首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The palladium‐mediated N‐arylation of indoles with 4‐[18F]fluoroiodobenzene as a novel radiolabelling method has been developed. Optimized reaction conditions were elaborated by variation of different catalyst systems (CuI/1,2‐diamines and Pd2(dba)3/phosphine ligands), bases and solvents in the reaction of indole with 4‐[18F]fluoroiodobenzene. Optimized reaction conditions (Pd2(dba)3/(2‐(dicyclohexyl‐phosphino)‐2′‐(N,N‐dimethylamino)‐biphenyl, NaOBut, toluene, 100°C for 20 min) were applied for the synthesis of 18F‐labelled σ2 receptor ligands [18F]‐11 and [18F]‐13 which were obtained in 91 and 84% radiochemical yields, respectively. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

2.
The radiosyntheses of 5‐(4′‐[18F]fluorophenyl)‐uridine [18F]‐11 and 5‐(4′‐[18F]fluorophenyl)‐2′‐deoxy‐uridine [18F]‐12 are described. The 5‐(4′‐[18F]fluoro‐phenyl)‐substituted nucleosides were prepared via a Stille cross‐coupling reaction with 4‐[18F]fluoroiodobenzene followed by basic hydrolysis using 1 M potassium hy‐droxide. The Stille cross‐coupling reaction was optimized by screening various palladium complexes, additives and solvents. By using optimized labelling conditions (Pd2(dba)3/CuI/AsPh3 in DMF/dioxane (1:1), 20 min at 65°C), 550 MBq of [4‐18F]fluoroiodobenzene could be converted into 120 MBq (33%, decay‐corrected) of 5‐(4′‐[18F]fluorophenyl)‐2′‐deoxy‐uridine [18F]‐12 within 40 min, including HPLC purification. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

3.
The first application of a Sonogashira cross‐coupling reaction in 18F chemistry has been developed. The reaction was exemplified by the cross‐coupling of terminal alkynes (ethynylcyclopentyl carbinol 6 , 17α‐ethynyl‐3,17β‐estradiol 7 and 17α‐ethynyl‐3‐methoxy‐3,17β‐estradiol 8 ) with 4‐[18F]fluoroiodobenzene. 4,4′‐Diiododiaryliodonium salts were used as precursors for the synthesis of 4‐[18F]fluoroiodobenzene, enabling the convenient access to 4‐[18F]fluoroiodobenzene in 13–70% yield using conventional heating or microwave activation. The Sonogashira cross‐coupling of 4‐[18F]fluoroiodobenzene with terminal alkynes gave the corresponding 4‐[18F]fluorophenylethynyl‐substituted compounds [18F]‐9 , [18F]‐10 and [18F]‐13 in yields up to 88% within 20 min of starting from 4‐[18F]fluoroiodobenzene. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

4.
A fully automated synthesis of N‐succinimidyl 4‐[18F]fluorobenzoate ([18F]SFB) was carried out by a convenient three‐step, one‐pot procedure on the modified TRACERlab FXFN synthesizer, including [18F]fluorination of ethyl 4‐(trimethylammonium triflate)benzoate as the precursor, saponification of the ethyl 4‐[18F]fluorobenzoate with aqueous tetrapropylammonium hydroxide instead of sodium hydroxide, and conversion of 4‐[18F]fluorobenzoate salt ([18F]FBA) to [18F]SFB treated with N,N,N′,N′‐tetramethyl‐O‐(N‐succinimidyl)uranium tetrafluoroborate (TSTU). The purified [18F]SFB was used for the labeling of Tat membrane‐penetrating peptide (containing the Arg‐Lys‐Lys‐Arg‐Arg‐Arg‐Arg‐Arg‐Arg‐Arg‐Arg‐Pro‐Leu‐Gly‐Leu‐Ala‐Gly‐Glu‐Glu‐Glu‐Glu‐Glu‐Glu‐Glu sequence, [18F]CPP) through radiofluorination of lysine amino groups. The uncorrected radiochemical yields of [18F]SFB were as high as 25–35% (based on [18F]fluoride) (n=10) with a synthesis time of~40 min. [18F]CPP was produced in an uncorrected radiochemical yields of 10–20% (n=5) within 30 min (based on [18F]SFB). The radiochemical purities of [18F]SFB and [18F]CPP were greater than 95%. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

5.
The well established M1 selective muscarinergic antagonist Pirenzepine 11‐[2‐(4‐methyl‐piperazin‐1‐yl)‐acetyl]‐5,11‐dihydro‐benzo[e]pyrido[3,2‐b][1,4]diazepin‐6‐one (1) exhibits an unusual behaviour in vivo, which cannot be explained with M1 antagonism exclusively. One of the aspects discussed is a specific interaction with poly ADP‐ribose polymerase (PARP‐1). 1 undergoes metabolism to form LS 75 5,11‐dihydro‐benzo[e]pyrido[3,2‐b][1,4]diazepin‐6‐one (2). In order to study deviations in Pirenzepine efficacy from pure M1 binding in vivo using PET, appropriate positron emitter labelled analogues of 1 and 2 were synthesised. Non‐radioactive reference compounds 3 and 4 were tested for PARP‐1 inhibition. The n‐octanol–water partition coefficients of compounds 1, 2, 3 and 4 at pH 7.4 (logD7.4) were determined. Both, 3 and 4 were labelled with 18F via 2‐[18F]fluoroalkylation in position 5 of the benzodiazepinone moiety to obtain N5‐[18F]fluoroethyl Pirenzepine [18F]‐3 and N5‐[18F]fluoroethyl LS 75 [18F]‐4. Radiotracers [18F]‐3 and [18F]‐4 were obtained in radiochemical yields of 15±4 % and 30±5% after 120 and 110 min, respectively. Metabolism of both compounds was investigated in vitro in human and rat plasma, respectively. Compound 3 did not show activity as an inhibitor of PARP‐1. Contrary, 4 displays moderate PARP‐1 inhibition potency. The new radiotracer [18F]‐4 can be applied for molecular imaging using autoradiography and PET. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

6.
The radiosynthesis of N‐(5‐(((5‐(tert‐butyl)oxazol‐2‐yl)methyl)thio)thiazol‐2‐yl)‐4‐[18F]fluoro‐benzamide [18F]2 as a potential radiotracer for molecular imaging of cyclin‐dependent kinase‐2 (CDK‐2) expression in vivo by positron emission tomography is described. Two different synthesis routes were envisaged. The first approach followed direct radiofluorination of respective nitro‐ and trimethylammonium substituted benzamides as labeling precursors with no‐carrier‐added (n.c.a.) [18F]fluoride. A second synthesis route was based on the acylation reaction of 2‐aminothiazole derivative with labeling agent [18F]SFB. Direct radiofluorination afforded 18 F‐labeled CDK‐2 inhibitor in very low yields of 1%–3%, whereas acylation reaction with [18F]SFB gave 18 F‐labeled CDK‐2 inhibitor [18 F]2 in high yields of up to 85% based upon [18 F]SFB during the optimization experiments. Large scale preparation afforded radiotracer [18 F]2 in isolated radiochemical yields of 37%–44% (n = 3, decay‐corrected) after HPLC purification within 75 min based upon [18 F]SFB. This corresponds to a decay‐corrected radiochemical yield of 13%–16% based upon [18F]fluoride. The radiochemical purity exceeded 95% and the specific activity was determined to be 20 GBq/µmol. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

7.
N‐(3‐[18F]fluoropropyl)‐2β‐carbomethoxy‐3β‐(4‐iodophenyl)nortropane ([18F]FP‐β‐CIT) was synthesized in a two‐step reaction sequence. In the first reaction, 1‐bromo‐3‐(nitrobenzene‐4‐sulfonyloxy)‐propane was fluorinated with no‐carrier‐added fluorine‐18. The resulting product, 1‐bromo‐3‐[18F]‐fluoropropane, was distilled into a cooled reaction vessel containing 2β‐carbomethoxy‐3β‐(4‐iodophenyl)‐nortropane, diisopropylethylamine and potassium iodide. After 30 min, the reaction mixture was subjected to a preparative HPLC purification. The product, [18F]FP‐β‐CIT, was isolated from the HPLC eluent with solid‐phase extraction and formulated to yield an isotonic, pyrogen‐free and sterile solution of [18F]FP‐β‐CIT. The overall decay‐corrected radiochemical yield was 25 ± 5%. Radiochemical purity was > 98% and the specific activity was 94 ± 50 GBq/µmol at the end of synthesis. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

8.
A reaction pathway via oxidation of [18F]fluorobenzaldehydes offers a very useful tool for the no‐carrier‐added radiosynthesis of [18F]fluorophenols, a structural motive of several potential radiopharmaceuticals. A considerably improved chemoselectivity of the Baeyer‐Villiger oxidation (BVO) towards phenols was achieved, employing 2,2,2‐trifluoroethanol as reaction solvent in combination with Oxone or m‐CPBA as oxidation agent. The studies showed the necessity of H2SO4 addition, which appears to have a dual effect, acting as catalyst and desiccant. For example, 2‐[18F]fluorophenol was obtained with a RCY of 97% under optimised conditions of 80°C and 30‐minute reaction time. The changed performance of the BVO, which is in agreement with known reaction mechanisms via Criegee intermediates, provided the best results with regard to radiochemical yield (RCY) and chemoselectivity, i.e. formation of [18F]fluorophenols rather than [18F]fluorobenzoic acids. Thus, after a long history of the BVO, the new modification now allows an almost specific formation of phenols, even from electron‐deficient benzaldehydes. Further, the applicability of the tuned, chemoselective BVO to the n.c.a. level and to more complex compounds was demonstrated for the products n.c.a. 4‐[18F]fluorophenol (RCY 95%; relating to 4‐[18F]fluorobenzaldehyde) and 4‐[18F]fluoro‐m‐tyramine (RCY 32%; relating to [18F]fluoride), respectively.  相似文献   

9.
We have developed an efficient synthesis method for the rapid and high‐yield automated synthesis of 4‐(2′‐methoxyphenyl)‐1‐[2′‐(N‐2″‐pyridinyl)‐p‐[18F]fluorobenzamido]ethylpiperazine (p‐[18F]MPPF). No‐carrier‐added [18F]F? was trapped on a small QMA cartridge and eluted with 70% MeCN(aq) (0.4 mL) containing Kryptofix 222 (2.3 mg) and K2CO3 (0.7 mg). The nucleophilic [18F]fluorination was performed with 3 mg of the nitro‐precursor in DMSO (0.4 mL) at 190 °C for 20 min, followed by the preparative HPLC purification (column: COSMOSIL Cholester, Nacalai Tesque, Kyoto, Japan; mobile phase: MeCN/25 mm AcONH4/AcOH = 200/300/0.15; flow rate: 6.0 mL/min) to afford p‐[18F]MPPF (retention time = 9.5 min). p‐[18F]MPPF was obtained automatically with a radiochemical yield of 38.6 ± 5.0% (decay corrected, n = 5), a specific activity of 214.3 ± 21.1 GBq/µmol, and a radiochemical purity of >99% within a total synthesis time of about 55 min. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

10.
Reductive coupling reactions between 4‐[18F]fluoro‐benzaldehyde ([18F] 1 ) and different alcohols by use of decaborane (B10H14) as reducing agent have the potential to synthesize 4‐[18F]fluoro‐benzylethers in one step. [18F] 1 was synthesized from 4‐trimethylammonium benzaldehyde (triflate salt) via a standard fluorination procedure (K[18F]F/Kryptofix® 222) in dimethylformamide at 90°C for 25 min and purified by solid‐phase extraction. Subsequently, reductive etherifications of [18F] 1 were performed as one‐step reactions with primary and secondary alcohols, mediated by B10H14 in acetonitrile at 60°C. Various 4‐[18F]fluorobenzyl ethers (6 examples are shown) were obtained within 1–2 h reaction time in decay‐corrected radiochemical yields of 12–45%. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

11.
Substitution of a halo atom (chloro or bromo) in easily prepared N‐haloacetyl‐anilines with no‐carrier added (NCA) cyclotron‐produced [18F]fluoride ion (18F, t1/2= 109.8 min; β+=96.9%), followed by reduction with borane–tetrahydrofuran (BH3–THF), provides an alternative route to NCA [18F]N‐(2‐fluoroethyl)‐anilines. This two‐step and one‐pot process is rapid (~50 min) and moderately high yielding (~40% decay‐corrected radiochemical yield (RCY) overall). In the nucleophilic substitution reaction, 18‐crown‐6 is preferred to Kryptofix® 222 as complexing agent for the solubilization of the counter‐ion (K+), derived from an added metal salt, in acetonitrile. Weakly basic potassium bicarbonate is preferred as the added metal salt. Inclusion of a small amount of water, equating to 4–5 molar equivalents relative to 18‐crown‐6, base or precursor (held in equimolar ratio), is beneficial in preventing the adsorption of radioactivity onto the wall of the glass reaction vessel and for achieving high RCY in the nucleophilic substitution reaction. BH3–THF is effective for the rapid reduction of the generated [18F]N‐fluoroacetyl‐aniline to the [18F]N‐(2‐fluoroethyl)‐aniline. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

12.
There is still no efficient fluorine‐18‐labeled dopamine D3 subtype selective receptor ligand for studies with positron emission tomography. We aim at improving the D3 selectivity and hydrophilicity of a candidate ligand by changing the substitution pattern to a 2,3‐dichlorophenylpiperazine and hydroxylation of the butyl chain. The compound [18F]3 exhibited D3 affinity of Ki = 3.6 nM, increased subtype selectivity (Ki(D2/D3) = 60), and low affinity to 5‐HT1A and α1 receptors (Ki (5‐HT1A/D3) = 34; Ki1/D3) = 100). The two‐step radiosynthesis was optimized for analog [18F]4 by reducing the necessary concentration of the precursor amine (57 mM), which reacted with [18F]fluorophenylazocarboxylic tert‐butylester under basic conditions. The optimization of the base (Cs 2CO3, 23 mM) and the adjustment of reaction temperature led to the radiochemical yield of 63% after 5 min at 35°C. The optimized reaction conditions were transferred on to the synthesis of [18F]3 with an overall non‐decay corrected yield of 8‐12% in a specific activity of 32‐102 GBq/µmol after a total synthesis time of 30‐35 min. This provides a D 3 radioligand candidate with improved attributes concerning selectivity and radiosynthesis for further preclinical studies.  相似文献   

13.
The availability of no‐carrier‐added (n.c.a.) 1‐bromo‐4‐[18F]fluorobenzene with high radiochemical yields is important for 18F‐arylation reactions using metallo‐organic 4‐[18F]fluorophenyl compounds (e.g. of lithium or magnesium) or Pd‐catalyzed coupling. In this study, different methods for the preparation of 1‐bromo‐4‐[18F]fluorobenzene by nucleophilic aromatic substitution reactions using n.c.a. [18F]fluoride were examined. Of six pathways compared, symmetrical bis‐(4‐bromphenyl)iodonium bromide proved most useful to achieve the title compound in a direct, one‐step nucleophilic substitution with a radiochemical yield (RCY) of 65% within 10 min. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

14.
The aim of this study was to develop a direct fluorination method for the preparation of [18F]‐(E)‐N‐(3‐iodoprop‐2‐enyl)‐2β‐carbofluoroethoxy‐3β‐(4′‐methyl‐phenyl)nortropane ([18F]FE‐PE2I (VI). The synthesis procedure relies on the conventional Kryptofix‐mediated nucleophilic 18F‐substitution of the tosylate group in the precursor, TsOE‐PE2I (V). Out of reaction conditions tested, the highest fluorination efficiency was obtained in dimethyl sulfoxide at 140°C. The reaction mixture was purified by semi‐preparative HPLC, followed‐up by a standard Sep‐Pak SPE procedure. On average, 1.0 GBq of [18F]FE‐PE2I was produced from 5‐min irradiation at 35 μA (dimethyl sulfoxide, 5 min/140°C). Decay‐uncorrected yield of the product after HPLC purification and formulation was in the order of 20%. Specific radioactivity of [18F]FE‐PE2I at 15 min after EOS was 3.3–5.1 Ci/µmol (n = 3); radiochemical purity was >98% (n = 4). This direct nucleophilic fluorination strategy is well suited for the automation of the entire synthesis of [18F]FE‐PE2I in a modern PET synthesizer for human PET application. In addition, the 18F‐incorporation rate into TsOE‐PE2I was evaluated using radio‐thin layer chromatography (TLC) and radio‐HPLC. The suggested HPLC method (ACE 5 C18‐HL column and acetonitrile/0.1 M NH4CO2 (80:20)) was found to be suitable for evaluation of ‘free’ 18F‐fluoride in the reaction mixture; in addition, this method allowed the detection of three radiolabelled by‐products that were not discernable with the TLC approach. Therefore, we conclude that the HPLC approach may serve as a good alternative to traditional radio‐TLC technique as it provides more detailed information about the fluorination process in the reaction kinetics or optimization studies. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

15.
Fluorine‐18 labeled (2S,4S)‐4‐fluoro‐l ‐proline (cis‐[18F]4‐FPro) has been reported to be a potential positron emission tomography tracer to study abnormal collagen synthesis occurring in pulmonary fibrosis, osteosarcomas, mammary and colon carcinomas. In this paper, we report the stereospecific radiofluorination of (2S,4R)‐N‐tert‐butoxycarbonyl‐4‐(p‐toluenesulfonyloxy) proline methyl ester (at 110°C) to produce diastereomerically pure cis‐[18F]4‐FPro in 38% radiochemical yield at the end of a 90‐min synthesis. Investigation of the effect of temperature on the stereospecificity of nucleophilic fluorination showed that diasteriomerically pure cis‐[18F]4‐FPro or trans‐[18F]4‐FPro was produced at lower temperatures (85°C–110°C) during the fluorination of (2S,4R) or (2S,4S) precursors, respectively. However, at higher temperatures (130°C–145°C), fluorination of (2S,4R) precursor produced a mixture of cis‐[18F]4‐FPro and trans‐[18F]4‐FPro diastereomers with cis‐[18F]4‐FPro as the predominant isomer. Hydrolysis of the purified fluorinated intermediate was carried out either in one step, using 2 m triflic acid at 145°C for 10 min, or in two steps where the intermediate was heated in 1 m HCl at 110°C for 10 min followed by stirring at room temperature in 1 N NaOH for 5 min. The aqueous hydrolysis mixture was loaded onto an anion exchange column (acetate form for one‐step hydrolysis) or an ion retardation column (two‐step hydrolysis) followed by a C18 Sep‐Pak® (Waters Corporation, Milford, MA, USA). Pure cis‐[18F]4‐FPro was then eluted with sterile water. We also report that epimerization of cis‐[18F]4‐FPro occurs during the two‐step hydrolysis (H+ followed by OH?) of the intermediate, resulting in 5 ± 3% trans‐[18F]4‐FPro, whereas the one‐step acid hydrolysis yielded pure cis‐[18F]4‐FPro in the final product. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

16.
As model reactions for the introduction of 18F into protected aromatic amino acids, the replacement of NO2, Cl, Br and F by [18F]fluoride in 2‐isophthalaldehyde and 2‐terephthalaldehyde derivatives which model 18F‐DOPA and 18F‐tyrosine was investigated by comparing labelling yields and reaction rates with those of corresponding mono‐aldehyde compounds. All isophthalaldehydes showed maximum radiochemical yields (79 to 86%) at 140°C and in comparison with the corresponding mono‐aldehydes the reaction proceeded faster. At lower temperature the reaction already resulted in high yields, e.g. 2‐nitroisophthalaldehyde was labelled with a yield of 78% at 25°C after 7 min, whereas 2‐nitrobenzaldehyde only reached a yield of 1.7% under the same reaction conditions. The 18F/NO2 exchange in nitroterephthalaldehydes proceeded more slowly and with lower radiochemical yields when compared with corresponding isophthalaldehydes and monoaldehydes. The decarbonylation of 18F‐labelled aromatic dialdehydic compounds with 4 eq. of Wilkinson's catalyst at 150°C in benzonitrile resulted in high yields, e.g. 2‐[18F]fluoro‐5‐methoxyisophthalaldehyde and 4‐[18F]fluoro‐2‐methoxy‐5‐methylisophthalaldehyde were decarbonylated efficiently with yields of 67±3% and 72±2%, respectively. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

17.
The first application of the Horner–Wadsworth–Emmons reaction in 18F‐chemistry is described. This carbonyl‐olefination reaction was performed via a ‘multi‐step/one‐pot’ reaction by the coupling of benzylic phosphonic acid esters (3,5‐bis‐methoxymethoxybenzyl)‐phosphonic acid diethyl ester 2e , (4‐methoxy‐methoxybenzyl)‐phosphonic acid diethyl ester 3e and (4‐dimethyl‐aminobenzyl)phosphonic acid diethyl ester 4d ) with 4‐[18F]fluorobenzaldehyde to give the corresponding 18F‐labelled stilbenes [18F]2g , [18F]3g and [18F]4e exclusively as the expected E‐isomers. The radiochemical yields ranged from 9% to 22% (based upon [18F]fluoride, including HPLC purification). The specific activity reached up to 90 GBq/µmol. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

18.
Although 3′‐deoxy‐3′‐[18F]fluorothymidine ([18F]FLT) is a prospective radiopharmaceutical for the imaging of proliferating tumor cell, it is difficult to prepare large amount of [18F]FLT. We herein describe the preparation of [18F]FLT in an ionic liquid, [bmim][OTf] (1‐butyl‐3‐methyl‐imidazolium trifluoromethanesulfonate). At optimized condition, [18F]fluorinationin ionic liquid with 5 µl of 1 M KHCO3 and 5 mg of the precursor yielded 61.5 ± 4.3% (n=10). Total elapsed time was about 70 min including HPLC purification. The rapid synthesis of [18F]FLT can be achieved by removing all evaporation steps. Overall radiochemical yield and radiochemical purity were 30 ± 5% and >95%, respectively. This method can use a small amount of a nitrobenzenesulfonate precursor and can be adapted for automated production. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

19.
Three 18F‐labelled PET tracers, 2‐[18F]fluoroethyl 1‐[(1R)‐1‐phenylethyl]‐1H‐imidazole‐5‐carboxylate ([18F]FETO), 6‐[(S)‐(4‐chlorophenyl)‐(1H)‐1,2,4‐triazol‐1‐yl)methyl]‐1‐(2‐[18F]fluoroethyl)‐1H‐benzotriazole ([18F]FVOZ) and 7‐[2‐(2‐[18F]fluoroethoxy)ethoxy]‐1‐9H‐ β ‐carboline ([18F]FHAR) were synthesized by a one‐step nucleophilic fluorination using the automated commercial platform TRACERLab FXFN. The labelled products were obtained with 16–20% isolated decay corrected radiochemical yields after 70–75 min synthesis time. The radiochemical and chemical purities were more than 98% in all cases. The synthesis using commercial platform may make these tracers more accessible for clinical research. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

20.
2‐exo‐(2′‐Fluoro‐3′‐(4‐fluorophenyl)‐pyridin‐5′‐yl)‐7‐azabicyclo[2.2.1]heptane (F2PhEP), a novel, epibatidine‐based, α4β2‐selective nicotinic acetylcholine receptor antagonist of low toxicity, as well as the corresponding N‐Boc‐protected chloro‐ and bromo derivatives as precursors for labelling with fluorine‐18 were synthesized from 7‐tert‐butoxycarbonyl‐7‐azabicyclo[2.2.1]hept‐2‐ene in 13, 19 and 8% overall yield, respectively. [18F]F2PhEP was prepared in 8–9% overall yield (non‐decay‐corrected) using 1 mg of the bromo derivative in the following two‐step radiochemical process: (1) no‐carrier‐added nucleophilic heteroaromatic ortho‐radiofluorination with the activated K[18F]F‐Kryptofix®222 complex in DMSO using microwave activation at 250 W for 90 s, followed by (2) quantitative TFA‐induced removal of the N‐Boc protective group. Radiochemically pure (>95%) [18F]F2PhEP (1.48–1.66 GBq, 74–148 GBq/µmol) was obtained after semi‐preparative HPLC (Symmetry® C18, eluent aqueous 0.05 M NaH2PO4 CH3CN: 78/22 (v:v)) in 75–80 min starting from an 18.5 GBq aliquot of a cyclotron‐produced [18F]fluoride production batch. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号