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 31 毫秒
1.
The excretion and metabolism of orally administered [14C]-labelled benzoic acid (100 mg/kg) was examined in the reptile Sphenedon punctatus (tuatara). The major excreted metabolite was chromatographically and electrophoretically identical with ornithuric acid. Conjugation with glycine or glucuronic acid was not detected. 7–21 percent of the dose was recovered from the urine and faeces, the bulk of the excreted radioactivity being eliminated in the first seven days. Free benzoic acid and conjugates were excreted in the first week but only conjugates could be detected in fauces collected at later intervals. These results are discussed in relation to the taxonomic position of tuatara.  相似文献   

2.
The comparative toxicity of lactic acid, acetic acid, and benzoic acid to tilapia (Oreochromis mossambicus), cladoceran crustacea (Moina micrura), and oligochaete worm (Branchiura sowerbyi) were determined using static bioassay tests. Worms were found most sensitive to all the acids whereas the cladoceran was found most resistant to lactic acid and the fish most resistant to acetic acid and benzoic acid. The 96h LC50 values of lactic acid, acetic acid, and benzoic acid, were, respectively, 257.73, 272.87, and 276.74 mg L?1 for O. mossambicus; 329.12, 163.72, and 71.65 mg L?1 for M. micrura and 50.82, 14.90, and 39.47 mg L?1 for B. sowerbyi. Tilapia lost appetite at sub-lethal concentrations as low as 2.18 mg L?1 lactic acid, 1.26 mg L?1 acetic acid, and 13.84 mg L? 1 of benzoic acid. Growth and reproduction of the fish were affected following 90-day chronic exposure to sub-lethal concentrations of the acids. Minimum effective concentration of the acids that significantly reduced food conversion efficiency (FCE), percent increase of weight, specific growth rate, yield and fecundity of the fish were 2.18, 1.47, and 3.95 mg · L?1 of lactic acid, acetic acid, and benzoic acid, respectively. Effects of acetic acid and benzoic acid on FCE, weight increase, and yield were not significantly different from each other whereas lactic acid produced different effects from acetic acid as well as benzoic acid. Mean values of dissolved oxygen, primary productivity, and plankton populations of the test medium significantly reduced from control at 16.94 mg L?1 lactic acid, 16.79 mg L?1 acetic acid, and 13.84 mg L?1 benzoic acid.  相似文献   

3.
Atmospheric concentrations of the greenhouse gas nitrous oxide (N2O) have continued to rise since the advent of the industrial era, largely because of the increase in agricultural land use. The urine deposited by grazing ruminant animals is a major global source of agricultural N2O. With the first commitment period for reducing greenhouse gas emissions under the Kyoto Protocol now underway, mitigation options for ruminant urine N2O emissions are urgently needed. Recent studies showed that increasing the urinary concentration of the minor urine constituent hippuric acid resulted in reduced emissions of N2O from a sandy soil treated with synthetic bovine urine, due to a reduction in denitrification. A similar effect was seen when benzoic acid, a product of hippuric acid hydrolysis, was used. This current laboratory experiment aimed to investigate these effects using real cow urine for the first time. Increased concentrations of hippuric acid or benzoic acid in the urine led to reduction of N2O emissions by 65% (from 17% to <6% N applied), with no difference between the two acid treatments. Ammonia volatilization did not increase significantly with increased hippuric acid or benzoic acid concentrations in the urine applied. Therefore, there was a net reduction in gaseous N loss from the soil with higher urinary concentrations of both hippuric acid and benzoic acid. The results show that elevating hippuric acid in the urine had a marked negative effect on both nitrification and denitrification rates and on subsequent N2O fluxes. This study indicates the potential for developing a novel mitigation strategy based on manipulation of urine composition through ruminant diet.  相似文献   

4.
Khurana JP  Cleland CF 《Plant physiology》1992,100(3):1541-1546
Lemna paucicostata LP6 does not normally flower when grown on basal Bonner-Devirian medium, but substantial flowering is obtained when 10 μm salicylic acid (SA) or benzoic acid is added to the medium. Benzoic acid is somewhat more effective than SA, and the threshold level of both SA and benzoic acid required for flower initiation is reduced as the pH of the medium is lowered to 4.0. SA- or benzoic acid-induced flowering is enhanced in the simultaneous presence of 6-benzylaminopurine (BAP), although BAP per se does not influence flowering in strain LP6. Continuous presence of SA or benzoic acid in the culture medium is essential to obtain maximal flowering. A short-term treatment of the plants (for first 24 h) with 10 μm SA or benzoic acid, followed by culture in the basal medium containing 1 μm BAP can, however, stimulate profuse flowering. Benzoic acid is more effective than SA, and the effect is more pronounced at pH 4 than at 5.5. Thus, under these conditions, flowering is of an inductive nature. Experiments with [14C]SA and [14C]benzoic acid have provided evidence that at pH 4 there is relatively more uptake of benzoic acid than SA, thus leading to an increased flowering response. The data obtained from the experiments designed to study the mobility of [14C]SA and [14C]-benzoic acid from mother to daughter fronds indicate that there is virtually no mobility of SA or benzoic acid between fronds.  相似文献   

5.
Metabolism of benzoic acids and benzaldehydes are crucial to produce hormones, defense compounds and attractants for pollinators in plants. Tricholoma matsutake contains benzoic acid and benzaldehyde, but their roles have not been fully studied. First we conducted tracer experiments to gain insight into benzoic acid and benzaldehyde biosynthesis in T. matsutake. 13C and 2H were incorporated into benzoic acid from uniformly 13C- and 2H- labelled l-phenylalanine and (E)-cinnamate 1?d after supplementation of the precursors without any substitution of 13C and 2H. In contrast, no 13C and 2H were incorporated into benzaldehyde from these precursors 10?d after the supplementation. The results indicate that T. matsutake has a metabolic pathway to biosynthesize benzoic acid from l-phenylalanine and (E)-cinnamate in which benzaldehyde is not a metabolic intermediate. However, 30?d after the supplementation of 13C- and 2H- labelled l-phenylalanine, 13C and 2H were incorporated into all the carbon and hydrogen atoms of benzaldehyde. In addition, 2H was not incorporated into benzaldehyde from exogenously supplemented 2H-labelled benzoic acid. This result indicates that T. matsutake has a metabolic pathway to biosynthesize benzaldehyde from l-phenylalanine in which benzoic acid is not an intermediate. Taken together, our results strongly suggest that T. matsutake biosynthesizes benzoic acid and benzaldehyde in separate pathways.  相似文献   

6.
The described method permits the determination of the five most important metabolites of the pyrethroids permethrin, cypermethrin, deltamethrin, λ-cyhalothrin, fenvalerate, phenothrin and β-cyfluthrin in human urine in one run. The major urinary metabolites of these substances are cis-3-(2,2-dichlorovinyl)-2,2-dimethylcyclopropane-1-carboxylic acid (cis-Cl2CA), trans-3-(2,2-dichlorovinyl)-2,2-dimethylcyclopropane-1-carboxylic acid (trans-Cl2CA), cis-3-(2,2-dibromovinyl)-2,2-dimethylcyclopropane-1-carboxylic acid (Br2CA), fluoro-3-phenoxybenzoic acid (F-PBA) and 3-phenoxybenzoic acid (3-PBA). After acidic hydrolysis to release the conjugated carboxylic acid metabolites, the analytes were separated from the matrix by means of solid-phase extraction using a reversed-phase column. The components of the eluate were converted to their methyl esters and extracted in hexane. Separation and quantitative analysis of the pyrethroid metabolites was carried out by capillary gas chromatography and mass selective detection. 2-Phenoxybenzoic acid served as an internal standard. The detection limits lay between 0.3 and 0.5 μg per litre urine. The relative standard deviations of the within-series imprecision were between 1% and 6%. The relative recovery rates ranged between 90% and 98%. Using this method we determined the elimination of pyrethroid metabolites in 24-h urine samples from eight pest controllers after indoor application of permethrin. The detected concentrations ranged from 1 to 70 μg g−1 creatinine.  相似文献   

7.
Previous genetic and proteomic studies identified altered activity of various enzymes such as those of fatty acid metabolism and glycogen synthesis after a single toxic dose of valproic acid (VPA) in rats. In this study, we demonstrate the effect of VPA on metabolite synthesis flux rates and the possible use of abnormal 13C labeled glucose-derived metabolites in plasma or urine as early markers of toxicity. Female CD-1 mice were injected subcutaneously with saline or 600 mg/kg) VPA. Twelve hours later, the mice were injected with an intraperitoneal load of 1 g/kg [U-13C]-d-glucose. 13C isotopomers of glycogen glucose and RNA ribose in liver, kidney and brain tissue, as well as glucose disposal via cholesterol and glucose in the plasma and urine were determined. The levels of all of the positional 13C isotopomers of glucose were similar in plasma, suggesting that a single VPA dose does not disturb glucose absorption, uptake or hepatic glucose metabolism. Three-hour urine samples showed an increase in the injected tracer indicating a decreased glucose re-absorption via kidney tubules. 13C labeled glucose deposited as liver glycogen or as ribose of RNA were decreased by VPA treatment; incorporation of 13C via acetyl-CoA into plasma cholesterol was significantly lower at 60 min. The severe decreases in glucose-derived carbon flux into plasma and kidney-bound cholesterol, liver glycogen and RNA ribose synthesis, as well as decreased glucose re-absorption and an increased disposal via urine all serve as early flux markers of VPA-induced adverse metabolic effects in the host.  相似文献   

8.
Co‐permeability of 3H‐labelled water and 14C‐labelled benzoic acid or 2,4‐dichlorophenoxyacetic acid across isolated cuticular membranes of Prunus laurocerasus L. was measured at temperatures ranging from 15 to 50 °C. The water and benzoic acid permeances were highly correlated over the whole temperature range investigated, whereas water and 2,4‐dichlorophenoxyacetic acid permeances were only correlated between 15 and 30 °C. The activation energies of cuticular permeability calculated from Arrhenius plots were 40 kJ mol?1 for water and benzoic acid and 115 kJ mol?1 for 2,4‐dichlorophenoxyacetic acid. The slopes of the Arrhenius plots of 2,4‐dichlorophenoxyacetic acid were linear between 15 and 50 °C, whereas pronounced phase transitions around 30 °C were observed for water and benzoic acid permeability. However, with isolated polymer matrix membranes, where cuticular waxes forming the transport‐limiting barrier of cuticles have been extracted, phase transitions were not observed for water and benzoic acid. It is concluded that temperatures above 30 °C caused structural changes in the transport‐limiting barrier of the cuticles leading to additional paths of diffusion for water and benzoic acid but not for 2,4‐dichlorophenoxyacetic acid.  相似文献   

9.
The white rot fungus Phanerochaete chrysosporium has the largest cytochrome P450 contingent known to date in fungi, but the study on the function of these P450s is limited. In this study, induction of functional P450 in P. chrysosporium was first shown and P450-mediate degradation of benzoic acid was demonstrated in this fungus. Carbon monoxide difference spectra indicated significant induction of P450 by benzoic acid, m-chlorobenzoic acid, p-chlorobenzoic acid and n-hexane, and showed the effect of inducer concentration and nutrient condition on the induction of P450. The high contents of P450 in the microsomal fractions facilitated the study on the function of P450. While the n-hexane-induced P450 could not interact with benzoic acid, the microsomal P450 induced by benzoic acid produced type I substrate binding spectra upon the addition of benzoic acid. The benzoic acid degradation by the microsomal P450 was NADPH-dependent at a specific rate of 194 ± 14 min−1, and significantly inhibited by piperonyl butoxide (a P450 inhibitor). However, inhibition of benzoic acid degradation by piperonyl butoxide was slight or not detectable in the cultures of this fungus, suggesting presumable involvement of other enzyme in benzoic acid degradation. The extracellular ligninolytic enzymes, lignin peroxidase and manganese-dependent peroxidase, were not involved in initial metabolism of benzoic acid under the test conditions.  相似文献   

10.
Biotransformation of [1-6-14C]benzene and [1-14C]toluene in English ryegrass (Lolium perenne L.) seedlings was investigated. Vapors of these compounds were absorbed by the leaves of this plant. Benzene and toluene were oxidized, forming phenol and benzoic acid, respectively. A portion of phenol and benzoic acid was bound by low-molecular-weight peptides forming conjugates. A qualitative amino acid composition of the peptides involved in the conjugation was determined. After removing plants from the atmosphere containing [1-6-14C]benzene and [1-14C]toluene, the radioactivity of the conjugates gradually decreased. This process was accompanied by the evolution of 14CO2, indicating the breakdown of these conjugates. Radioactive compounds thus formed were oxidized, yielding carbon dioxide. A portion of phenol and benzoic acid, along with peptide conjugation, was subjected to further oxidative transformations up to disruption of the aromatic ring. By this pathway, nonvolatile carboxylic acids, such as muconic, fumaric, succinic, malic, malonic, glycolic, and glyoxylic, were formed. Using electron microscopy, a damaging effect of benzene on the cell ultrastructure of English ryegrass leaves was shown, and this toxic effect depended on the benzene concentration.  相似文献   

11.
Halothane (CF3CHClBr), a widely used volatile anesthetic, undergoes extensive biotransformation in humans. Oxidative halothane metabolism yields the stable metabolites trifluoroacetic acid and bromide which can be detected in plasma and urine. To date, analytical methodologies have either required extensive sample preparation, or two separate analytical procedures to determine plasma and urine concentrations of these analytes. A rapid and sensitive method utilizing high-performance liquid chromatography-ion chromatography (HPLC-IC) with suppressed conductivity detection was developed for the simultaneous detection of both trifluoroacetic acid and bromide in plasma and urine. Sample preparation required only ultrafiltration. Standard curves were linear (r2≥0.99) from 10 to 250 μM trifluoroacetic acid and 2 to 5000 μM bromide in plasma and 10 to 250 μM trifluoroacetic acid and 2 to 50 μM bromide in urine. The assay was applied to quantification of trifluoroacetic acid and bromide in plasma and urine of a patient undergoing halothane anesthesia.  相似文献   

12.
Analytical methods are described for the selective, rapid and sensitive determination of R- and S-apomorphine, apocodeine and isoapocodeine and the glucuronic acid and sulfate conjugates in plasma and urine. The methods involve liquid-liquid extraction followed by high-performance liquid chromatography with electrochemical detection. The glucuronide and sulfate conjugates are determined after enzymatic hydrolysis. For the assay of R- and S-apomorphine a 10 μm Chiralcel OD-R column is used and the voltage of the detector is set at 0.7 V. The mobile phase is a mixture of aqueous phase (pH 4.0)-acetonitrile (65:35, v/v). At a flow-rate of 0.9 ml min−1 the total run time is ca. 15 min. The detection limits are 0.3 and 0.6 ng ml−1 for R- and S- apomorphine, respectively (signal-to-noise ratio 3). The intra- and inter-assay variations are <5% in the concentration range of 2.5-25 ng ml−1 for plasma samples, and <4% in the concentration range of 40-400 ng ml−1 for urine samples. For the assay of apomorphine, apocodeine and isoapocodeine, a 5 μm C18 column was used and the voltage of the detector set at 0.825 V. Ion-pairing chromatography was used. The mobile phase is a mixture of aqueous phase (pH 3.0)-acetonitrile (75:25, v/v). At a flow-rate of 0.8 ml min−1 the total run time is ca. 14 min. The detection limits of this assay are 1.0 ng ml−1 for apomorphine and 2.5 ng ml−1 for both apocodeine and isoapocodeine (signal-to-noise ratio 3). The inter-assay variations are 5% in the concentration range of 5-40 ng ml−1 for plasma samples and 7% in the concentration range of 50-500 ng ml−1 for urine samples. The glucuronic acid and sulfate conjugates of the various compounds are hydrolysed by incubation of the samples with β-glucuronidase and sulfatase type H-1, respectively. Hydrolysis was complete after 5 h of incubation. No measurable degradation of apomorphine, apocodeine and isoapocodeine occurred during the incubation. A pharmacokinetic study of apomorphine, following the intravenous infusion of 30 μg kg−1 for 15 min in a patient with Parkinson's disease, demonstrates the utility of the methods: both the pharmacokinetic parameters of the parent drug and the appearance of apomorphine plus metabolites in urine could be determined.  相似文献   

13.
Biliary phospholipids have been hypothesized to be important for essential fatty acid homeostasis. We tested this hypothesis by investigating the intestinal absorption and the status of linoleic acid in mdr2 Pgp-deficient mice which secrete phospholipid-free bile. In mice homozygous (?/?) for disruption of the mdr2 gene and wild-type (+/+) mice, dietary linoleic acid absorption was determined by 72 h balance techniques. After enteral administration, [13C]-linoleic acid absorption was determined by measuring [13C]-linoleic acid concentrations in feces and in plasma. The status of linoleic acid was determined in plasma and in liver by calculating the molar percentage of linoleic acid and the triene:tetraene ratio. Although plasma concentration of [13C]-linoleic acid at 2 h after enteral administration was significantly lower in (?/?) compared to (+/+) mice (P≤0.05), net intestinal absorption of dietary linoleic acid or of [13C]-linoleic acid was similar in (+/+) and (?/?) mice. Molar percentage of linoleic acid and the triene:tetraene ratio were not different in whole plasma or in liver of (?/?) compared to (+/+) mice. Present data indicate that biliary phospholipids are involved in the rate of appearance in plasma of enterally administered linoleic acid, but are not required for net intestinal absorption or plasma status of linoleic acid.  相似文献   

14.
Heavy chain disease proteins (FOR) were isolated from human plasma. These proteins were also detected immunochemically in the urine of the patient. The proteins were disulphide-linked Fc-like dimers with molar mass 64.2 kg/mol and sedimentation rate S 20,w 0 = = 0.356 ps (3.56 S). Similar amounts of aspartic acid and pyroglutamic acid were found at the N-terminus. After cyanogen bromide cleavage of the FOR proteins, three peptides were isolated and their amino acid composition and partial amino acid sequence was determined. We suggest that two Fc-like proteins of similar sizes are present in the plasma: (1) the first with N-terminal aspartic acid corresponding to position 221 of γ1 EU chain and (2) the second with N-terminal pyroglutamic acid. The first protein and small amounts of related low-molar mass fragments found also in the plasma could be degradation products of the second protein. Evidence is given on structural differences between the FOR proteins and the corresponding portion of the γ1 EU chain.  相似文献   

15.
The fate of benzoic acid in various species   总被引:4,自引:2,他引:2       下载免费PDF全文
1. The urinary excretion of orally administered [14C]benzoic acid in man and 20 other species of animal was examined. 2. At a dose of 50mg/kg, benzoic acid was excreted by the rodents (rat, mouse, guinea pig, golden hamster, steppe lemming and gerbil), the rabbit, the cat and the capuchin monkey almost entirely as hippuric acid (95–100% of 24h excretion). 3. In man at a dose of 1mg/kg and the rhesus monkey at 20mg/kg benzoic acid was excreted entirely as hippuric acid. 4. At 50mg/kg benzoic acid was excreted as hippuric acid to the extent of about 80% of the 24h excretion in the squirrel monkey, pig, dog, ferret, hedgehog and pigeon, the other 20% being found as benzoyl glucuronide and benzoic acid, the latter possibly arising by decomposition of the former. 5. On increasing the dose of benzoic acid to 200mg/kg in the ferret, the proportion of benzoyl glucuronide excreted increased and that of hippuric acid decreased. This did not occur in the rabbit, which excreted 200mg/kg almost entirely as hippuric acid. It appears that the hedgehog and ferret are like the dog in respect to their metabolism of benzoic acid. 6. The Indian fruit bat produced only traces of hippuric acid and possibly has a defect in the glycine conjugation of benzoic acid. The main metabolite in this animal (dose 50mg/kg) was benzoyl glucuronide. 7. The chicken, side-necked turtle and gecko converted benzoic acid mainly into ornithuric acid, but all three species also excreted smaller amounts of hippuric acid.  相似文献   

16.
A simple chromatographic method to separate the arsenic metabolites in plasma and urine following iv administration of 74As to dogs is described. Plasma and urine samples were chromatographed on a cation-exchange resin column. Inorganic, monomethylated and dimethylated arsenic compounds were eluted with 0.5 n HCl, H2O, and 20% (vv) NH4OH solution respectively. Experiments on two dogs slow ~90% of the 74As in urine present in the form of dimethylarsinic acid 3 days after administration. It is also the predominant species in plasma 2 hr after dosing.  相似文献   

17.
The effects of two inclusion levels of benzoic acid (5 and 20 g/kg diet; B5 and B20, respectively) vs. a control (C) and an antibiotic (tiamulin) supplemented (A; 150 mg/kg) diet on nutrient digestibility and selected biochemical parameters were investigated in 48 weaned (n = 12 per treatment) rabbits (35 days old). Blood samples were obtained (at 45 and 85 days of age) to measure erythrocyte glutathione peroxidase (GSH-Px) activity, plasma alkaline phosphatase (ALP) activity and Ca and P levels. At 56 days of age total tract apparent digestibility (TTAD) was determined. At the end of the trial (85 days) rabbits were slaughtered for subsequent measurements. GSH-Px activity was also determined in liver homogenates post-mortem. Benzoic acid inclusion level decreased linearly (P<0.05) the weight of total full gastrointestinal tract, resulting in a trend (P=0.058) to increased dressing yield. The TTAD of organic matter (OM) increased linearly (P<0.05), whereas that of dry matter, crude protein and gross energy tended to increase (P=0.067, P=0.090 and P=0.098, respectively) with benzoic acid increment. The TTAD of aNDFom, hemicelluloses and cellulose was affected quadratically (P<0.05) by benzoic acid supplementation, with the B5-fed rabbits showing the higher values. Plasma ALP activity and P levels were linearly reduced (P<0.001) by benzoic acid inclusion, but no influence was observed on GSH-Px activity in erythrocytes or liver. In conclusion, dietary benzoic acid affects the TTAD of OM and fibre fractions in a dose dependent manner, without having any adverse effect on the systemic markers (biochemical parameters) examined.  相似文献   

18.
A method for the determination of unconjugated phentolamine at concentrations down to 6 ng/ml in human plasma, and of free and total (free plus conjugated) phentolamine down to 25 ng/ml in urine is described. After addition of 2-[N-(p-chlorophenyl)-N-(m-hydroxyphenyl)-aminomethyl]-2-imidazoline as internal standard, both compounds are extracted into benzene—ethyl acetate (1:1, v/v) at pH 10, transferred into an acidic aqueous solution and back-extracted at pH 10 into benzene—ethyl acetate. They are then derivatized with N-heptafluorobutyrylimidazole. The derivatives are determined by gas chromatography using a 63Ni electron-capture detector. In urine, total (free plus conjugated) phentolamine is determined after enzymatic hydrolysis. The technique was applied for the study of the plasma concentrations and urinary elimination after oral administration to man.  相似文献   

19.
The mutant mouse strain HPH2 (hyperphenylalaninemia) was isolated after N-ethyl-N-nitrosourea (ENU) mutagenesis on the basis of delayed plasma clearance of an injected load of phenylalanine. Animals homozygous for the recessive hph2 mutation excrete elevated concentrations of many of the neutral amino acids in the urine, while plasma concentrations of these amino acids are normal. In contrast, mutant homozygotes excrete normal levels of glucose and phosphorus. These data suggest an amino acid transport defect in the mutant, confirmed in a small reduction in normalized values of 14C-labeled glutamine uptake by kidney cortex brush border membrane vesicles (BBMV). The hyperaminoaciduria pattern is very similar to that of Hartnup Disorder, a human amino acid transport defect. A subset of Hartnup Disorder cases also show niacin deficiency symptoms, which are thought to be multifactorially determined. Similarly, the HPH2 mouse exhibits a niacin-reversible syndrome that is modified by diet and by genetic background. Thus, HPH2 provides a candidate mouse model for the study of Hartnup Disorder, an amino acid transport deficiency and a multifactorial disease in the human. Received: 16 May 1996 / Accepted: 25 September 1996  相似文献   

20.
The reference standard 2-fluoro-4-(1-(3,5,5,8,8-pentamethyl-5,6,7,8-tetrahydronaphthalen-2-yl)vinyl)benzoic acid was synthesized from 2,5-dimethyl-2,5-hexanediol and 2-fluoro-4-methylbenzoic acid in 10 steps with 3% overall chemical yield. The precursor 2-nitro-4-(1-(3,5,5,8,8-pentamethyl-5,6,7,8-tetrahydronaphthalen-2-yl)vinyl)benzoic acid was synthesized from 2,5-dimethyl-2,5-hexanediol and dimethyl-2-nitroterephthalate in seven steps with 2% overall chemical yield. The target tracer 2-[18F]fluoro-4-(1-(3,5,5,8,8-pentamethyl-5,6,7,8-tetrahydronaphthalen-2-yl)vinyl)benzoic acid was synthesized from its nitro-precursor by the nucleophilic substitution with K[18F]F/Kryptofix 2.2.2 and isolated by HPLC combined with solid-phase extraction (SPE) purification in 20–30% radiochemical yield with 37–370 GBq/μmol specific activity at end of bombardment (EOB).  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