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 46 毫秒
1.
Thalassemia is a genetic blood disorder requiring life‐long blood transfusions. This process often results in iron overload and can be treated by an iron‐chelating agent, like deferiprone (3‐hydroxy‐1,2‐dimethylpyridin‐4‐one), C7H9NO2, in an oral formulation. The first crystal structure of deferiprone, (Ia), was reported in 1988 [Nelson et al. (1988). Can. J. Chem. 66 , 123–131]. In the present study, two novel polymorphic forms, (Ib) and (Ic), of deferiprone were identified concomitantly with polymorph (Ia) during the crystallization experiments. Polymorph (Ia) was redetermined at low temperature for comparison of the structural features and lattice energy values with polymorphs (Ib) and (Ic). Polymorph (Ia) crystallized in the orthorhombic space group Pbca, whereas both polymorphs (Ib) and (Ic) crystallized in the monoclinic space group P21/c. The asymmetric units of (Ia) and (Ib) contain one deferiprone molecule, while polymorph (Ic) has three crystallographically independent molecules (A, B and C). All three polymorphs have similar hydrogen‐bonding features, such as an R22(10) dimer formed by O—H…O hydrogen bonds, an R43(20) tetramer formed by C—H…O hydrogen bonds and π–π interactions, but the polymorphs differ in their molecular arrangements in the solid state and are classified as packing polymorphs. O—H…O and C—H…O hydrogen bonds lead to the formation of two‐dimensional hydrogen‐bonded parallel sheets which are interlinked by π–π stacking interactions. In the three‐dimensional crystal packing, the deferiprone molecules were aggregated as corrugated sheets in polymorphs (Ia) and (Ic), whereas in polymorph (Ib), they were aggregated as a square‐grid network. The characteristic crystalline peaks of polymorphs (Ia), (Ib) and (Ic) were established through powder X‐ray diffraction analysis. The Rietveld analysis was also performed to estimate the contribution of the polymorphs to the bulk material.  相似文献   

2.
Crystal and molecular structure of the ethanol solvate of 2-[bis(diphenylphosphoryl)methyl]phenol, C31H26O3P2·3/4 C2H5OH (I), was determined by X-ray diffraction analysis. Crystals I are monoclinic: a=11.896(3), b=17.207(3), c=28.421(6) Å, β=90.75(2)°, Z=8, space group P21/c. The structure of I was solved by direct methods and refined anisotropically by large-block least-squares analysis to R=0.066 (CAD-4 automatic diffractometer, λMoKα radiation, 3650 independent reflections with I≥3σ). The crystal structure of I contains two independent molecules, Ia and Ib, of the basic substance and two independent solvate molecules of ethanol. The P atoms in molecules Ia and Ib have distorted tetrahedral environments; the average bond lengths are: P=O 1.476(3), P-C(sp3)=1.833(4) and P-C(Ar)=1.809(4) Å. Molecule Ia has an approximate symmetry plane Cs. The conformation of molecule Ib differs from that of Ia in rotation of one diphenylphosphoryl substituent around the P-C(sp3) bond by an angle of ≈34°. In molecule Ia, the O=P-C-P=O fragment has the form of a chela with two nearly parallel P=O bonds. In the crystal structure of I there are O-H...O=P intermolecular H-bonds; in molecules Ia and Ib there is an intramolecular H-bond of C-H...OH type.  相似文献   

3.
A rapid and sensitive liquid chromatography–tandem mass spectrometry (LC‐MS/MS) method was developed and validated for simultaneous quantification of escin Ia and escin Ib in human plasma. After a solid‐phase extraction (SPE), the analytes were separated on a Zorbax Extend C18 column by isocratic elution with a mobile phase of methanol–acetonitrile–10 mm ammonium acetate (27:27:46, v/v/v) at a flow rate of 1.0 mL/min and analyzed by mass spectrometry in the positive ion multiple reaction monitoring mode. The precursor to product ion transitions of m/z 1131.8 → 807.6 was used to quantify escin Ia and escin Ib. Good linearity was achieved over a wide range of 2.00–900 ng/mL for escin Ia and 1.50–662 ng/mL for escin Ib. The intra‐ and inter‐day precisions (as relative standard deviation) were less than 11% for each QC level of escin Ia and escin Ib. The accuracies (as relative error) were within ±5.27% for escin Ia and within ±4.07% for escin Ib. The method was successfully employed in a pharmacokinetic study after a single intravenous infusion administration of sodium aescinate injection containing 10 mg escin to each of the 10 healthy volunteers. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

4.
C7H7Mo(CO)(PN)I (I) (PN  (S)(+)-(C6H5)2PN(CH3)CH(CH3)(C6H5)) is prepared in 90% yield by reaction of C7H7Mo(CO)2I and PN. The two diastereo-isomers Ia and Ib differing only in the Mo-configuration exhibit chemical shift differences of their C7H7 and CH3 signals. Ia and Ib can be separated by fractional crystallization. In solution Ia epimerizes with respect to the Mo configuration. The half lives in benzene for the equilibration Ia ? Ib are 5.5, 30, and 104 min at 60, 50, and 40°C, respectively. Phosphine exchange experiments show that the epimerization proceeds via PN dissociation.An X-ray structure analysis was carried out on a single crystal of Ia. The absolute configuration at Mo was determined to be (R).  相似文献   

5.
The absorption spectra of Ru(2,2′-bipyridine)2 (2,2′-bipyridine-4,4′-dicarboxylic acid)2+ (I) and its diethyl ester (II) are closely related and are both significantly different from the spectra of the mono-protonated (Ia) and deprotonated (Ib) complexes. Luminescence polarization measurements show that for I and II the luminescent states have the transferred electron in the bipy-4,4′(COOH)2 and bipy-4,4′(COOEt)2 ligands, respectively, rather than in the unsubstituted bipy ligands.  相似文献   

6.
Diethyl 2-vinylcyclopropane-1,1-dicarboxylate (Ia), 2-vinylcyclopropane-1,1-dicarbonitrile (Ib), ethyl 1-cyano-2-vinylcyclopropanecarboxylate (Ic), and 1,1-diphenyl-2-vinylcyclopropane (Id) were radically homo- and copolymerized. Ia, Ib, and Ic polymerized cleanly in 1,5-type ring-opening fashion to yield polymers of high molecular weight. Id failed to homopolymerize but copolymerized with common monomers which included maleic anhydride. Infrared (IR) spectra indicated that the double bonds of the polymer chains were in trans form. All polymers were soluble in solvents and solution-cast films were clear and flexible, showing Tg values in the 25–40°C range. Model reactions of those monomers with benzenethiol also indicated predominant 1,5-addition reactions. From the results of our investigation it was concluded that the driving force for the facile radical 1,5-polymerization of Ia, Ib, and Ic was the stabilization of growing radicals by two substituents.  相似文献   

7.
2′, 3′-Di-O-protected guanosine derivatives (Ia and Ib) were allowed to react with phosphorus trichloride in acetone to give the N2-(1-methyl-1-phosphono)ethylguanosine derivatives (IIIa and IIIb).  相似文献   

8.
2-Phenyl-2-hydroxyethyl(aquo)cobaloxime(Ia) and the p-methyl (Ib) and p-cyano (Ic) derivatives have been synthesized and characterized. The parent compound and the p-methyl derivative decompose spontaneously both in the solid state and in methanol to give mixtures of the appropriate styrenes and acetophenones. Kinetics and product ratios of the decomposition of Ia and Ib in methanol have been studied as a function of temperature. Evidence is presented that both reaction pathways are ionic in nature and that an intermediate (presumed to be a phenethylcobaloxime carbonium ion) is involved in the styrene forming pathway. Acetophenones are apparently formed via a 1,2-hydride shift mechanism with solvent acting as a general base. Both reaction pathways show a large substituent effect with electron donating substituents increasing reactivity, Ic proved to be extraordinarily stable in methanol but decomposed readily in aqueous sulfuric acid to produce primarily p-cyanoacetophenone. Kinetic evidence for formation of a cationic intermediate is presented.  相似文献   

9.
Tri(meta-tolyl)antimony chloro(benzenesulfonate) (I) has been synthesized by the reaction between tri(meta-tolyl)antimony bis(benzenesulfonate) and tri(meta-tolyl)antimony dichloride (toluene; 0.5 h, 100°C). According to X-ray diffraction data, the antimony atoms in two crystallographically independent molecules Ia and Ib have a distorted trigonal bipyramidal coordination to chlorine and oxygen atoms in axial positions: OSbCl, 177.8(4)° (Ia), 178.6(4)° (Ib); CSbC, 114.0(9)°?127.6(8)° (Ia), CSbC, 114.4(8)°?128.0(8)° (Ib); Sb?C, 2.01(2)?2.139(18) Å, Sb?Cl, 2.453(7) Å, Sb?O, 2.163(13) Å (Ia), Sb?C, 2.07(2)?2.14(2) Å, Sb?Cl, 2.442(7) Å; Sb?O, 2.187(14) Å (Ib).  相似文献   

10.
Mass spectrometrical and 1H-NMR.-analyses of the exo-5-norbornen-2-yl acetate, formed by acetolysis of endo-5-norbornen-2-yl-2-exo-d brosylate, demonstrate that the deuterium initially on C(2) migrates partially (30%) onto C(1) (mechanism Ia or Ib). No deuterium could be detected on the other positions, which shows that C(1–7) migration is insignificant. 13C-NMR.-analysis of the deuteriated nortricyclyl acetate obtained as main product shows that the deuterium is equally and uniquely distributed between positions C(1) and C(6). This indicates that the nortricyclyl derivatives do not arise from nucleophilic attack on C(5) of asymmetrical norbornenyl intermediates, but from the reaction of a symmetrical nortricyclyl cation intermediate with solvent (mechanism Ib). Since the pioneering work of Roberts [1] and Winstein [2] on the solvolysis of exo- and endo-5-norbornen-2-yl derivatives 1-X and 2-X many papers have dealt with the cationic intermediate, the nature of which has still not been established satisfactorily [3]. We discuss briefly the main features of this homoallylic system and present experimental results that allow, for the first time, a clear distinction between five possible mechanisms Ia, Ib, II, III and IV of the degenerate rearrangement of the cationic intermediate formed in the acetolysis of the endo-5-norbornen-2-yl brosylate.  相似文献   

11.
The electron impact mass spectra of the new synthesized dioximes of o-diacyl benzenes (2) are reported. In addition to the molecular ion, characteristic peaks appear at values corresponding to the [M ? OH] +, [M ? NOH]+ and [M ? NHOH]+ ions. No initial dehydration of the molecular ion has been observed.  相似文献   

12.
3-Acetoxy-1-oxaphenalene4 (Ia) is conveniently prepared from naphthalene-1,8-carbalactone5 (IIa) by hydrolysis, condensation of the resulting hydroxy compound with chloroacetic acid and cyclisation of the resulting di-acid. Combined hydrolysis and reduction of (Ia) using alkaline potassium or sodium borohydride gives 2,3-dihydro-1-oxaphenalen-3-ol which may be converted into the 3-chloro-compound from which 1-oxaphenalene (Ib) is obtained by de-hydrochlorination.4  相似文献   

13.
The rates of the mononuclear heterocyclic rearrangement of the p-methoxyphenylhydrazone (Ib) and the m-nitrophenylhydrazone (1c) of 3-benzoyl-5-phenyl-1,2,4-oxadiazole have been measured in the range of pS+ 3.8–11.5 (solvent: dioxane/water 1:1, v:v) and compared with those of the unsubstituted phenylhydrazone (Ia). The obtained results show that in the base-catalyzed range, electron-repelling as well as electron-withdrawing substituents accelerate the rearrangement rates.  相似文献   

14.
1,1-Dichloro-2-vinylcyclopropane ( Ia ), 1,1-dichloro-2-methyl-2-vinylcyclopropane ( Ib ), 1,1,2-trichloro-2-vinylcyclopropane ( Ic ) were prepared from the corresponding dienes and chloroform in the presence of a phase transfer catalyst (PTC), R4N+Cl?. Monomers Ia – Ic underwent a clean 1,5-type radical ring-opening process to afford the corresponding polymers in good yield. Further, the relative rate of polymerization and reaction of ( I ) with thiophenol were studied.  相似文献   

15.
Authentic samples of 1,3,3-trimethyl-2-(3,4,5-trimethoxyphenyl)-4-methylenepiperidine (Ia) and 2-(p-chlorophenyl)-1,3,3-trimethyl-4-methylenepiperidine (Ib) are prepared by Mannich condensation between 4-methyl-1-methylamino-3-pentanone hydrochloride (VI) and an aromatic aldehyde, followed by a Wittig reaction on the resulting 4-piperidone. Comparing the physical and spectroscopic properties of Ia and Ib with those of the methylene derivatives IIa and IIb obtained as by-products in the Stevens rearrangement of 1-benzyl-1,3,4-trimethyl-1,2,5,6-tetrahydropyridinium salts IIIa and IIIb, respectively, it is shown that the assignment previously made for IIa and IIb is incorrect. Spectroscopic analysis (ir, 1H nmr, 13C nmr, ms) of these compounds and of its hydrogenation products VIII allows the structural and stereochemical assignment of 11a as cis-3-isopropenyl-1,3-dimethyl-2-(3,4,5-trimethoxyphenyl)pyrrolidine and of IIb as cis-2-(p-chlorophenyl)-3-isopropenyl-1,3-dimethylpyrrolidine. The formation of these rearrangement products is mechanistically interpreted as a Stevens [3,2] type process.  相似文献   

16.
The synthesis of 3-methyl-2-phenylbenzimidazole-1-oxide (VIII) has been accomplished. Consequently, it has been possible to establish the tautomeric structure of I as 1-hydroxy-2-phenylbenzimidazole (Ia). Contrary to recent observations on the parent 1-oxybenzimidazole (IV), I shows no tendency to exist as an N-oxide (Ib) in aqueous solution.  相似文献   

17.
Comparison of the NMR. spectra in CDCl3 of the heterocyclic bases obtained from the cyclisation of ω-(N-thiocarbamoylamino) ethyl (or propyl)-alcohols (or their orthophosphoric or sulfuric monoesters) to those of model compounds II (n = 1 or 2) and III (n = 1 or 2) has shown that: (1) In the case of five membred rings the C?N double bond is always endocyclic (Ib, n = 1) should R be aromatic, araliphatic or aliphatic; (2) In the case of six membered rings the C?S double bond is cnclocyclic when R is aliphatic or araliphatic (Ib, n = 2), and exocyclic when R is aromatic (I a, n = 2), with the exception of 2-(o-carboxyphcnylamino)-dihydro-δ2-m (Ib, n = 2, K = o-carboxyphcnyle). In CF, COOH, all five membered rings (I b, n = 1) show a triplet for the C-4 methylenic protons, whereas all the six membered rings (Ia or I b, n = 2) with the exception of I b, n = 2, R = o-carboxyphenyle, are represented b y a double triplet for the C-4 protons (samt. protonated spccics). Only one triplet is observed when the 3 position is substituted. Thiocarbamoylation of hydrazinoethanol or its orthophosphoric or sulfuric monoesters canoccur at either of the two nitrogen atoms, thus yielding upon cyclization five- (IT′) or six-membered rings (Va or Vb). The NMR spectra of compounds I V in (CIl,), SO show a singlet for 2 amino pro-tons (3-amino) and there is no further structural problem. The NMR spectra of compounds T′ in (CT), SO show a triplet for one amino proton coupling with the neighboring methylenic protons. I n this case, mode1 compounds are needed to assign the position of the C?N double bond ( e x cyclic V a or cndocyclic V b). When R = o-carboxyphenylc, the C?N double bond is probably endoc, yclic (Vb) because this ccimpound and 2-(o-carboxyphenvlarnino)-dihydro-δ2 have very similar UV spectra.  相似文献   

18.
Among the four 1,1-disubstituted 2-vinylcyclopropanes, diethyl 2-vinylcyclopropane-1,1-dicarboxylate (Ia), 2-vinylcyclopropane-1,1-dicarbonitrile (Ib), ethyl 1-cyano-2-vinylcyclopropanecarboxylate (Ic), and 1,1-diphenyl-2-vinylcyclopropane (Id), Ib and Ic polymerized well with sodium cyanide in N,N-dimethylformamide. Ib was most reactive and a polymer (IIb) from Ib exhibited an inherent viscosity of 1.05 dl/g (concentration of 1.0 g in 100 ml of 95% H2SO4). All experimental results indicated that the polymerization proceeded by ring opening and that the structure of the polymers had pendant vinyl groups. The polymer IIc from Ic was soluble in common solvents like acetone, but IIb was soluble only in 95% H2SO4. Reactions of those compounds with benzenethiolate ion in ethanol yielded addition products that supported the ring-opening polymerization of those monomers. In the postulated mechanism of polymerization cyanide ion attacks the carbon of a cyclopropane ring with electron-releasing vinyl group and the resulting anion is thereby stabilized by two electron-withdrawing substituents. The propagation takes place by the reaction of the anion with another monomer molecule.  相似文献   

19.
The ionization (or basicity) constants (pKb) were determined for many 2‐substituted 4,6‐diamino‐s‐tri‐azines ( I ) by means of the electrometric titration. I includes 2‐alkoxy or aryloxy‐( Ia ), 2‐alkyl‐ or 2‐aryl‐( Ib ), and 2‐alkylamino‐ or 2‐arylamino‐4,6‐diamino‐s‐triazines ( Ic ). For the series with the same alkyl or aryl group, the order of the basicity was found to be Ic < Ib < Ia . A study was made of relationships between the pKb, values of I , and the substituent constants, σp, σm, σp+, σm+, σpO, σmo, σI, σn, and σ*. The Hammett relationships were observed between the pKa values of I, and the substituent constants σm, (or the combination ones, [0.97σm + 0.03σp] as well as another [0.77σI + 0.23σR]). The Taft relationships were also found between the pKa values of Ia , Ib , and Ic and the constants σ*, respectively. Furthermore, in the case of Ic a linear relationship was observed between the pKa values and Σσ8.  相似文献   

20.
cis-2-Aryl-3-isopropenyl-1,3-dimethylpyrrolidines Ha and IIb have been synthesized by an unambiguous way, thus confirming the structure of the methylene derivatives obtained as by-products in the Stevens rearrangement of 1-benzyl-1,3,4-trimethyl-1,2,5,6-tetrahydropyridinium salts Ia and Ib. The synthesis is based on the acid-induced intramolecular cyclization between an iminium salt and the α-position of a ketal group. Thus, condensation between amino ketal XXI, prepared via Gabriel synthesis from 5-chloro-3-methyl-2-pentanone, and the appropriate aldehyde afforded imines XXI. Their treatment with dry hydrogen chloride followed by acid hydrolysis and methylation gave 3-acetylpyrrolidines IV, which were transformed into the isopropenyl derivatives II by reaction with methyl-lithium and further dehydration.  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