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 0 毫秒
1.
4‐[18F]Fluoroiodobenzene ([18F]FIB) is a versatile building block in 18F radiochemistry used in various transition metal‐mediated C–C and C–N cross‐coupling reactions and [18F]fluoroarylation reactions. Various synthesis routes have been described for the preparation of [18F]FIB. However, to date, no automated synthesis of [18F]FIB has been reported to allow access to larger amounts of [18F]FIB in high radiochemical and chemical purity. Herein, we describe an automated synthesis of no‐carrier‐added [18F]FIB on a GE TRACERlab? FX automated synthesis unit starting from commercially available (4‐iodophenyl)diphenylsulfonium triflate as the labelling precursor. [18F]FIB was prepared in high radiochemical yields of 89 ± 10% (decay‐corrected, n = 7) within 60 min, including HPLC purification. The radiochemical purity exceeded 95%, and specific activity was greater than 40 GBq/µmol. Typically, from an experiment, 6.4 GBq of [18F]FIB could be obtained starting from 10.4 GBq of [18F]fluoride. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

2.
[18F]2‐Fluoroethyl‐p‐toluenesulfonate also called [18F]2‐fluoroethyl tosylate has been widely used for labeling radioligands for positron emission tomography (PET). [18F]2‐Fluoroethyl‐4‐bromobenzenesulfonate, also called [18F]2‐fluoroethyl brosylate ([18F]F(CH2)2OBs), was used as an alternative radiolabeling agent to prepare [18F]FEOHOMADAM, a fluoroethoxy derivative of HOMADAM, by O‐fluoroethylating the phenolic precursor. Purified by reverse‐phase HPLC, the no‐carrier‐added [18F]F(CH2)2OBs was obtained in an average radiochemical yield (RCY) of 35%. The reaction of the purified and dried [18F]F(CH2)2OBs with the phenolic precursor was performed by heating in DMF and successfully produced [18F]FEOHOMADAM, after HPLC purification, in RCY of 21%. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

3.
The scope of the nucleophilic aromatic ortho‐fluorinations from the corresponding ortho‐halonitrobenzene precursors (halo‐to‐fluoro substitutions) with no‐carrier‐added [18F]fluoride ion as its activated K[18F]F–K222 complex has been evaluated via the radiosynthesis of ortho‐[18F]fluoronitrobenzene, chosen as a model reaction. The parameters studied include the influence of the leaving group in the ortho position of the phenyl ring (–Cl, –Br, –l), the quantity of precursor used, the type of activation (conventional heating or microwave irradiations), the solvent, the temperature and the reaction time. The iodo‐precursor was completely unreactive and the bromo‐precursor gave only low incorporation (<10%) in the optimal conditions used (conventional heating at 145°C or microwave activation, 100 W for 120 s). Only the chloro‐precursor was found reactive in the conditions described above and up to 70% yield was observed for the formation of ortho‐[18F]fluoronitrobenzene ([18F]‐ 1 ). In all the experiments, the unwanted ortho‐[18F]fluoro‐halobenzenes, potentially resulting from the nitro‐to‐fluoro substitution, could not be detected. These results will be applied for the radiosynthesis of 5‐[18F]fluoro‐6‐nitroquipazine, a potent radioligand for the imaging of the serotonin transporter with PET. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

4.
In an attempt to visualize folate receptors that over‐express on many cancers, [18F]‐fluorobenzene and pyridine carbohydrazide‐folates were synthesized using two different synthetic approaches starting from nucleophilic displacement reactions on ethyl‐trimethylammonium‐benzoate and pyridine carboxylate precursors. The intermediates ethyl [18F]‐fluorinated benzene and pyridine esters were reacted with hydrazine to produce the [18F]‐fluorobenzene and pyridine carbohydrazides followed by coupling with NHS‐folate 11 in the first approach. Whereas hydrazide‐folate 5 was reacted with 2,5‐dioxoazolidinyl [18F]‐fluorobenzenecarboxylate in the second approach. Based on starting [18F]‐fluoride, radiochemical yields and synthesis times were found to be around 80% (45 min) and 35% (80 min) for the first and the second approaches, respectively. The first synthetic approach holds considerable promise as a rapid and simple method for the radiofluorination of folic acid with high radiochemical yield and short time. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

5.
An efficient, fully automated, enantioselective multi‐step synthesis of no‐carrier‐added (nca) 6‐[18F]fluoro‐L‐dopa ([18F]FDOPA) and 2‐[18F]fluoro‐L‐tyrosine ([18F]FTYR) on a GE FASTlab synthesizer in conjunction with an additional high‐ performance liquid chromatography (HPLC) purification has been developed. A PTC (phase‐transfer catalyst) strategy was used to synthesize these two important radiopharmaceuticals. According to recent chemistry improvements, automation of the whole process was implemented in a commercially available GE FASTlab module, with slight hardware modification using single use cassettes and stand‐alone HPLC. [18F]FDOPA and [18F]FTYR were produced in 36.3 ± 3.0 % (n = 8) and 50.5 ± 2.7 % (n = 10) FASTlab radiochemical yield (decay corrected). The automated radiosynthesis on the FASTlab module requires about 52 min. Total synthesis time including HPLC purification and formulation was about 62 min. Enantiomeric excesses for these two aromatic amino acids were always >95 %, and the specific activity of was >740 GBq/µmol. This automated synthesis provides high amount of [18F]FDOPA and [18F]FTYR (>37 GBq end of synthesis (EOS)). The process, fully adaptable for reliable production across multiple PET sites, could be readily implemented into a clinical good manufacturing process (GMP) environment.  相似文献   

6.
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   

7.
No‐carrier‐added (n.c.a.) 2‐[18F]fluoromethyl‐l‐phenylalanine (2‐[18F]FMP) was found to be very sensitive to hydrolysis in aqueous solutions. In this paper, the defluorination reaction was studied in detail to elucidate its mechanism. Therefore, besides 2‐[18F]FMP and 4‐[18F]FMP, 2‐[18F]fluoromethyl‐phenethylamine (2‐[18F]FMPAM) and 4‐[18F]FMPAM were synthesized, both ‘mimetic’ molecules of the decarboxylated amino acid analogues. Radiosynthesis, using a customized Scintomics automatic synthesis hotboxthree module, resulted in a high overall yield and a radiochemical purity of >99%. The defluorination rates of all compounds were studied by HPLC. The defluorination rate of 2‐[18F]FMLP at 50°C was approximately 300 times faster than that of n.c.a. 4‐[18F]FMLP. The defluorination rate of 2‐[18F]FMPAM is somewhat lower than of 2‐[18F]FMP but still very high in comparison with 4‐[18F]FMPAM, which is virtually stable. It allowed to elucidate the reaction mechanism ruled by two distinct intramolecular interactions. First, the hydrogen bond interaction between the amine and the benzylic fluorine weakening the carbon–fluorine bond. Secondly, the formation of a second hydrogen bond between the carboxyl oxygen atom and one of the benzylic hydrogen atoms rendering the benzyl fluoride group even more susceptible to hydrolysis. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

8.
Reproducible methods for [18F]radiolabeling of biological vectors are essential for the development of new [18F]radiopharmaceuticals. Molecules such as carbohydrates, peptides and proteins are challenging substrates that often require multi‐step indirect radiolabeling methods. With the goal of developing more robust, time saving, and less expensive procedures for indirect [18F]radiolabeling of such molecules, our group has synthesized ethynyl‐4‐[18F]fluorobenzene ([18F]2, [18F]EYFB) in a single step (14 ± 2% non‐decay corrected radiochemical yield (ndc RCY)) from a readily synthesized, shelf stable, inexpensive precursor. The alkyne‐functionalized synthon [18F]2 was then conjugated to two azido‐functionalized vector molecules via CuAAC reactions. The first ‘proof of principle’ conjugation of [18F]2 to 1‐azido‐1‐deoxy‐β‐d ‐glucopyranoside (3) gave the desired radiolabeled product [18F]4 in excellent radiochemical yield (76 ± 4% ndc RCY (11% overall)). As a second example, the conjugation of [18F]2 to matrix‐metalloproteinase inhibitor (5), which has potential in tumor imaging, gave the radiolabeled product [18F]6 in very good radiochemical yield (56 ± 12% ndc RCY (8% overall)). Total preparation time for [18F]4 and [18F]6 including [18F]F? drying, two‐step reaction (nucleophilic substitution and CuAAC conjugation), two HPLC purifications, and two solid phase extractions did not exceed 70 min. The radiochemical purity of synthon [18F]2 and the conjugated products, [18F]4 and [18F]6, were all greater than 98%. The specific activities of [18F]2 and [18F]6 were low, 5.97 and 0.17 MBq nmol?1, respectively.  相似文献   

9.
The SUZUKI reaction of organoboron compounds with 4‐[18F]fluoroiodobenzene has been developed as a novel radiolabelling technique in 18F chemistry. The cross‐coupling reaction of p‐tolylboronic acid with 4‐[18F]fluoroiodobenzene was used to screen different palladium complexes, bases and solvents. Optimized reaction conditions (Pd2(dba)3, Cs2CO3, acetonitrile, 60°C for 5 min) were further applied to the synthesis of various 18F‐labelled biphenyls bearing different functional groups. The reaction proceeded in excellent radiochemical yields of up to 94% within 5 min while showing good compatibility to many functional groups. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

10.
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   

11.
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   

12.
18F‐labeled fluorobenzaldehydes and fluorobenzylbromides are useful synthons for the preparation of positron emission tomography radiopharmaceuticals. Although ortho‐ and para‐[18F]fluorobenzaldehydes can easily be prepared with high yields, the corresponding meta‐derivatives are more problematic. In order to improve the yield of meta‐[18F]fluorobenzaldehyde, we used the corresponding diaryliodonium salt precursors, since diaryliodonium salts had already been used as precursors in the preparations of 18F‐labeled electron‐rich, as well as electron‐deficient, aromatic rings. Diaryliodonium salts with different counter ions [PhIPhCHO]X (X = Cl, Br, OTs, OTf) were synthesized. 18F radiolabeling was performed using different bases at different temperatures in the presence of a radical scavenger, 2,2,6,6‐tetramethylpiperidine‐N‐oxyl (TEMPO). The best conversion (~80%) to meta‐[18F]fluorobenzaldehyde was obtained using CsHCO3 base at a reaction temperature of 110°C. To study iodonium salt counter ion effects on radiofluorination, each precursor was separately treated with Cs[18F]F/CsHCO3 in DMF at 110°C for 5 min in the presence of TEMPO. Our observed reactivity order was OTsMeta‐[18F]fluorobenzaldehyde thus obtained was reduced to the corresponding alcohol with aqueous NaBH4 at room temperature and then converted to meta‐[18F]fluorobenzylbromide using triphenylphosphine dibromide. Formation of meta‐[18F]fluorobenzylbromide was confirmed using high‐performance liquid chromatography and the desired product was purified on a silica Sep ‐ Pak® plus cartridge. Published in 2011 by John Wiley & Sons, Ltd.  相似文献   

13.
Compared to homoaromatic and aliphatic nucleophilic radiofluorinations, only few references can be found in the literature describing nucleophilic substitutions with [18F]fluoride ion of heteroaromatic compounds such as pyridines and only reactions involving fluorination processes at the ortho‐position (2‐position) have been more intensively studied. In the present paper, the scope of the nucleophilic aromatic fluorinations at the meta‐ and para‐position of the pyridine ring with no‐carrier‐added [18F]fluoride ion as its activated K[18F]F‐K222 complex has been evaluated and compared to the nucleophilic aromatic fluorinations at the ortho‐position in this pyridine series. The syntheses of 3‐ and 4‐[18F]fluoropyridines were chosen as model reactions and compared to the radiosynthesis of 2‐[18F]fluoropyridine. The parameters studied include the influence of the position of the leaving group at the pyridine ring, as well as the quantity of the precursor used, the type of activation (conventional heating, microwave irradiation), the solvent, the temperature and the reaction time. Using the corresponding nitro precursor, high yields were obtained at the 2‐position (94% yield) using microwaves (100 W) for 2 min in DMSO. Good yields (up to 72%) were observed at the 4‐position using the same conditions while practically no reaction was observed at the 3‐position. About 60% yield was also obtained at both the 2‐ and 4‐position using the corresponding nitro precursor at 145°C for 10 min in DMSO. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

14.
To assess the potential of intermolecular hydroacylation reactions as a new fluorine‐18 labeling method, model reactions of [18F]fluorobenzaldehyde with three different olefins (1‐hexene ( 2a ), allylbenzene ( 2b ), and 3‐phenoxypropene ( 2c )) in the presence of Wilkinson's catalyst were performed. The procedure gave high radiochemical yields (38–62%) of [18F]fluorophenylketones with short reaction times (15 min). The intermolecular hydroacylation reaction provides a new method for the preparation of fluorine‐18 labeled compounds. Copyright © 2002 John Wiley & Sons, Ltd.  相似文献   

15.
Radiolabeled prostate‐specific membrane antigen (PSMA) targeting PET‐tracers have become desirable radiopharmaceuticals for the imaging of prostate cancer (PC). Recently, the PET radiotracer [18F]PSMA‐1007 was introduced as an alternative to [68Ga]Ga‐PSMA‐11, for staging and diagnosing biochemically recurrent PC. We incorporated a one‐step procedure for [18F]PSMA‐1007 radiosynthesis, using both Synthra RNplus and GE TRACERlab FxFN automated modules, in accordance with the recently described radiolabeling procedure. Although the adapted [18F]PSMA‐1007 synthesis resulted in repeatable radiochemical yields (55 ± 5%, NDC), suboptimal radiochemical purities of 87 ± 8% were obtained using both modules. As described here, modifications made to the radiolabeling and the solid‐phase extraction purification steps reduced synthesis time to 32 minutes and improved radiochemical purity to 96.10%, using both modules, without shearing the radiochemical yield.  相似文献   

16.
Four different no carrier added (n.c.a.) 4‐[18F]fluorophenylurea derivatives are synthesized as model compounds via two alternative routes. In both cases carbamate‐4‐nitrophenylesters are used as intermediates. Either n.c.a. 4‐[18F]fluoroaniline reacts with carbamates of several amines, or the carbamate of n.c.a. 4‐[18F]fluoroaniline is formed at first and an amine is added subsequently to yield the urea derivative. The choice of the appropriate way of reaction depends on the possibilities of precursor synthesis. The radiochemical yields reach up to 80% after 50 min of synthesis time while no radiochemical by‐products can be determined. These high yields were possible due to an optimized preparation of n.c.a. 4‐[18F]fluoroaniline with a radiochemical yield of up to 90%. From the various ways of its radiosynthesis, the substitution with n.c.a. [18F]fluoride on dinitrobenzene is chosen, using phosphorous acid and palladium black for reduction of the second nitro group. Copyright © 2006 John Wiley & Sons, Ltd.  相似文献   

17.
Currently there is still a need for more potent amino acid analogues as tumour imaging agents for peripheral tumour imaging with PET as it was recently reported that the success of O‐(2′‐[18F]fluoroethyl)‐L ‐tyrosine ([18F]FET) is limited to brain, head and neck tumours. As the earlier described 2‐Amino‐3‐(2‐[18F]fluoromethyl‐phenyl)‐propionic acid (2‐[18F]FMP) suffered from intramolecular‐catalysed defluorination, we synthesized 2‐Amino‐3‐(4‐[18F]fluoromethyl‐phenyl)‐propionic acid (4‐[18F]FMP) as an alternative for tumour imaging with PET. Radiosynthesis of 4‐[18F]FMP, based on Br for [18F] aliphatic nucleophilic exchange, was performed with a customized modular Scintomics automatic synthesis hotboxthree system in a high overall yield of 30% and with a radiochemical purity of \gt 99%. 4‐[18F]FMP was found to be stable in its radiopharmaceutical formulation, even at high radioactivity concentrations. Additionally, for a comparative study, [18F]FET was synthesized using the same setup in 40% overall yield, with a radiochemical purity \gt 99%. The described automated radiosynthesis allows the production of two different amino acid analogues with minor alternations to the parameter settings of the automated system, rendering this unit versatile for both research and clinical practice. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
4‐[18F]Fluorobenzyltriphenylphosphonium cation (18F‐FBnTP) is a promising negative membrane potential targeting positron emission tomography tracer. However, the reported multistep radiolabeling approach for the synthesis of 18F‐FBnTP poses a challenge for routine clinical applications. In this study, we demonstrated that 18F‐FBnTP can be prepared in good conversion yields (~60%, nondecay corrected) in just one step via a copper‐mediated 18F‐fluorination reaction using a pinacolyl arylboronate precursor. In addition, our data suggest that 18F‐labeled (phosphonium) cations can be efficiently prepared via a copper‐mediated 18F‐fluoronation by using triflate as the counterion.  相似文献   

19.
With the goal of developing a PET radioligand for the in vivo assessment of glucose transport, 6-deoxy-6-[18F]fluoro-D -glucose ([18F]6FDG) was prepared in two steps from 18F. Starting with D -glucose, the tosyl- and mesyl-derivatives of 3,5-O-benzylidene-1,2-O-isopropylidene-α-D -glucofuranose were prepared by known methods. Reaction of either of these precursors with 18F resulted in the formation of 3,5-O-benzylidene-6-deoxy-6-[18F]-fluoro-1,2-O-isopropylidene-α-D -glucofuranose in high yield. Subsequent hydrolysis resulted in the production of [18F]6FDG. Under optimal conditions, [18F]6FDG is produced 60–70 min after end of bombardment (EOB) in 71 ± 12% yield (decay corrected, based upon fluoride) with a radiochemical purity of ⩾96%. Preliminary experiments have indicated that [18F]6FDG may be a more representative in vivo tracer for the glucose transporter than 2FDG. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

20.
A synthesis method has been developed for the labelling of N‐(3‐[18F]fluoropropyl)‐2β‐carbomethoxy‐3β‐(4‐fluorophenyl)nortropane ([18F]β‐CFT‐FP), a potential radioligand for visualization of the dopamine transporters by positron emission tomography. The two‐step synthesis includes preparation of [18F]fluoropropyl tosylate and its use without purification in the fluoroalkylation of 2β‐carbomethoxy‐3β‐(4‐fluorophenyl)nortropane (nor‐β‐CFT). The final product is purified by HPLC. Optimization of the two synthesis steps resulted in a greater than 30% radiochemical yield of [18F]β‐CFT‐FP (decay corrected to end of bombardment). The synthesis time including HPLC‐purification was approximately 90 min. The radiochemical purity of the final product was higher than 99% and the specific radioactivity at the end of synthesis was typically 20 GBq/µmol. In comparison to alkylation by [18F]fluoropropyl bromide, the procedure described here results in an improved overall radiochemical yield of [18F]β‐CFT‐FP in a shorter time. Copyright © 2005 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