size anion exchange cartridges (Sep-Pak Plus QMA Light, PSHCO3). A typical kryptofix-based eluent contains 2.0–3.5 mg (15–25 μmol) of K2CO3, which is an unacceptably large amount when dealing with base-sensitive precursors. Resulting basic reaction conditions often lead to the formation of undesired radiolabeled by-products and the generation of chemical impurities resulting from the decomposition of the precursor.
In order to address these issues, considerable efforts have been expended to develop shortened [18F]F adsorbtion/elution protocols omitting the azeotropic drying procedure and, more importantly, providing less basic and milder 18F-fluorination conditions for base-sensitive precursors. A number of innovative methods and techniques were introduced by different groups through optimization of the base and phase-transfer agent used, eluting solvent composition, cartridge conditioning solvent (anion), and other parameters. As a result, a number of fine-tuned protocols for the preparation of particular radiotracers have been developed and published.
In this review, our focus is on the improvement of “classical” methods for the activation of [18F]fluoride facilitating their implementation into the routine automated production of radiotracers, as well as emerging methodologies and techniques. One needs to acknowledge that in recent years a number of microfluidic and microscale approaches have been introduced for small-scale production. However, they, as well as electrochemical methods used for reprocessing of [18F]fluoride, lie outside of the scope of the current review.

2. [18F]Fluoride Activation for Aliphatic Nucleophilic SN2 Substitution

Aliphatic nucleophilic SN2 substitution in the presence of K2.2.2./K2CO3 or Bu4NHCO3 is currently the most prominent method for the production of the majority of clinically relevant radiotracers, including [18F]FDG, 3-deoxy-3-[18F]fluorothymidine ([18F]FLT), [18F]fluoromisonidazole (1-[18F]fluoro-3-(2-nitroimidazol-1-yl)-propan-2-ol; [18F]FMISO), O-(2′-[18F]fluoroethyl-L-tyrosine ([18F]FET), 16α-[18F]fluoroestradiol ([18F]FES) (Figure 1), and others [2,5,6,7,8,9,10,11,12].
Typical one-vessel two-step synthesis of a radiotracer is achieved by nucleophilic displacement of an appropriate leaving group in the precursor with activated [18F]F in polar aprotic solvents (MeCN, DMF, DMSO, etc.) at 80–130 °C followed by (if necessary) removal of the protective groups. This methodology was also adapted to the labeling of a number of 18F-labeled bioactive and drug-like molecules prepared via direct radiofluorination, as well as for the preparation of the most common 18F-fluoroalkylating agents [14] and prosthetic groups [15] for indirect radiolabeling approaches. Despite the fact that conventional azeotropic-drying methodologies are well established in the aliphatic fluorinations and are implemented in the majority of automated synthesizers, the drying-free approaches have shown distinct advantages such simplicity of the automation and shorter production time.
Traditionally the methods to circumvent the drying process relied upon the variation of the composition of kryptofix-based AEC eluent containing either K2CO3 or less basic KHCO3, K2C2O4, and reduction of water fraction in the organic/water mixture. For base-sensitive precursors encouraging results were achieved using chemically inert quaternary ammonium salts such as Bu4NOTs, Bu4NOTf, and Et4NOTf that have been introduced recently in lieu of “classical” Bu4NHCO3.

2.1. Azeotropic Drying Free Methods for Kryptofix-Mediated Radiofluorinations

In the majority of protocols for kryptofix-mediated fluorinations the AEC resin-bound [18F]F is eluted from the cartridge with a solution of K2.2.2./K2CO3 dissolved in a mixture of organic solvent and water.
The type of anion-exchange cartridge and the amount of packing material substantially influence the efficacy of [18F]F elution as well as the subsequent radiolabeling. Currently, the capture-and-release process is usually accomplished using standard AccellTM Plus QMA light Sep-Pak cartridge (filled with 130 mg of quaternary ammonium chloride polymer) or less common Chromafix® PS-HCO3 (filled with 46 mg of quaternary ammonium bicarbonate polymer). Before use, the QMA*Cl cartridge is conditioned with aqueous K2CO3 or KHCO3 and water, while the PSHCO3 cartridge is just wetted with water. The QMA resin is also available in a carbonate form, the Sep-Pak Light (46 mg) AccelTM Plus QMA Carbonate cartridge that can be used as is, without any conditioning. These cartridges are used in all types of automated synthesizers as components of pre-assembled disposable cassettes (GE FASTlab) or the appropriate reagents kits (IBA Synthera). Barrel-type cartridges such as Vac QMA 1 cc (100 mg), Oasis WAX 1 cc Vac Cartridge (30 mg), StrataTM-XL-A polymeric strong anion, 100 mg/3 mL, and others are commonly encountered in research and development situations but are less common in routine applications.
Choosing the right eluent solvent mixture can significantly reduce drying time later on. A 100% aqueous eluent employed in the first published methods provided 100% elution efficiency (EE). However, the azeotropic drying time was too long for the method to be of significant practical use. In the current practice, the acetonitrile-water mixtures with different percentages of water are employed to shorten the drying procedure. Using eluent that contains only 4% of water and 96% of acetonitrile allowed for the dried reactive [K/K2.2.2.]+[18F] complex to be obtained without further addition of acetonitrile. According to this protocol, the radionuclide is trapped on a standard strong anion-exchange cartridge AccellTM Plus QMA light Sep-Pak (130 mg) pre-conditioned with 10 mL of 0.5 M K2CO3 solution and water, each. Trapped [18F]fluoride is eluted with 2 mL of MeCN/H2O (96/4 v/v) mixture containing 9.8 mg (25 µmol) of K2.2.2. and 1.8 mg (12 µmol) of K2CO3, with EE >98%. Evaporation of solvents at 110 °C under a stream of nitrogen (with no vacuum) produces reactive [18F]fluoride directly. The general applicability of this methodology, originally developed by our group for automated synthesis of [18F]FDG [16,17], was confirmed by applying it to the routine preparations of the tumor imaging agents [18F]FMeChol [18], [18F]FES [19], [18F]FEDBT [20]. Numerous other examples of this method’s application to production of other radiotracers were reported in the literature: [18F]Flumazenil [21], [18F]FEPE2I [22], [18F]FP-06684511 [23], [18F]MCL524 [24], and others [Figure 2].
Pascali and co-authors [25,26] have reported an azeotropic-drying free method using QMA eluent containing 12% of water in acetonitrile with less basic KHCO3. The [18F]F retained on the cartridge (QMA, 130 mg) was eluted with 1 mL of K2.2.2./KHCO3 solution; with vacuum-assisted evaporation to dryness at 72 °C under a nitrogen flow following, affording reactive [18F]fluoride species. Using this protocol, two highly-sought radiotracers, [18F]FET [25] and [18F]FMISO [26] were synthesized. The latter one was obtained in greater radiochemical yield (RCY) (54% against reported 30–40% for conventional methods), with only 30 min total synthesis time.
However, even as techniques described [16,17,18,19,20,21,22,23,24,25,26] avoid azeotropic drying procedure, the eluting solvent must still be removed by distillation to be replaced with the reaction solvent. There has been a recent focus on developing methodologies with the aim of completely excluding the need for solvent replacement prior to radiofluorination step.
Wester and co-workers [27] described an approach using highly concentrated K2.2.2./KOH solution in anhydrous MeCN, allowing for quantitative elution of [18F]F under water-free conditions. No solvent evaporation was required to obtain reactive [18F]fluoride species. The applicability of this methodology to the preparation of several well-known radiotracers ([18F]FDG, [18F]FET, [18F]FLT, [18F]FAZA, and [18F]fallypride) was confirmed when radiolabeling was carried out using small aliquots of the eluate. An example of application of this so-called “Munich method” in a full-batch production is the synthesis of 18F-rhPSMA-7 [28], a novel hybrid PSMA radiloligand that contains either 18F or 68Ga in the same structure. In a manual synthesis [28] [18F]F was eluted from the QMA cartridge (46 mg) with a mixture of K2.2.2. (110 μmol) and KOH (100 μmol) in 0.5 mL of MeCN; the eluent was freshly prepared from the lyophilized kit. The eluate was collected directly into a reaction vessel containing labeling precursor (50–150 nmol) and an oxalic acid (to provide less basic conditions [29]) in 150 μL DMSO/30 μL MeCN mixture. Isotopic 19F/18F exchange reaction was completed in 5 min at r.t. providing an RCY of over 55% after cartridge-based purification. Despite the distinct advantages of this time-saving and easy-to-automate protocol, it has not yet been implemented into routine 18F-radiochemistry, possibly due to the need for high amounts of base and subsequently relatively large amount of precursors required.
Very recently [30] use of water-free eluent containing KHCO3 (1.6 mg) and K2.2.2. (6.0 mg) in MeOH (0.4 mL) for the recovery of [18F]fluoride from the QMA cartridge (46 mg, carbonate form) followed by direct use of the eluate in labeling reaction has been reported. This methodology was previously developed [30] for microscale 18F-radiolabeling with minimization of the amounts of base and cryptand. Synthesis of a 18F-labeled stilbene [30] for β-amyloid imaging was achieved using 2 mg of precursor in 0.2 mL of DMSO. When employing conventional azeotropic drying methodology for the same synthesis, a substantial amount of the unlabeled impurities in the reaction mixture was observed. The authors speculate that heating acetonitrile in the presence of K2CO3 and K2.2.2. promoted dimerization of acetonitrile, leading to the formation of butyronitrile. The latter then reacted with the sulfonyl leaving group of the precursor producing unlabeled impurities. In contrast, no impurities were observed using the water-free methanolic solution in the elution stage with no need for azeotropic drying. Still, in this approach, the eluting solvent needs to be removed before fluorination. Noteworthy, this is the first report pointing out the possibility of MeCN dimerization under azeotropic drying conditions. Small volumes of the solvents employed in the method [31] might be a limitation of this approach’s amenability to automation due to possible losses of the radioactivity in the transfer lines.
A relatively straightforward method [32] avoiding the need to remove the water is a kryptofix-mediated radiofluorination in a mixed organic solvent system containing only a small amount of water. Good tolerance to the presence of trace water was demonstrated in the preparation of three radiotracers, [18F]PSS232, [18F]MISO, and [18F]fallypride when using [18F]F eluent containing K2.2.2. (7.2 mg, 19 μmol) and K2CO3 (1.3 mg, 9.6 μmol) in 600 μL of MeCN-H2O (49:1 v/v). [18F]Fluoride was recovered from 46 mg QMA cartridge (EE 86%) directly into a solution of the precursor in DMA; fluorination was carried out in DMA/MeCN (70/30) mixture containing 97 µL of water. This operationally simple methodology was easily implemented on a GE TRACERlab FX FN platform (GE Healthcare, Madison, WI, USA) exemplified by the automated production of [18F]fallypride, a clinically useful D2/D3 antagonist radioligand with a moderate RCY of 25%.
In another report [33], eluate containing a certain amount of water (4850 ppm, eluted volume 1 mL) was used directly in the radiolabeling of several aliphatic and aromatic precursors, including mannose triflate ([18F]FDG precursor). To trap cyclotron-produced [18F]fluoride a custom-prepared water-wettable macroporous co-polymer loaded with a long alkyl chain quaternary ammonium salt was used instead of the conventional QMA cartridge. Limitations of this methodology are the need for high precursor amounts (15–40 mg) and the necessity to add the additional PTCs, either K2CO3/K2.2.2. (7 mg/22 mg) or Et4NHCO3 (50 mg) to increase the fluorination efficiency. Noteworthy, since only an aliquot of the eluate was used in the radiofluorinations, the reaction conditions were effectively anhydrous. Due to the exclusion of the evaporation steps, this approach appears to be attractive for deployment in microfluidic systems, which was one of the goals of this study.
A successful aliphatic substitution reaction with [18F]fluoride in the presence of 2–3% of water (12.000–20.000 ppm) was demonstrated by Kniess and co-workers [34] in the azeotropic-drying free synthesis of 2-[18F]fluoroethyl tosylate, a frequently used 18F-fluorialkylating agent. With this so-called “hydrous fluoroethylation” protocol, [18F]F loaded onto the QMA cartridge was eluted with 0.7 mL of solution containing 21 µmol K2CO3 and 42 µmol K2.2.2.; the collected eluate was used directly in reaction with 1,2-ethylene glycol-bistosylate (7 mg, 19 µmol) with further one-pot 18F-fluoroalkylation reaction promoted by Cs2CO3 as a base. Avoiding any solvent removal steps has resulted in five-fold increase in the isolated RCY of three model COX-2 inhibitors not only due to shortened synthesis procedure but also due to minimal losses of the radioactivity on the inner walls of the reaction vessel.

2.2. Tetraalkylammonium Salts (Bu4N+, Et4N+) as the PTCs

Tetraalkylammonium salts together with basic, non-nucleophilic anions (e.g., CO32− or HCO3) have been used for the years [13,34] with Bu4NHCO3 as less basic alternatives to K2.2.2./K2CO3. They are characterized by counter-ion exchanging ability with [18F]fluoride and good solubility in most organic solvents. From this group of the PTCs, Bu4NHCO3 has been the most widely used, demonstrating higher [35,36] or similar [37] RCYs when compared to conventional kryptofix-mediated fluorinations. The [18F]F is quantitatively recovered from the AEC with the eluent composed of MeCN (2 mL) and Bu4NHCO3 (0.075 mL, 20% aq., pH 8) [38,39] followed by conventional azeotropic drying. Quite recently [38], an elution protocol using ethanolic solution of the PTC (0.075 M Bu4NHCO3 in 0.5 mL of EtOH) was described. The reactive [18F]fluoride species thus produced was further used for the synthesis of [18F]FES immediately after simple solvent evaporation without the need for azeotropic drying. For this radiotracer similar RCYs of around 50% (decay-corrected) were obtained regardless of the choice of the PTC—K2.2.2./K2CO3 or Bu4NHCO3.
Similar to Bu4NHCO3, tetraethylammonium salts were found to be effective in enhancing [18F]fluoride reactivity in both aliphatic and aromatic nucleophilic substitution scenarios. Brichard and Aigbirhio [39] demonstrated that Et4NHCO3 (15 mg) dissolved in a polar aprotic solvent (MeCN, DMF, DMSO) containing up to 5% of water could be used to elute QMA-loaded [18F]F with elution performed in the reverse (opposite to loading) direction. The obtained eluate can be used directly in further fluorinations eliminating solvents evaporation steps. The feasibility of this approach was confirmed for a wide range of the aliphatic substrates ([18F]FDG, [18F]fallypride, [18F]FEDAA1106) and model aromatic compound demonstrating similar or increased RCY when compared with conventional kryptofix-mediated procedures. However, it is worth noting that all the labeling reactions were performed using small aliquots of the eluate (50–100 µL). Thus, the applicability of this methodology for full-batch production needs to be evaluated further, which reduces both the amount of base and the amount of water in the reaction mixture. Thus the applicability of this methodology for full-batch production needs to be evaluated further.
In general, in routine clinical application use of Et4NHCO3 is not as common as Bu4NHCO3. In one of the recent protocols [40] that has been implemented into GMP production of [18F]DPA-714, preparation of reactive Et4N+[18F]F still incorporated steps for “traditional” solvent removal. Automated synthesis on a Trasis AllinOne (AIO) cassette-based module involved elution of the [18F]fluoride from the QMA cartridge with a solution of Et4NHCO3 in water/acetonitrile (85/15 v/v), followed by azeotropic distillation. Fluorination of the corresponding tosyl precursor in MeCN provided significantly higher RCY (55–71%) if compared with earlier kryptofix-based methods. The authors speculate that Et4NHCO3 could be a better enhancer of [18F]fluoride nucleophilicity than K2.2.2./K2CO3 or K2.2.2./KHCO3, probably due to better stabilization of the precursor-fluoride intermediate during the reaction.
Apart from developments concerning the elimination of the azeotropic drying step and improvements in automation, considerable efforts have been directed at the control of the basicity of the reaction medium, with aim to reduce amount of precursors used in the reaction. Aliphatic radiofluorination of base-sensitive precursors has received much attention in particular. A recent review [41] summarises various new developments in this field. In particular, success [42,43,44] has been achieved by exploiting tetraalkylammonium salts in conjunctions with less basic anions such as mesylate, tosylate, and triflate. The first application of these quaternary ammonium salts (Bu4NOMs and Bu4NOTf) dates back to 2011 [45]. It was shown that the solution of Bu4NOMs in methanol (0.6 mL, 10 mg (30 μmol)) could release over 90% of the radioactivity from the AEC (Chromafix® PS-HCO3) and provide over 90% RCC in the radiofluorination of model aliphatic substrates when carried out in tertiary alcohols. The effects induced by conditioning anion replacing bicarbonate ion in the PSHCO3 resin with other anionic bases such as Na2CO3, K2HPO4, and K3PO4 was also examined in that study. It turned out that conditioning the cartridge with phosphate-based anions (KHPO4 and K2PO4) did not only provide the highest RCC in radiofluorination of aliphatic substrates but contributed significantly to creating a mild reaction environment with unreacted precursor remaining almost intact.
Although that study demonstrated distinct advantages of Bu4NOMs as a neutral PTC, results obtained in fluorination of model compounds were not extended further to the preparation of clinically relevant radiopharmaceuticals. Thus with the aim of wider implementation of this concept our group has recently [42] suggested a general protocol where the use of the corresponding tosylate salt, Bu4NOTs, was adapted to the routine automated synthesis of a series of 18F-labeled radiotracers [18F]FDG, [18F]FLT, [18F]FMISO, [18F]FET, [18F]FES as well as common radiolabeling synthon, 2-[18F]fluoroethyl tosylate ([18F]FEOTs). The radionuclide initially retained on a QMA cartridge (conditioned with an aq. KHCO3) was recovered in a solution of Bu4NOTs (1 mg, 2.5 µmol) in MeOH or EtOH. Following solvent removal, reactive [18F]fluoride was reconstituted in 1 mL of MeCN. In order to achieve release of [18F]F from conventional 130 mg QMA cartridge while using such a low amount of the PTC, a reversed loading/elution procedure was employed of the type that has been demonstrated in radiofluorination of electron-rich arenes [46]. However, in conjunction with a smaller QMA Light carbonate cartridge (46 mg) a straightforward eluting protocol was found to be just as effective in providing high EE using the same amount of Bu4NOTs dissolved in 2 mL of EtOH. The radiofluorinations were carried out in MeCN, a solvent commonly employed for the aliphatic fluorinations, under typical reaction conditions (100 °C, 10 min). High RCCs (70–95%) were achieved using significantly reduced amounts (0.2–1.2 mg/0.5–1.2 µmol) (Table 1) of precursors compared to traditional approaches. In addition, by reducing the precursor amounts, the sources of chemical impurities in the final preparations are minimized, allowing for simplification of the purification procedures. The practical utility of the method (standard elution protocol, 46 mg QMA cartridge) was demonstrated in automated synthesis of [18F]FET [47], [18F]FES [48], and [18F]FDG [42] on the GE TRACERlab FX N Pro module, one of most common 18F-labeling platform.
To further explore the potential of tetraalkylammonium salts in respect to reaction basicity and resolubilization of the [18F]fluoride, effects of three different salts Bu4NOMs, Bu4NOTf, and Bu4NH2PO4 were recently investigated by Bratteby and co-workers [43]. In model experiments, the key parameters evaluated were trapping of [18F]fluoride on the standard QMA cartridge, its subsequent elution, and the effects of different conditioning regiments, which led to the identification of a number of relevant factors affecting RCC of the 18Ffluorination of base-sensitive precursors. Through testing of a great variety of combinations for conditioning and elution anions, suitable general protocols applicable to the synthesis of currently used radiotracers were developed.
In the following publication by the same group [49], the radiolabeling protocol for [18F]FEE2I, a well-established dopamine transporter radioligand, was developed and implemented in a cassette-based Synthera® + HPLC (IBA) synthesizer. Despite the fact that a previous study [43] revealed the highest RCC (72.5 ± 13.1%) for [18F]FEE2I was achieved in the presence of Bu4NOMs, due to the potential toxicity of mesylate ions, the Bu4NH2PO4 was chosen for the clinical production implementation. For elution procedure, 20 µmol of this PTC in MeCN/H2O (50:50) was applied to the standard QMA cartridge pre-conditioned with CO32−. The 18F eluate was then concentrated to dryness via three short cycles of combined application of N2 flow and vacuum, resulting in the complete removal of the eluting solvent. Using DMSO as the reaction solvent, the product was obtained with the isolated RCY of 33 ± 2% (n = 3) from the 45 GBq starting activity. Due to radiolysis, the synthesis with higher starting activities (up to 140 GBq) proceeded with the decrease in RCY even under optimal conditions.
In line with these studies [42,43], Inkster and co-workers [44] recently employed water-organic solutions (up to 50% of water) of tetraethyl ammonium salts (Et4NOTs, Et4NClO4) to recover [18F]F that was trapped on an miniature MP-1 anion-exchange column or a commercial QMA anion-exchange column (both in carbonate form, 10–12 mg; MedChem Imaging, LLC, Boston, United States), commonly used in microfluidic synthesis. The elution was performed in reverse mode; the eluate was directly used for 18F-substitution reactions, i.e., without any evaporation steps. The low volume of the concentrated eluent (Et4NOTs, 23.5 mg/mL, 100 µL in 7:3 MeCN:H2O) was sufficient to release radioactivity from the AEC. The eluate was mixed with the precursor solution in 900 µL of MeCN with a final reaction volume equal to 1 mL with 3% water. The radiofluorinations were carried out with microwave heating (150 °C, 10 min), showing high RCC in the synthesis of [18F]fallypride, a widely used D2/D3 antagonist tracer, on a sub-hundred-MBq scale. More recently [50] this protocol was employed in a small scale (190–225 MBq of [18F]F) manual synthesis of [18F]MCL-524, a D2 agonist moderately selective for the high-affinity state of the D2 receptor. Despite the RCC of radiofluorination of the precursor being high (ca. 50%), due to lengthy procedure of deprotection with In(OTf)3 and HPLC purification, the RCY of 5–9% (decay corrected) was obtained (with synthesis time of 146–199 min) providing enough [18F]MCL-524 for preclinical imaging applications.
This proof-on-concept study supports the conjecture that non-basic tetraalkylammonium salts could serve as the PTCs in non-anhydrous solvent mixtures where no solvent removal steps would be required to produce reactive [18F]F species. However, to serve as a useful production method, the manual protocol [49,50] developed with its use of miniature anion-exchange columns, small eluent volumes as well as microwave heating, needs to be adapted to commercial synthesizers with conventional heaters and employing “normal size” AEC for scalable radionuclide processing.

2.3. PTC Free Cartridge-Based Radiofluorination of Aliphatic Substrates

In principle, immobilizing precursor on the solid-phase support to perform an online 18F-fluorination avoiding [18F]F elution, azeotropic evaporation, and PTC additions is a hugely attractive concept. It was successfully applied as early as 1990 [51] in the synthesis of [18F]FDG via resin-based nucleophilic 18F-fluorination of mannose triflate, using a quaternary 4-(N,N-dialkylamino)-pyridinium functionalized polystyrene anion-exchange resin (“4-AP” resin) as solid support. [18F]fluoride was captured on a small resin-packed column. The column was then dried by rinsing it with anhydrous MeCN, and the solution of mannose triflate was slowly pushed through the column heated to 100 °C. Labeled intermediate was then eluted into reaction vessel for removal of acetyl protecting groups through acidic hydrolysis. This procedure was implemented in the first cassette-based synthesizer (FDG Microlab, GE Medical Systems, Uppsala, Sweden) for routine productions of [18F]FDG [52]. Due to the issues with the manually-packed “4-AP” resin columns that had to be stored in methanol and other technical issues, the manufacture, and supply of the cassettes and FDG Microlab modules were seized around 2005.
Alternatively, [53] the use of commercially available polymer-supported phosphazene base PS-P2tBu was found to be effective for resin-based 18F-fluorinations of a series of aliphatic substrates. The phosphazene super bases such as P2Et were earlier introduced by Lemaire and co-workers [54] for elution of [18F]F trapped on the QMA cartridge using acetonitrile containing carefully controlled amounts of water (1 mL MeCN, 5–10,000 ppm of water, 45 µmmol of base). The highly reactive [18F]fluoride was obtained without the need of azeotropic drying procedure and effectively used in radiolabeling of the appropriate labeling precursor (up to 40 mg) in a solution containing BTMG (2-tert-butyl-1,1,3,3-tetramethylguanidine) as an additional base.
For resin-supported fluorination [53], a glass column or PTFE tube filled with 100 µmol of PS-P2tBu and 1200 mg of glass beads were employed for trapping [18F]F with 98% efficiency. Precursor solution (i.e., mannose triflate, 25 mg in 3 mL of MeCN) was then passed through the column heated to 85 °C at a flow of 0.55 mL/min. The subsequent online basis hydrolysis on conventional tC18 Sep-Pak cartridge afforded [18F]FDG in 40% RCY in automated synthesis, with starting activity of 120 GBq. However, the success of the method depends to a great extent on the amount of resin, reaction temperature, and column packing efficiency, which, taken together, are considered a major limitation of this elegant but technically demanding production method.
The aforementioned methodologies address the synthesis of radiotracers via direct nucleophilic substitution of the appropriate leaving groups in aliphatic substrates with activated [18F]fluoride as the most common route to routinely used radiotracers. The methods have been evaluated in view of possibilities for their implementation in the automated synthesizers with the aim of production of clinically useful amounts of the radiotracers. In recent years a number of highly innovative strategies to introduce the 18F-label into aliphatic structures have been introduced, unfortunately, few of them are currently considered to be suitable for routine preparations of radiotracers. There have been several in-depth reviews [3,41,55,56,57] that offer a good overview of the current state of the art in aliphatic radiofluorinations, and deeper discussion of those more exotic approaches really lies outside the scope of the current review.

3. [18F]Fluoride Activation for Conventional Aromatic Nucleophilic Substitution

Traditional nucleophilic aromatic substitution (SNAr) reactions, including those using [18F]fluoride as the nucleophile, require arene substrates with strong electron-withdrawing groups (EWG) such as NO2, CN, CHO, COR, COOR, and CF3 in the ortho- or para-positions of the aromatic structure combined with an appropriate leaving group (NO2, halides, Me3N+X; X = TfO, TsO, ClO4, I). Radiofluorinations are usually performed in polar, anhydrous aprotic solvents, such as DMSO, DMF, DMA, or sulfolane at moderate to high temperatures (100–180 °C) in the presence K2.2.2./K2CO3 (KHCO3) or Bu4NHCO3. A direct labeling strategy is an effective approach for the synthesis of 18F-labeled benzaldehydes or other prosthetic groups [15] obtained through radiolabeling of substituted nitrobenzaldehydes. When it comes to the direct introduction of [18F]F into biologically active molecules or drug candidates, they usually do not contain a suitably activated aryl ring for fluorination through the addition–elimination mechanism. Notable radiotracers examples accessible via direct nucleophilic substitution include (imaging target is in parenthesis): [18F]altanserin [58,59] (5-HT2A receptors); [18F]MPPF [60] (5-HT1A receptors); [18F]flumazenil [21] (GABAA benzodiazepine receptors); [18F]FPEB [61] (metabotropic glutamate 5 receptors) (Figure 3).
These radiotracers are usually prepared through classical kryptofix-mediated substitution of a nitro leaving group in low to moderate RCYs; microwave-assisted heating has been used to improve yields [58]. In the case of heteroaromatic substitutions, particularly in the pyridine series, direct SNAr fluorination on the ortho- or para-positions is feasible with no need for activation of the heteroaromatic ring. Over the years, a number of radiotracers for nicotine acetylcholinic receptors [62,63] as well as for imaging β-amyloid aggregates [64] and other targets have been prepared from the appropriate heteroaromatic precursors.
A recent example is the synthesis of [18F]PSMA-1007, a widely-used radioligand for the diagnosis and non-invasive staging of PSMA-positive prostate cancer developed by Kopka’s group [65]. The radiotracer has been synthesized in high RCY (50–70%) from its trimethylammonium salt precursor (1–2 mg) in the presence of Bu4NHCO3 with heating to 80–90 °C for 10 min in DMSO. To produce reactive [18F]fluoride species, the radionuclide was released from the QMA cartridge (130 mg) with 750 µL 0.075 M aqueous TBAHCO3 followed by an azeotropic drying procedure. In general, aryltrimethylammonium salts tend to be more reactive than neutral leaving groups. An additional benefit is that the charged precursor is easily separated from a neutral end-product using cartridge-based purification procedures. Thus, the synthesis of [18F]PSMA-1007 with SPE purification can be implemented in the majority of currently available automated synthesizers [65] allowing for easy access to this important PSMA imaging agent.
The second strategy [65] to introduce 18F into PSMA-1007 relies on using 6-[18F]fluoronicotinic acid 2,3,5,6-tetrafluorophenyl ester ([18F]F-Py-TFP), a common prosthetic reagent for the indirect peptides labeling. The unique feature of [18F]F-Py-TFP introduced in 2010 by Oberg and co-workers [66] is that the fluorination of the quaternary ammonium salt precursor proceeds readily at 40 °C with RCC over 65% (10 min heating in tBuOH:MeCN (8:2)) or even at room temperature in cartridge supported [67] radiofluorination approach. The later protocol included elution of the radionuclide retained on pre-wetted Chromafix 30-PS-HCO3 with quaternary ammonium triflate precursor (5–7 mg, in 0.5 mL 1:4 MeCN:tBuOH). The eluate was passed through Oasis MCX Plus cartridge, from which the product was eluted with 1 mL acetonitrile. The study in question presents a very effective use of so-called “minimalist” approach proposed by the Neumaier group [68], based on the direct elution of [18F]F from the AECs using solutions of the precursors bearing quaternary ammonium, diaryliodonium, or triarylsulfonium functionality in alcohols, were precursors themselves act as the PTCs and do not require the addition of base. The “minimalist” method eliminates not only the need for a base or PTC but also azeotropic evaporation steps, enabling efficient fluorinations of the base-sensitive precursors.

4. Copper-Mediated Late-Stage Radiofluorination of Non-Activated Arenes

In the last decade or so, significant efforts have been expended on overcoming the general limitations of the conventional nucleophilic approach in direct introduction of 18F-label into the non-activated aromatic ring (general structural component in a great number of potentially useful PET radiotracers), leading to approach commonly referred to as “late-stage radiofluorinations.” As the result of those efforts, several innovative 18F-labeling approaches for labeling of electron-rich arenes (heteroarenes) have been introduced using, for example, (mesytyl)(aryl)iodonium salts, spirocyclic hypervalent iodine (III) complexes, and organoboron and stannane precursors (for recent reviews see [69,70,71,72,73]).
Among the methods described, Cu-catalyzed radiofluorination reactions using commercially-available CuOTf(CH3CN)4 and Cu(OTf)2(Py)4 complexes as sources of copper have been found to be a particularly useful route to a diverse range of complex radiopharmaceuticals that cannot be readily accessed through conventional approaches; numerous examples are presented in recent reviews [74,75]. However, attempts at the application of those methods to full-batch clinical production revealed several problems, including modest or fluctuating radiochemical yields, long synthesis times, and problematic protocols with relatively large amounts of precursors and catalysts, base sensitivities, and others. Therefore, in parallel to the development of methods for the Cu-mediated fluorinations, substantial efforts were expended on designing novel protocols for reprocessing [18F]fluoride that could be more easily implemented in automated modules.
Copper-mediated fluorination of (mesityl)(aryl)iodonium (MAI) salts promoted by CuOTf(CH3CN)4 complex was recently introduced by the Sanford and Scott groups [76,77] as an alternative to the use of diaryl iodonium salt precursors pioneered in 1995 by Pike’s group [78]. Owing to their electron-deficient nature at the iodine center and simultaneous presence of an excellent leaving-group in the form of a Ph-I fragment, diaryl iodonium salts are exceptionally reactive towards nucleophilic fluorination. The most common application has been the preparation of 18F-labeled aromatic building blocks [15,79] such as 4-[18F]fluoroiodobenzene, often employed in the radiolabeling via Pd-catalyzed cross-coupling reactions [80]. Notable examples of the radiopharmaceuticals accessible via fluorination of diaryl iodonium salts are 6-[18F]fluorodopamine [81] and 6-L-[18F]FDOPA [82]. Despite the straightforward labeling approach, the later radiotracer, with growing clinical application in both neurology and oncology fields, was obtained from commercially available anisyl(aryl)iodonium triflate salt (Scheme 1, upper part) in relatively poor RCY of 15% (decay-corrected) with ca. 120 min synthesis time. Using the same precursor but applying an optimized minimalist protocol [68] under cryptand/base free conditions, RCY of 6-L-[18F]FDOPA was increased to 27–35% (decay corrected) with ca. 70 min synthesis time [83]. The major drawbacks of these methods are the requirement for high temperatures (140 °C) as well as the use of “non-conventional” solvents such as diglyme [82] or toluene [81,83] that are not well compatible with the automated module’s components [84].
In contrast, (mesityl)(aryl)iodonium (MAI) salts [77] are easily fluorinated under moderate temperatures (typically 85 °C, 20 min) in the commonly used polar aprotic solvents such as DMF. In the original method [77], 18-crown-6/K2CO3 was suggested for the activation of the [18F]fluoride obtained in the conventional azeotropic drying procedure. Under these conditions, high RCC values (up to 80%) were demonstrated for a series of model substrates, however, for complex molecules, such as protected 6-L-[18F]FDOPA RCC values in order of 20% were obtained using a tosylate counter ion, increasing up to 30% when a tetrafluoroborate was employed in the same role.
The high fluorination efficiency (in the order of 60–90%, depending on the position of fluorine in the aromatic ring) was achieved [85] when implementing “minimalist” protocol [68] in the direct preparation of a series of 2,3- and 4-[18F]labeled-L-phenylalanines, starting from respective MAI precursors (Scheme 1). [18F]fluoride retained on the anion-exchange resin was eluted with a solution of the MAI precursor in MeOH/DMF, followed by fluorination reaction in the same combination of solvents without any drying or solvent replacement steps, thus significantly simplifying automation. However, the utilization of this methodology for labeling electron-rich arenes is hampered by a relatively difficult preparation of the iodonium salt precursors, in particular, polyfunctionalized iodonium salts, as well as the limited shelf-life of these compounds.
Consequently, the synthesis of radiofluorinated arenes labeled in inactivated positions using arylboronic acid pinacol esters (ArylBPin) as precursors and Cu(OTf)2(Py)4 as a catalyst has been recognized as a versatile route to a number of innovative 18F-radiotracers and clinically significant radiopharmaceuticals [73,74,75]. This approach is based on copper-promoted Chan-Lam C-F cross-coupling reaction and was brought to 18F-fluorine radiochemistry in 2014 by the Gouverneur’s group [86] followed cold chemistry researches by the Sanford’s group [87]. In the follow-up studies, Cu-mediated radiofluorination of organoborons [88] and (hetero)aryl organostannanes [89] were developed as well, however, the applications compared to ArylBPin precursors are more limited. As with most new approaches, this methodology was initially developed through small-scale experiments based on the use of small aliquots of aqueous [18F]F and subsequent scale-up to production quantities required for generating clinical doses revealed numerous issues.
In the first application of the method [86], a conventional K2.2.2./K2CO3 protocol with azeotropic drying was used to obtain reactive [18F]fluoride for the reaction catalyzed by Cu(OTf)2Py4 in the presence of O2 (or atmospheric air). Due to the sensitivity of the Cu-mediated process to basic conditions the “low-base” [18F]F reprocessing protocol [46] (substantially reducing both base and kryptofix amounts) or the use of less basic cryptand/base combinations [90] was suggested. Results by Gouverneur’s group [86] in collaboration with other laboratories [90] demonstrated the applicability of the method to the synthesis of eight widely used radiotracers: [18F]FMTEB, [18F]FPEB, [18F]flumazenil, [18F]DAA1106, [18F]MFBG, 6-L-[18F]FDOPA, [18F]FMT, and [18F]FDA (Figure 4).
The methodology was shown to be adaptable to different automation platforms where conventional ACEs (QMA, 46 mg, or PSHCO3) were used for trapping and elution of the radionuclide. Depending on the particular radiotracer, different kryptofix-based eluents were employed with the replacement of “standard” K2CO3 with less basic K2C2O4 or KH2PO4. For example, in the preparation of 6-L-[18F]FDOPA on Synthra automated module following eluent composition was used: K2.2.2. (6.3 mg), K2C2O4 (1 mg) and K2CO3 (0.1 mg) in 1 mL of MeCN/H2O, 4:1. To the azeotropically dried [18F]fluoride 20 µmol Cu(OTf)2(Py)4 and 20 µmol of di-Boc precursor in anhydrous DMF (400 µL) were added and the reaction mixture heated to 120 °C for 20 min, followed by acid hydrolysis (57% HI, 150 °C for 10 min). 6-L-[18F]FDOPA was isolated in the RCY of 9% (non-decay corrected) after ca. 150 min synthesis time [90].
Further improvements of the methodology came with the employment of the weakly or non-basic organic salts in suitable aprotic or protic solvents such as Et4NHCO3 [91] and Bu4NOTf [92] with good solubility in organic solvents. In addition to improved radiofluorination efficiency, using solutions of these PTCs in alcohols for elution of [18F]F from conventional AECs gives additional benefits for automation through the possibility to avoid time-consuming azeotropic drying and solvent replacement steps.
In addition to the aforementioned commercially available PTCs, pyridinium sulfonate salts in organic solutions [93] have been found to be efficient for recovery of radionuclide from WAX column (1 cc, 30 mg) with about 80% EE. Following such procedure Cu-mediated radiofluorination of ArylBPin precursors using Cu(OTf)2(Py)4 catalyst can be performed directly in the same solvent bypassing any solvent replacement. Using 4-dimethylaminopyridinium triflate (DMAP.OTf) in DMA (110 °C, 20 min, commonly used conditions for fluorination of ArylBPin [86,90]) the RCCs of 80–90% were reported for seven model boronates and ArylBPin precursor for protected 4-[18F]fluoro-D,L-phenylalanine. In the following study [94] DMAP.OTf was used in combination with Cu(OTf)2. The RCCs values obtained for various model substrates in this case varied from 23 to 87%; [18F]FMZ, GABAA receptors radioligand, was obtained on a preparative scale in 47% isolated RCY with an average synthesis time of 60 min. The general utility of Cu(OTf)2 in conjunction with pyridine as an alternative to Cu(OTf)2(Py)4 complex was previously demonstrated by Scott’s group in high-yield examples of radiofluorinations of organoboron compounds, including ArylBPin precurors, arylboronic acids [88], and arylstannanes [89] using K [18F]F/KOTf/K2CO3 as the source of nucleophilic [18F]fluoride. A more recent publication of Scott’s group [95] on the customized [18F]F elution techniques reports the enhancement of copper (II) triflate mediated radiofluorination of arylBPin substrates when using an aqueous solution of a weak non-ionic base such as 4-DMPA. However, using an aqueous solution necessitates the inclusion of azeotropic drying steps, making this procedure less attractive for automation. As for organic solutions of pyridinium sulfonate salts [93,94], despite of their distinctive features allowing for operationally simple [18F]fluoride processing protocol, their application in routine productions is quite limited: for clinical use commercially available materials (catalysts) is preferable.
In current practice, the procedure for alcohol-enhanced Cu-mediated radiofluorination [91], which consists of the direct elution of the radionuclide with tetraalkylammonium salt in nBuOH into a solution of the precursor and Cu(OTf)2(Py)4 in DMA followed by heating of resulting mixture under air has found broad application using various substrates. The presence of alcohol as co-solvent in Cu-mediated process has a beneficial effect on the radiofluorination efficiency leading to substantial increases in the radiochemical yield of various arylBPin substrates [91,96,97,98,99,100,101,102]. The limitation of this approach in the original version [91] suggested by Neumaier’s group is the need for a high amount of the reactants (up to 60 μmol of precursor and 30 μmol of Cu-catalyst). Therefore, in the follow-up studies of the same group possibility of reduction in the amount of Et4NHCO3 was investigated using less viscous and more polar MeOH as an eluting solvent. Correspondingly, the quantities of precursor/copper catalyst could be reduced to 10–15 and 5–10 μmol, respectively [96,97,98,99].
The notable feature of the “alcohol enhanced” method is the use of reverse loading/elution protocol (Figure 5) that starts with loading radionuclide onto an AEC (QMA cartridge) from the male side followed by rinsing of the resin with MeOH (2 mL) and drying with air from the same side. Then [18F]F is eluted from the female side using a solution of Et4NHCO3 (2 mg, 10 µmol) in MeOH (0.5 mL). MeOH is evaporated using a flow of air at 80 °C within 10 min. A solution of ArylBPin precursor (16 µmol, 1 eq) and Cu(OTf)2Py4 (20 µmol, 1.25 eq) in DMA/nBuOH 2:1 (750 µL) was added to [18F]Et4NF, and the resulting solution was heated at 110 °C for 10–20 min under air or argon. This typical protocol has been shown to be robust, displaying broad substrate scope. However, it still required the evaporation of the solvent (MeOH) from the eluent before fluorination.
To overcome this problem, further improvements of the methodology were introduced by our group [92] via the employment of Bu4NOTf as non-basic PTC and low-boiling iPrOH for eluting [18F]fluoride. Using reverse loading/elution protocol in conjunction with QMA cartridge of 130 mg the EE > 90% was achieved using a solution of Bu4NOTf (5 mg, 12.5 µmol) in iPrOH (0.8 mL). Notably, the following Cu-mediated radiofluorination reaction can be performed directly with alcohol as co-solvent with DMA [92] or MeCN [92], offering significant practical benefits through elimination of not only azeotropic drying but any solvent removal steps preceding radiofluorination. Radiofluorination in DMA/iPrOH was fully compatible with various model substrates and ArylBPin precursors for two aromatic amino acids. Radiofuorination in neat iPrOH was substrate-dependent, being particularly effective for the preparation of 6-L-[18F]FDOPA (reaction conditions are shown under Scheme 2). More recently100, this approach was employed in the automated synthesis of 6-L-[18F]FMT, the radiotracer representing a valuable alternative to 6-L-[18F]FDOPA in the diagnosis and staging of Parkinson’s disease with simplified kinetic modeling. Using Bpin-substituted chiral Ni(II) complex as a labeling precursor [99], Cu-mediated radiofluorination was performed in highly boiling iPrOH/MeCN mixture in the presence of Bu4NOTf as the PTC affording the target radiotracer in ca.20% activity yield (non-decay corrected) within 70 min. The radiofluorination proceeds under nitrogen instead of air, usually applied for Cu-mediated 18F-fluorination of boronates. This could be advantageous if the module is also used for the production of other PET tracers, which require inert conditions.
The use of Bu4NOTf as a replacement for KOTf and K2CO3 was suggested by Scott’s group [103] for the automated synthesis of 6-L-[18F]FDOPA (Scheme 2). An aqueous eluent consisting of 15 mg/mL Bu4NOTf and 0.2 mg/mL Cs2CO3 (0.5 mL), respectively, gave good recovery of [18F]F from the QMA cartridge and provided over 50% RCC in radiofluorination in DMF with addition of pyridine. However, the performance of this protocol necessitated a conventional azeotropic drying step resulting in a longer synthesis time (Scheme 2).
To this end, various protocols (Scheme 2) for activating [18F]F for Cu-mediated fluorinations of arylBPin substrates have been introduced by different groups [86,91,92,103] to be implemented onto automated production of 6-L-[18F]FDOPA, including GMP compliant synthesis using commercially available precursor 3. Due to easy-to-cleave MOM protecting groups on the catechol moiety and tert-butyl ester group protecting the amino acid fragment in the structure of 3, hydrolysis/deprotection step can be performed using less aggressive (to compare with 57% HI [86]) aqueous HCl solutions. Despite the reported Cu-mediated fluorinations of arylBPin precursors allowing the generation of multiple clinical doses using high initial radionuclide activity, they still require significant improvement in terms of activity yields.

5. Conclusions

[18F]fluoride as a first step in the route to various classes of radiopharmaceuticals through SN2 and SNAr nucleophilic substitution reactions has been explored extensively over several decades and continues to remain an important area of research when it comes to the development of new PET radiotracers. In parallel with the appearance of new radiolabeling approaches, substantial efforts are focused on the preceding reprocessing steps aimed to enhance the reactivity of the radionuclide that traditionally arrives from cyclotrons as an aqueous [18F]fluoride ion. Usually, the [18F]fluoride is trapped on an anion-exchange resin, eluted with a mixture of MeCN and water containing complex of inorganic base and cryptand (typically kryptofix-K2.2.2./K2CO3), and dried by repeated azeotropic distillation using MeCN. However, the time-consuming azeotropic drying is associated with radioactivity losses due to decay of the isotope, adsorption on the walls of the reactor vessel, and often is difficult to automate. Another limitation of this approach is the use of the substantial amounts of base to facilitate the elution of [18F]fluoride from standard anion exchange cartridges commonly used. Basic reaction conditions resulting often lead to the formation of undesired radiolabeled by-products and the generation of chemical impurities from the decomposition of base-sensitive precursors. In order to resolve these issues considerable efforts were expended to develop novel [18F]fluoride adsorbtion/elution protocols omitting azeotropic drying procedure and, more importantly, providing milder fluorination conditions for base-sensitive precursors. It must be stressed that the basicity of the reaction environment is one of the critical aspects that influence the yield of SN2 and SNAr radioflurinations, evermore so in the case of base-sensitive precursor being used. Moving away from kryptofix to tetraalkylammonium salts as PTCs often allows not only to increase the yield but can also lead to substantial simplification of overall synthetic procedure. Higher yields, lower levels of by-products, smaller amounts of precursor used, and fewer impurities formed in several cases have allowed replacing HPLC purification procedure with a cartridge-based one altogether—an immense benefit for any clinical production protocol. In this mini-review, we focused on both well-established [18F]fluoride processing protocols already used in routine clinical productions as well as new and innovative methods and techniques in regards to possibilities of their implementation in modern automated synthesizers. By providing a comprehensive introduction and analysis of distinctive features and pitfalls of various approaches followed by a detailed description of optimized protocols for the preparation of most important radiotracers, we hope to facilitate the implementation of the new techniques described into future clinical-production scale 18F-fluorination.

Author Contributions

R.N.K. conceived and designed the review concept; V.V.O. reviewed the papers on Cu-mediated fluorination; R.N.K. and V.V.O. wrote the draft manuscript; R.N.K. finalized the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the state assignment of the Ministry of Education and Science of the Russian Federation, theme number 0133-2019-0004. The contribution of V.V.O. was funded by the Russian Foundation of Basic Researches 20-53-12030\21.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Eberl, S.; Eriksson, T.; Svedberg, O.; Norling, J.; Henderson, D.; Lam, P.; Fulham, M.J. High Beam Current Operation of a PETtraceTM Cyclotron for 18F Production. Appl. Radiat. Isot. 2012, 70, 922–930. [Google Scholar] [CrossRef]
  2. Cai, L.; Lu, S.-Y.; Pike, V.W. Chemistry with [18F]Fluoride Ion. Eur. J. Org. Chem. 2008, 17, 2853–2873. [Google Scholar] [CrossRef]
  3. Coenen, H.H.; Ermert, J. 18F-labelling innovations and their potential for clinical application. Clin. Trans. Imaging 2018, 6, 16. [Google Scholar] [CrossRef]
  4. Deng, X.; Rong, J.; Wang, L.; Vasdev, N.; Zhang, L.; Josephson, L.; Liang, S.H. Chemistry for Positron Emission Tomography: Recent Advances in 11C-, 18F-, 13N-, and 15O-Labeling Reactions. Angew. Chem. Int. Ed. 2019, 58, 2580–2605. [Google Scholar] [CrossRef]
  5. Jacobson, O.; Kiesewetter, D.O.; Chen, X. Fluorine-18 Radiochemistry, Labeling Strategies and Synthetic Routes. Bioconjugate Chem. 2014, 26, 1–18. [Google Scholar] [CrossRef] [PubMed]
  6. Schlyer, D.J.; Bastos, M.A.V.; Alexoff, D.; Wolf, A.P. Separation of [18F]fluoride from [18O]water using anion exchange resin. Appl. Radiat. Isot. 1990, 41, 531–533. [Google Scholar] [CrossRef]
  7. Sachinidis, J.I.; Poniger, S.; Tochon-Danguy, H.J. Automation for Optimised Production of Fluorine-18-Labelled Radiopharmaceuticals. Curr. Radiopharm. 2010, 3, 248–253. [Google Scholar] [CrossRef]
  8. Hjelstuen, O.K.; Svadberg, A.; Olberg, D.E.; Rosser, M. Standardization of fluorine-18 manufacturing processes: New scientific challenges for PET. Eur. J. Pharm. Biopharm. 2011, 78, 307–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Krasikova, R. PET Radiochemistry automation: State of the art and future trends in 18F-nucleophilic fluorination. Curr. Org. Chem. 2013, 17, 2097–2107. [Google Scholar] [CrossRef]
  10. Coenen, H.H.; Schuller, M.; Stocklin, G.; Klatte, B.; Knochel, A. Preparation of N.C.A. [17-18F]-fluoroheptadecanoic acid in high yields via aminopolyether supported, nucleophilic fluorination. J. Labelled. Compds. Radiopharm. 1986, 23, 455–466. [Google Scholar] [CrossRef]
  11. Hamacher, K.; Coenen, H.; Stocklin, G. Efficient stereospecifc synthesis of no-carrier-added 2-[F-18]-fuoro-2-deoxy-d-glucose using aminopolyether supported nucleophilic-substitution. J. Nucl. Med. 1986, 27, 235–238. [Google Scholar] [PubMed]
  12. Coenen, H.H.; Elsinga, P.H.; Iwata, R.; Kilbourn, M.R.; Pillai, M.R.; Rajan, M.G.; Wagner, H.N.; Zagnun, J.J. Fluorine-18 radiopharmaceuticals beyond [18F] FDG for use in oncology and neurosciences. Nucl. Med. Biol. 2010, 37, 727–740. [Google Scholar] [CrossRef] [PubMed]
  13. Brodack, J.W.; Kilbourn, M.R.; Welch, M.J.; Katzenellenbogen, J.A. NCA 16 alpha-[18F]fluoroestradiol-17 beta: The effect of reaction vessel on fluorine-18 resolubilization, product yield, and effective specific activity. Int. J. Rad. Appl. Instrum. A 1986, 37, 217–221. [Google Scholar] [CrossRef]
  14. Zhang, M.-R.; Suzuki, K. [18F]Fluoroalkyl agents: Synthesis, reactivity and application for development of PET ligands in molecular imaging. Curr. Top. Med. Chem. 2007, 7, 1817–1828. [Google Scholar] [CrossRef]
  15. Van der Born, D.; Pees, A.; Poot, A.J.; Orru, R.V.A.; Windhorst, A.D.; Vugts, D.G. Fluorine-18 labelled building blocks for PET tracer synthesis. Chem. Soc. Rev. 2017, 46, 4709–4773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Gomzina, N.A.; Zaitsev, V.V.; Krasikova, R.N. Optimization of nucleophilic fluorination step in the synthesis of various compounds labelled with fluorine-18 for their use as PET radiotracers. J. Label. Compds. Radiopharm. 2001, 44 (Suppl. S1), S895–S897. [Google Scholar] [CrossRef]
  17. Gomzina, N.A.; Vassiliev, D.A.; Krasikova, R.N. Optimization of robotic preparation of 2-[18F]fluoro-2-deoxy-D-glucose based on alkali hydrolysis. Radiochemistry 2002, 44, 366–372. [Google Scholar] [CrossRef]
  18. Fedorova, O.S.; Vaitekhovich, F.P.; Krasikova, R.N. Automated synthesis of [18F] fluoromethylcholine for positron emission tomography imaging. Pharm. Chem. J. 2018, 52, 730–734. [Google Scholar] [CrossRef]
  19. 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. [Google Scholar] [CrossRef]
  20. Li, G.Y.; Vaulina, D.D.; Li, J.J.; Fedorova, O.S.; Wang, H.E.; Liu, R.S.; Krasikova, R.N.; Chen, C.L. Synthesis and biological evaluation of 2-(3,4-dimethoxyphenyl)-6-(2-[18F]fluoroethoxy)benzothiazole ([18F]FEDBT) for PET imaging of breast cancer. Bioorg. Med. Chem. Lett. 2017, 27, 3460–3463. [Google Scholar] [CrossRef]
  21. Ryzhikov, N.N.; Seneca, N.; Krasikova, R.N.; Gomzina, N.A.; Shchukin, E.; Fedorova, O.S.; Vassiliev, D.A.; Gulyás, B.; Hall, H.; Savic, I.; et al. Preparation of high specific radioactivity [18F]flumazenil and its evaluation in cynomolgus monkey by positron emission tomography. Nucl. Med. Biol. 2005, 32, 109–116. [Google Scholar] [CrossRef]
  22. Stepanov, V.; Krasikova, R.; Raus, L.; Loog, O.; Hiltunen, J.; Halldin, C. An Efficient One-Step Radiosynthesis of [18F]FEPE2I, a PET Radioligand for Imaging of Dopamine Transporters. J. Label. Compd. Radiopharm. 2012, 55, 206–210. [Google Scholar] [CrossRef]
  23. Arakawa, R.; Takano, A.; Stenkrona, P.; Stepanov, V.; Nag, S.; Jahan, M.; Grybäck, P.; Bolin, M.; Chen, L.; Zhang, L.; et al. PET imaging of beta-secretase 1 in the human brain: Radiation dosimetry, quantification, and test-retest examination of [18F]PF-06684511. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 2429–2439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Finnema, S.J.; Stepanov, V.; Nakao, R.; Sromek, A.W.; Zhang, T.; Neumeyer, J.L.; George, S.R.; Seeman, P.; Stabin, M.G.; Jonsson, C.; et al. (18)F-MCL-524, an (18)F-Labeled Dopamine D2 and D3 Receptor Agonist Sensitive to Dopamine: A Preliminary PET Study. J. Nucl. Med. 2014, 55, 1164–1170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Bogni, A.; Laera, L.; Cucchi, C.; Iwata, R.; Seregni, E.; Pascali, C. An improved automated one-pot synthesis of O’-(2-[18F]fluoroethyl)-L-tyrosine ([18F]FET) based on a purification by cartridges. Nucl. Med. Biol. 2019, 72–73, 11–19. [Google Scholar] [CrossRef] [PubMed]
  26. Cucchi, C.; Bogni, A.; Casanova, C.; Seregini, E.; Pascali, C. An improved one-pot preparation of [18F]FMISO based on solid phase extraction purification: Pitfalls on the analytical method reported in the Ph.Eur.’s monograph. J. Label. Compd. Radiopharm. 2021, 1–7. [Google Scholar] [CrossRef]
  27. Wessmann, S.H.; Henriksen, G.; Wester, H.J. Cryptate mediated nucleophilic 18F-fluorination without azeotropic drying. Nuklearmedizin 2012, 51, 1–8. [Google Scholar] [CrossRef]
  28. Wurzer, A.; Di Carlo, D.; Schmidt, A.; Beck, R.; Eiber, M.; Schwaiger, M.; Wester, H.-J. Radiohybrid Ligands: A Novel Tracer Concept Exemplified by 18F- or 68Ga-Labeled rhPSMA Inhibitors. J. Nucl. Med. 2020, 61, 735–742. [Google Scholar] [CrossRef] [Green Version]
  29. Kostikov, A.P.; Chin, J.; Orchowski, K.; Niedermoser, S.; Kovacevic, M.M.; Aliaga, A.; Jurkschat, K.; Wängler, B.; Wängler, C.; Wester, H.J.; et al. Oxalic acid supported Si-18F-radiofluorination: One-step radiosynthesis of N-succinimidyl 3-(di-tert-butyl[18F]fluorosilyl)benzoate ([18F]SiFB) for protein labeling. Bioconjug Chem. 2012, 3, 106–114. [Google Scholar] [CrossRef]
  30. Iwata, R.; Pascali, C.; Terasaki, K.; Ishikawa, Y.; Furumoto, S.; Yanai, K. Minimization of the amount of Kryptofx 222—KHCO3 for applications to microscale 18F-radiolabeling. Appl. Radiat. Isot. 2017, 125, 113–118. [Google Scholar] [CrossRef]
  31. Song, R.; Tago, T.; Tatsuta, M.; Shiraishi, N.; Iwai, K.; Hirano, K.; Toyohara, D.; Tanaka, H. N-Alkyl 3-aminobut-2-enenitrile as a Non-radioactive Side Product in Nucleophilic 18F-Fluorination. ChemistrySelect 2021, 6, 2826–2831. [Google Scholar] [CrossRef]
  32. Kwon, Y.D.; Son, J.; Yun, M.J.; Chun, J.H. Azeotropic drying-free aliphatic radiofuorination to produce PET radiotracers in a mixed organic solvent system. Tetrahedron Lett. 2018, 59, 2848–2852. [Google Scholar] [CrossRef]
  33. Aerts, J.; Voccia, S.; Lemaire, C.; Giacomelli, F.; Goblet, D.; Thonon, D.; Plenevaux, A.; Warnock, G.; Luxen, A. Fast production of highly concentrated reactive [18F] fluoride for aliphatic and aromatic nucleophilic radiolabelling. Tetrahedron Lett. 2010, 51, 64–66. [Google Scholar] [CrossRef]
  34. Culbert, P.A.; Adam, M.J.; Hurtado, E.T.; Huser, J.M.A.; Jivan, S.; Lu, J.; Ruth, T.J.; Zeisler, S.K. Automated synthesis of [18F]FDG using tetrabutylammonium bicarbonate. Appl. Radiat. Isot. 1995, 46, 887–891. [Google Scholar] [CrossRef]
  35. Hamacher, K.; Coenen, H.H. Efficient routine production of the 18F-labelled amino acid O’-(2-[18F]fluoroethyl)-L-tyrosine. Appl. Radiat. Isot. 2002, 57, 853–856. [Google Scholar] [CrossRef]
  36. Krasikova, R.N.; Kuznetsova, O.F.; Fedorova, O.S.; Maleev, V.I.; Savel’eva, T.F.; Belokon, Y.N. No carrier added synthesis of O-(2′-[18F]fluoroethyl)-L-tyrosine via a novel type of chiral enantiomerically pure precursor, NiII complex of a (S)-tyrosine Schiff base. Bioorg. Med. Chem. 2008, 16, 4994–5003. [Google Scholar] [CrossRef]
  37. Jiang, X.; Li, Y.; Wang, X.; Shen, T.; Li, X.; Yao, Y.; Zhang, G.; Kou, Y.; Shen, J.; Luo, Z.; et al. Quick Automatic Synthesis of Solvent-Free 16α-[18F]Fluoroestradiol: Comparison of Kryptofix 222 and Tetrabutylammonium Bicarbonate. Front. Oncol. 2020, 10, 577979. [Google Scholar] [CrossRef]
  38. Krasikova, R.N.; Kuznetsova, O.F.; Fedorova, O.S.; Belokon, Y.N.; Maleev, V.I.; Mu, L.; Ametamey, S.; Schubiger, P.A.; Friebe, M.; Berndt, M.; et al. 4-[18F]fluoro glutamic acid (BAY 85-8050)—A new amino acid radiotracer for PET imaging of tumors: Synthesis and in vitro characterization. J. Med. Chem. 2011, 54, 406–410. [Google Scholar] [CrossRef]
  39. Brichard, L.; Aigbirhio, F.I. An Efficient Method for Enhancing the Reactivity and Flexibility of [18F]Fluoride Towards Nucleophilic Substitution Using Tetraethylammonium Bicarbonate. Eur. J. Org. Chem. 2014, 28, 6145–6149. [Google Scholar] [CrossRef]
  40. Cybulska, K.A.; Bloemers, V.; Perk, L.R.; Laverman, P. Optimised GMP-compliant production of [18F]DPA-714 on the Trasis AllinOne module. EJNMMI Radiopharm. Chem. 2021, 6, 20. [Google Scholar] [CrossRef]
  41. Bratteby, K.; Shalgunov, V.; Herth, M. Aliphatic 18F-Radiofluorination: Recent Advances in the Labeling of Base-Sensitive Substrates. ChemMedChem. 2021, 16, 2612–2622. [Google Scholar] [CrossRef]
  42. Orlovskaya, V.; Antuganov, D.; Fedorova, O.; Timofeev, V.; Krasikova, R. Tetrabutyl ammonium tosylate as inert phase-transfer catalyst: The key to high efficiency SN2 radiofluorinations. Appl. Radiat. Isot. 2020, 163, 109195. [Google Scholar] [CrossRef]
  43. Bratteby, K.; Shalgunov, V.; Battisti, U.M.; Petersen, I.N.; Broek, S.L.; Ohlsson, T.; Gillings, N.; Erlandsson, M.; Herth, M. Insights into Elution of Anion Exchange Cartridges: Opening the Path towards Aliphatic 18F-Radiolabeling of Base-Sensitive Tracers. ACS Pharmacol. Transl. Sci. 2021, 4, 1556–1566. [Google Scholar] [CrossRef]
  44. Inkster, J.A.H.; Akurathi, V.; Sromek, A.W.; Chen, Y.; Neumeyer, J.L.; Packard, A.B. A non-anhydrous, minimally basic protocol for the simplification of nucleophilic 18F-fluorination chemistry. Sci. Rep. 2020, 10, 6818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Seo, J.W.; Lee, B.S.; Lee, S.J.; Oh, S.J.; Chi, D.Y. Fast and easy drying method for the preparation of activated [18F]fluoride using polymer cartridge. Bull. Korean Chem. Soc. 2011, 32, 71–76. [Google Scholar] [CrossRef] [Green Version]
  46. Zlatopolskiy, B.D.; Zischler, J.; Krapf, P.; Zarrad, F.; Urusova, E.A.; Kordys, E.; Endepols, H.; Neumaier, B. Copper-mediated aromatic radiofluorination revisited: Efficient production of PET tracers on a preparative scale. Chem. Eur. J. 2015, 21, 5972–5979. [Google Scholar] [CrossRef] [PubMed]
  47. 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. 2019, 152, 135–139. [Google Scholar] [CrossRef]
  48. Fedorova, O.S.; Orlovskaya, V.V.; Nadporojskii, M.; Krasikova, R.N. Automated synthesis of the 16α-[18F]fluoroestradiol ([18F]FES): Minimization of precursor amount and resulting benefits. Radiochim. Acta 2020, 108, 979–988. [Google Scholar] [CrossRef]
  49. Bratteby, K.; Denholt, C.L.; Lehel, S.; Petersen, I.N.; Madsen, J.; Erlandsson, M.; Ohlsson, T.; Herth, M.; Gillings, N. Fully Automated GMP-Compliant Synthesis of [18F]FE-PE2I. Pharmaceuticals 2021, 14, 601. [Google Scholar] [CrossRef]
  50. Inkster, J.A.H.; Sromek, A.W.; Akurathi, V.; Neumeyer, J.L.; Packard, A.B. The Non-Anhydrous, Minimally Basic Synthesis of the Dopamine D2 Agonist [18F]MCL-524. Chemistry 2021, 3, 1047–1056. [Google Scholar] [CrossRef]
  51. Toorongian, S.A.; Mulholland, K.; Jewett, D.M.; Bachelor, M.A.; Kilbourn, M.R. Routine production of 2-deoxy-2-[18F]fluoro-D-glucose by direct nucleophilic exchange on a quaternary 4-aminopyridinium resin. Int. J. Radiat. Appl. Instrum. B 1990, 17, 273–279. [Google Scholar] [CrossRef] [Green Version]
  52. Kuge, Y.; Tsukamoto, E.; Katoh, C.; Seki, K.; Ohkura, K.; Ohmiya, Y.; Nishijima, K.; Tanaka, A.; Sasaki, M.; Tamaki, N. Synthesis of 18F-FDG with FDG MicroLab system: Basic studies for clinical application. Jpn. J. Nucl. Med. 1999, 36, 873–878. [Google Scholar] [PubMed]
  53. Mathiessen, B.; Zhuravlev, F. Automated Solid-Phase Radiofluorination Using Polymer-Supported Phosphazenes. Molecules 2013, 18, 10531–10547. [Google Scholar] [CrossRef] [PubMed]
  54. Lemaire, C.F.; Aerts, J.; Voccia, S.; Libert, L.C.; Mercier, F.; Goblet, D.; Plenevaux, A.R.; Luxen, A.J. Fast Production of Highly Reactive No-Carrier-Added [18F]Fluoride for the Labeling of Radiopharmaceuticals. Angew. Chem. Int. Ed. 2010, 49, 3161–3164. [Google Scholar] [CrossRef] [PubMed]
  55. Rong, J.; Liang, S.H. Aliphatic [18F]Fluorination Chemistry for Positron Emission Tomography. In Fluorination, Synthetic Organofluorine Chemistry 1; Hu, J., Umemoto, T., Eds.; Springer Nature Singapore Pte Ltd.: Singapore, 2018. [Google Scholar] [CrossRef]
  56. Liang, S.H.; Vasdev, N. Aliphatic 18F Bond Formation via Transition Metal Based [18F]Fluorination. Angew. Chem. Int. Ed. Engl. 2014, 53, 11416–11418. [Google Scholar] [CrossRef] [Green Version]
  57. Thompson, S.; Lee, S.J.; Jackson, I.M.; Ichiishi, N.; Brooks, A.F.; Sanford, M.S.; Scott, P.J.H. Synthesis of [18F]-γ-Fluoro-α,β-unsaturated Esters and Ketones via Vinylogous 18F-Fluorination of α-Diazoacetates with [18F]AgF. Synthesis 2019, 51, 4401–4407. [Google Scholar] [CrossRef]
  58. Lemaire, C.; Cantineau, R.; Guillaume, M.; Plenevaux, A.; Christiaens, L. Fluorine-18-altanserin: A radioligand for the study of serotonin receptors with PET: Radiolabeling and in vivo biologic behavior in rats. J. Nucl. Med. 1991, 32, 2266–2272. [Google Scholar] [PubMed]
  59. Hasler, F.; Kuznetsova, O.F.; Krasikova, R.N.; Cservenyak, T.; Quednow, B.B.; Vollenweider, F.X.; Ametamey, S.M.; Westera, G. GMP-compliant radiosynthesis of [18F]altanserin and human plasma metabolite studies. Appl. Radiat. Isot. 2009, 67, 598–601. [Google Scholar] [CrossRef] [Green Version]
  60. Le, B.D.; Lemaire, C.; Ginovart, N.; Plenevaux, A.; Aerts, J.; Brihaye, C.; Hassoun, W.; Leviel, V.; Mekhsian, P.; Weissmann, D.; et al. High-yield radiosynthesis and preliminary in vivo evaluation of p-[18F]MPPF, a fuoro analog of WAY-100635. Nucl. Med. Biol. 1998, 25, 343–350. [Google Scholar] [CrossRef]
  61. Wong, D.F.; Waterhouse, R.; Kuwabara, H.; Kim, J.; Brašić, J.R.; Chamroonrat, W.; Stabins, M.; Holt, D.P.; Dannals, R.F.; Hamill, T.G.; et al. 18F-FPEB, a PET Radiopharmaceutical for quantifying metabotropic glutamate 5 receptors: A first-in-human study of radiochemical safety, biokinetics, and radiation dosimetry. J. Nucl. Med. 2013, 54, 388–396. [Google Scholar] [CrossRef] [Green Version]
  62. Dollé, F. Fluorine-18-labelled fuoropyridines: Advances in radiopharmaceutical design. Curr. Pharm. Des. 2005, 11, 3221–3235. [Google Scholar] [CrossRef] [PubMed]
  63. Horti, A.; Koren, A.O.; Ravert, H.T.; Musachio, J.L.; Mathews, W.B.; London, E.D.; Dannals, R.F. Synthesis of a radiotracer for studying nicotinic acetylcholine receptors: 2-[18F]Fluoro-3-(2(S)- azetidinylmethoxy)pyridine (2-[18F]F-A-85380). J. Label. Compds. Radiopharm. 1998, 41, 309–318. [Google Scholar] [CrossRef]
  64. Cselényi, Z.; Jönhagen, M.E.; Forsberg, A.; Halldin, C.; Julin, P.; Schou, M.; Johnström, P.; Varnäs, K.; Svensson, S.; Farde, L. Clinical validation of 18F-AZD4694, an amyloid-β-specific PET radioligand. J. Nucl. Med. 2012, 53, 415–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Cardinale, J.; Martin, R.; Remde, Y.; Schäfer, M.; Hienzsch, A.; Hübner, S.; Zerges, A.M.; Marx, H.; Hesse, R.; Weber, K.; 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Olberg, D.E.; Arukwe, J.M.; Hjelstuen, O.K.; Solbakken, M.; Kindberg, G.M.; Cuthbertson, A. One Step Radiosynthesis of 6-[18F]Fluoronicotinic Acid 2,3,5,6-Tetrafluorophenyl Ester ([18F]F-Py-TFP): A New Prosthetic Group for Efficient Labeling of Biomolecules with Fluorine-18. J. Med. Chem. 2010, 53, 1732–1740. [Google Scholar] [CrossRef]
  67. Basuli, F.; Zhang, X.; Jagoda, E.M.; Choyke, P.L.; Swenson, R.E. Facile room temperature synthesis of fluorine-18 labeled fluoronicotinic acid-2,3,5,6-tetrafluorophenyl ester without azeotropic drying of fluorine-18. Nucl. Med. Biol. 2016, 43, 770–772. [Google Scholar] [CrossRef] [Green Version]
  68. Richarz, R.; Krapf, P.; Zarrad, F.; Urusova, E.A.; Neumaier, B.; Zlatopolskiy, B.D. Neither azeotropic drying, nor base nor other additives: A minimalist approach to 18F-labeling. Org. Biomol. Chem. 2014, 12, 8094–8099. [Google Scholar] [CrossRef]
  69. Lee, E.; Kamlet, A.S.; Powers, D.C.; Neumann, C.N.; Boursalian, G.B.; Furuya, T.; Choi, D.C.; Hooker, J.M.; Ritter, T. A Fluoride-Derived Electrophilic Late-Stage Fluorination Reagent for PET Imaging. Science 2011, 334, 639–642. [Google Scholar] [CrossRef] [Green Version]
  70. Brooks, A.F.; Topczewski, J.J.; Ichiishi, N.; Sanford, M.S.; Scott, P. Late-Stage [18F]fluorination: New Solutions to Old Problems. Chem. Sci. 2014, 5, 4545–4553. [Google Scholar] [CrossRef] [Green Version]
  71. Taylor, N.J.; Emer, E.; Preshlock, S.; Schedler, M.; Tredwell, M.; Verhoog, S.; Mercier, J.; Genicot, C.; Gouverneur, V. Derisking the Cu-Mediated 18F-Fluorination of Heterocyclic Positron Emission Tomography Radioligands. J. Am. Chem. Soc. 2017, 139, 8267–8276. [Google Scholar] [CrossRef]
  72. Cole, E.L.; Stewart, M.N.; Littich, R.; Hoareau, R.; Scott, P. Radiosyntheses Using Fluorine-18: The Art and Science of Late Stage Fluorination. Curr. Top. Med. Chem. 2014, 14, 875–900. [Google Scholar] [CrossRef] [Green Version]
  73. Preshlock, S.; Tredwell, M.; Gouverneur, V. 18F-Labeling of Arenes and Heteroarenes for Applications in Positron Emission Tomography. Chem. Rev. 2016, 116, 719–766. [Google Scholar] [CrossRef]
  74. Zarganes-Tzitzikas, T.; Clemente, G.S.; Elsinga, P.H.; Dömling, A. MCR Scaffolds Get Hotter With 18F-Labeling. Molecules 2019, 24, 1327. [Google Scholar] [CrossRef] [Green Version]
  75. Wright, J.S.; Kaur, T.; Preshlock, S.; Tanzey, S.S.; Winton, W.P.; Sharninghausen, L.S.; Wiesner, N.; Brooks, A.F.; Sanford, M.S.; Scott, P. Copper-Mediated Late-Stage Radiofluorination: Five Years of Impact on Preclinical and Clinical PET Imaging. Clin. Transl. Imaging 2020, 8, 167–206. [Google Scholar] [CrossRef]
  76. Ichiishi, N.; Canty, A.J.; Yates, B.F.; Sanford, M.S. Cu-Catalyzed Fluorination of Diaryliodonium Salts With KF. Org. Lett. 2013, 15, 5134–5137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Ichiishi, N.; Brooks, A.F.; Topczewski, J.J.; Rodnick, M.E.; Sanford, M.S.; Scott, P. Copper-Catalyzed [18F]Fluorination of (Mesityl)(aryl)iodonium Salts. Org. Lett. 2014, 16, 3224–3227. [Google Scholar] [CrossRef] [Green Version]
  78. Pike, V.W.; Aigbirhio, F.I. Reactions of Cyclotron-Produced [18F]fluoride With Diaryliodonium salts—A novel Single-Step Route to No-Carrier-Added [18F]fluoroarenes. J. Chem. Soc. Chem. Commun. 1995, 21, 2215. [Google Scholar] [CrossRef]
  79. Pike, V.W. Hypervalent Aryliodine Compounds as Precursors for Radiofluorination. J. Label. Compd. Radiopharm. 2018, 61, 196–227. [Google Scholar] [CrossRef] [PubMed]
  80. Way, J.D.; Wuest, F. Automated radiosynthesis of no-carrier-added 4-[18F]fluoroiodobenzene: A versatile building block in 18F radiochemistry. J. Label. Compds. Radiopharm. 2014, 57, 104–109. [Google Scholar] [CrossRef]
  81. Vāvere, A.L.; Neumann, K.D.; Butch, E.R.; Hu, B.; DiMagno, S.G.; Snyder, S.E. Improved, one-pot synthesis of 6-[18 F]fluorodopamine and quality control testing for use in patients with neuroblastoma. J. Label. Comp. Radiopharm. 2018, 61, 1069–1080. [Google Scholar] [CrossRef]
  82. Kuik, W.-J.; Kema, I.P.; Brouwers, A.H.; Zijlma, R.; Neumann, K.D.; Dierckx, R.A.J.O.; DiMagno, S.G.; Elsinga, P.H. 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-L-DOPA, Prepared by Conventional Electrophilic Substitution. J. Nucl. Med. 2014, 56, 106–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Maisonial, A.; Serre, A.; Ouadi, A.; Schmitt, S.; Canitrot, D.; Léal, F.; Miot-Noirault, E.; Brasse, D.; Marchand, P.; Chezal, J.M. Base/Cryptand/Metal-Free Automated Nucleophilic Radiofluorination of [18F]FDOPA from Iodonium Salts: Importance of Hydrogen Carbonate Counterion. Eur. J. Org. Chem. 2018, 2018, 7058–7065. [Google Scholar] [CrossRef]
  84. Krasikova, R.N. Nucleophilic Synthesis of 6-L-[18F]FDOPA. Is Copper-Mediated Radiofluorination the Answer? Molecules 2020, 25, 4365. [Google Scholar] [CrossRef] [PubMed]
  85. Orlovskaya, V.V.; Modemann, D.J.; Kuznetsova, O.F.; Fedorova, O.S.; Urusova, E.A.; Kolks, N.; Neumaier, B.; Krasikova, R.N.; Zlatopolskiy, B.D. Alcohol-Supported Cu-Mediated 18F-Fluorination of Iodonium Salts under “Minimalist” Conditions. Molecules 2019, 24, 3197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Tredwell, M.; Preshlock, S.M.; Taylor, N.J.; Gruber, S.; Huiban, M.; Passchier, J.; Mercier, J.; Génicot, C.; Gouverneur, V. A General Copper-Mediated Nucleophilic 18F Fluorination of Arenes. Angew. Chem. Int. Ed. 2014, 53, 7751–7755. [Google Scholar] [CrossRef] [PubMed]
  87. Ye, Y.; Schimler, S.D.; Hanley, P.S.; Sanford, M.S. Cu(OTf)2-Mediated Fluorination of Aryltrifluoroborates with Potassium Fluoride. J. Am. Chem. Soc. 2013, 135, 16292–16295. [Google Scholar] [CrossRef] [PubMed]
  88. Mossine, A.V.; Brooks, A.F.; Makaravage, K.J.; Miller, J.M.; Ichiishi, N.; Sanford, M.S.; Scott, P.J.H. Synthesis of [18F]Arenes via the Copper-Mediated [18F]Fluorination of Boronic Acids. Org. Lett. 2015, 17, 5780–5783. [Google Scholar] [CrossRef]
  89. Makaravage, K.J.; Brooks, A.F.; Mossine, A.V.; Sanford, M.S.; Scott, P.J.H. Copper-Mediated Radiofluorination of Arylstannanes with [18F]KF. Org. Lett. 2016, 18, 5440–5443. [Google Scholar] [CrossRef] [PubMed]
  90. Preshlock, S.; Calderwood, S.; Verhoog, S.; Tredwell, M.; Huiban, M.; Hienzsch, A.; Gruber, S.; Wilson, T.C.; Taylor, N.J.; Cailly, T.; et al. Enhanced copper-mediated 18F-fluorination of aryl boronic esters provides eight radiotracers for PET applications. Chem. Commun. 2016, 52, 8361–8364. [Google Scholar] [CrossRef]
  91. Zischler, J.; Kolks, N.; Modemann, D.; Neumaier, B.; Zlatopolskiy, B.D. Alcohol-Enhanced Cu-Mediated Radiofluorination. Chem. Eur. J. 2017, 23, 3251–3256. [Google Scholar] [CrossRef]
  92. Orlovskaya, V.; Fedorova, O.; Kuznetsova, O.; Krasikova, R. Cu-Mediated Radiofluorination of Aryl Pinacolboronate Esters: Alcohols as Solvents with Application to 6-L-[18F]FDOPA Synthesis. Eur. J. Org. Chem. 2020, 45, 7079–7086. [Google Scholar] [CrossRef]
  93. Antuganov, D.; Zykov, M.; Timofeev, V.; Timofeeva, K.; Antuganova, Y.; Fedorova, O.; Orlovskaya, V.; Krasikova, R. Copper-mediated radiofluorination of aryl pinacolboronate esters: A straightforward protocol using pyridinium sulfonates. Eur. J. Org. Chem. 2019, 2019, 918–922. [Google Scholar] [CrossRef]
  94. Zhang, X.; Basuli, F.; Swenson, R.E. An azeotropic drying-free approach for copper-mediated radiofluorination without addition of base. J. Label. Comp. Radiopharm. 2019, 62, 139–145. [Google Scholar] [CrossRef]
  95. Mossine, A.V.; Brooks, A.F.; Ichiishi, N.; Makaravage, K.J.; Sanford, M.S.; Scott, P.J. Development of Customized [18F]Fluoride Elution Techniques for the Enhancement of Copper-Mediated Late-Stage Radiofuorination. Sci. Rep. 2017, 7, 233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Zlatopolskiy, B.D.; Zischler, J.; Schäfer, D.; Urusova, E.A.; Guliyev, M.; Bannykh, O.; Endepols, H.; Neumaier, B. Discovery of 7-[18F]Fluorotryptophan as a Novel Positron Emission Tomography (PET) Probe for the Visualization of Tryptophan Metabolism in Vivo. Med. Chem. 2018, 61, 189–206. [Google Scholar] [CrossRef] [PubMed]
  97. Zlatopolskiy, B.D.; Neumaier, F.; Rüngeler, T.; Drewes, B.; Kolks, N.; Neumaier, B. Preparation of a First 18F-Labeled Agonist for M1 Muscarinic Acetylcholine Receptors. Molecules 2020, 25, 2880. [Google Scholar] [CrossRef] [PubMed]
  98. Zlatopolskiy, B.D.; Endepols, H.; Krasikova, R.N.; Fedorova, O.S.; Neumaier, B. 11C- and 18F-labelled tryptophans as PET-tracers for imaging of altered tryptophan metabolism in age-associated disorders. Russ. Chem. Rev. 2020, 89, 879–896. [Google Scholar] [CrossRef]
  99. Craig, A.; Kolks, N.; Urusova, E.A.; Zischler, J.; Brugger, M.; Endepols, H.; Neumaier, B.; Zlatopolskiy, B.D. Preparation of labeled aromatic amino acids via late-stage 18F-fluorination of chiral nickel and copper complexes. Chem. Commun. 2020, 56, 9505–9508. [Google Scholar] [CrossRef] [PubMed]
  100. Orlovskaya, V.V.; Craig, A.S.; Fedorova, O.S.; Kuznetsova, O.F.; Neumaier, B.; Krasikova, R.N.; Zlatopolskiy, B.D. Production of 6-L-[18F]Fluoro-m-tyrosine in an Automated Synthesis Module for 11C-Labeling. Molecules 2021, 26, 5550. [Google Scholar] [CrossRef]
  101. Bernard-Gauthier, V.; Mossine, A.V.; Mahringer, A.; Aliaga, A.; Bailey, J.J.; Shao, X.; Stauff, J.; Arteaga, J.; Sherman, P.; Grand’Maison, M.; et al. Identification of [18F]TRACK, a Fluorine-18-Labeled Tropomyosin Receptor Kinase (Trk) Inhibitor for PET Imaging. J. Med. Chem. 2018, 61, 1737–1743. [Google Scholar] [CrossRef]
  102. Yuan, G.; Shoup, T.M.; Moon, S.H.; Brownell, A.L. A concise method for fully automated radiosyntheses of [18F]JNJ-46356479 and [18F]FITM via Cu-mediated 18F-fluorination of organoboranes. RSC Adv. 2020, 10, 25223–25227. [Google Scholar] [CrossRef]
  103. Mossine, A.V.; Tanzey, S.S.; Brooks, A.F.; Makaravage, K.J.; Ichiishi, N.; Miller, J.M.; Henderson, B.D.; Skaddan, M.B.; Sanford, M.S.; Scott, P.J.H. One-pot synthesis of high molar activity 6-[18F]fluoro-l-DOPA by Cu-mediated fluorination of a BPin precursor. Org. Biomol. Chem. 2019, 17, 8701–8705. [Google Scholar] [CrossRef]
Figure 1. The structures of the commonly used tumor seeking agents prepared by conventional aliphatic radiofluorination.
Figure 1. The structures of the commonly used tumor seeking agents prepared by conventional aliphatic radiofluorination.
Applsci 12 00321 g001
Figure 2. The structures of PET brain radioligands prepared by aliphatic radiofluorination.
Figure 2. The structures of PET brain radioligands prepared by aliphatic radiofluorination.
Applsci 12 00321 g002
Figure 3. The structures of selected brain receptors radioligands prepared via conventional kryptofix-mediated aromatic nucleophilic radiofluorination of the appropriate nitro-precursors.
Figure 3. The structures of selected brain receptors radioligands prepared via conventional kryptofix-mediated aromatic nucleophilic radiofluorination of the appropriate nitro-precursors.
Applsci 12 00321 g003
Scheme 1. Preparation of 6-L-[18F]FDOPA [82] via 18F-fluorination of anisyl(aryl)iodonium salt precursors (upper part) and synthesis of 3-[18F]labeled-L-phenylalanine (lower part) using minimalist protocol [85].
Scheme 1. Preparation of 6-L-[18F]FDOPA [82] via 18F-fluorination of anisyl(aryl)iodonium salt precursors (upper part) and synthesis of 3-[18F]labeled-L-phenylalanine (lower part) using minimalist protocol [85].
Applsci 12 00321 sch001
Figure 4. Synthesis of eight radiotracers via Cu-mediated 18F-fluorination of arylBPin precursors using Cu(OTf)2(Py)4 as a copper source [90].
Figure 4. Synthesis of eight radiotracers via Cu-mediated 18F-fluorination of arylBPin precursors using Cu(OTf)2(Py)4 as a copper source [90].
Applsci 12 00321 g004
Figure 5. The schematic presentation of reverse loading-elution (back-flushed) protocol [92]. Step 1: [18F]fluoride is loaded onto QMA cartridge (130 mg) from the male side; Step 2: the cartridge is flushed from the male side by nitrogen gas for 2–3 min; Step 3: the cartridge is rinsed with 2 mL of iPrOH following drying with nitrogen from the male side; Step 4: [18F]fluoride is eluted from the female side using a solution of Bu4NOTf (5 mg, 12.5 µmol) in iPrOH (0.8 mL).
Figure 5. The schematic presentation of reverse loading-elution (back-flushed) protocol [92]. Step 1: [18F]fluoride is loaded onto QMA cartridge (130 mg) from the male side; Step 2: the cartridge is flushed from the male side by nitrogen gas for 2–3 min; Step 3: the cartridge is rinsed with 2 mL of iPrOH following drying with nitrogen from the male side; Step 4: [18F]fluoride is eluted from the female side using a solution of Bu4NOTf (5 mg, 12.5 µmol) in iPrOH (0.8 mL).
Applsci 12 00321 g005
Scheme 2. Cu-mediated synthesis of 6-L-[18F]FDOPA using various precursors (13) and different protocols. Precursor 1 [86]: elution—K2.2.2./K2C2O4/K2CO3 in 4:1 MeCN:H2O, azeotropic drying; fluorination—1 (20 µmol), Cu-complex (20 µmol), DMA, 120 °C, 20 min, under air; hydrolysis: 57% HI, 150 °C 15 min; RCY 9%, 146 min synthesis; Precursor 2 [91]: elution—Et4NHCO3/nBuOH, no azeotropic drying or solvent removal; fluorination—2 (60 µmol), Cu-complex (53 µmol), DMA/nBuOH, 110 °C, 20 min, under air; hydrolysis: 12 M HCl, 130 °C, 5 min; RCY 40% (decay corrected), synthesis time not reported; Precursor 3 [103]: elution—Bu4NOTf/Cs2CO3/H2O, azeotropic drying; fluorination—3 (4 µmol), Cu-complex (20 µmol), pyridine 500 µL, DMF, inert atmosphere; hydrolysis: 12 M HCl, 0.25M ascorbic acid, 130 °C, 5 min; RCY 6% (not decay corrected), 110 min synthesis; Precursor 3 [92]: elution—Bu4NOTf/iPrOH, no azeotropic drying or solvent removal; fluorination—3 (10 µmol), Cu-complex (16 µmol), iPrOH, 90 °C, 20 min, inert atmosphere; hydrolysis: 6 M HCl, MeOH, 0.25M ascorbic acid, 110 °C, 5 min; RCY 14% (not decay corrected), 70 min synthesis.
Scheme 2. Cu-mediated synthesis of 6-L-[18F]FDOPA using various precursors (13) and different protocols. Precursor 1 [86]: elution—K2.2.2./K2C2O4/K2CO3 in 4:1 MeCN:H2O, azeotropic drying; fluorination—1 (20 µmol), Cu-complex (20 µmol), DMA, 120 °C, 20 min, under air; hydrolysis: 57% HI, 150 °C 15 min; RCY 9%, 146 min synthesis; Precursor 2 [91]: elution—Et4NHCO3/nBuOH, no azeotropic drying or solvent removal; fluorination—2 (60 µmol), Cu-complex (53 µmol), DMA/nBuOH, 110 °C, 20 min, under air; hydrolysis: 12 M HCl, 130 °C, 5 min; RCY 40% (decay corrected), synthesis time not reported; Precursor 3 [103]: elution—Bu4NOTf/Cs2CO3/H2O, azeotropic drying; fluorination—3 (4 µmol), Cu-complex (20 µmol), pyridine 500 µL, DMF, inert atmosphere; hydrolysis: 12 M HCl, 0.25M ascorbic acid, 130 °C, 5 min; RCY 6% (not decay corrected), 110 min synthesis; Precursor 3 [92]: elution—Bu4NOTf/iPrOH, no azeotropic drying or solvent removal; fluorination—3 (10 µmol), Cu-complex (16 µmol), iPrOH, 90 °C, 20 min, inert atmosphere; hydrolysis: 6 M HCl, MeOH, 0.25M ascorbic acid, 110 °C, 5 min; RCY 14% (not decay corrected), 70 min synthesis.
Applsci 12 00321 sch002
Table 1. The synthesis of radiotracers [42] via radiofluorination in the presence of Bu4NOTs (1 mg, 2.5 µmol in 2 mL of EtOH, QMA carb, 46 mg); MeCN, 100 °C, 10 min; RCY—radiochemical yields, non-decay corrected.
Table 1. The synthesis of radiotracers [42] via radiofluorination in the presence of Bu4NOTs (1 mg, 2.5 µmol in 2 mL of EtOH, QMA carb, 46 mg); MeCN, 100 °C, 10 min; RCY—radiochemical yields, non-decay corrected.
Radiotracer
or Synthon
Precursor Amount, mg/µmolRCC, %
by TLC (n = 3)
RCY (%)/
Synthesis Time, Min
[18F]FETs0.2/0.596 ± 2-
[18F]FDG1.2/1.291 ± 160/25 [42]
[18F]FMISO0.2/0.574 ± 1-
[18F]FET0.4/0.6 *83 ± 1-
2/0.3 **93 ± 440/35 [47]
[18F]FLT1/1.165 ± 4-
[18F]FES0.2/0.594 ± 133/32 [48]
* TET precursor, ABX [35], ** chiral nickel (II) complex [36].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Krasikova, R.N.; Orlovskaya, V.V. Phase Transfer Catalysts and Role of Reaction Environment in Nucleophilc Radiofluorinations in Automated Synthesizers. Appl. Sci. 2022, 12, 321. https://doi.org/10.3390/app12010321

AMA Style

Krasikova RN, Orlovskaya VV. Phase Transfer Catalysts and Role of Reaction Environment in Nucleophilc Radiofluorinations in Automated Synthesizers. Applied Sciences. 2022; 12(1):321. https://doi.org/10.3390/app12010321

Chicago/Turabian Style

Krasikova, Raisa N., and Viktoriya V. Orlovskaya. 2022. "Phase Transfer Catalysts and Role of Reaction Environment in Nucleophilc Radiofluorinations in Automated Synthesizers" Applied Sciences 12, no. 1: 321. https://doi.org/10.3390/app12010321

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop