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.
Diacenaphtho[1,2-c:1,2-e]-1,2-dithiin 2 was synthesized in 23% yield by the reaction of acenaphthylene with elemental sulfur at 120 °C. This reaction also afforded either diacenaphtho[1,2-b:1,2-d]thiophene 1 or diacenaphtho[1,2-b:1,2-e]-dihydro[e]-1,4-dithiin 3 depending on the reaction time. Compound 2 was desulfurized and converted to 1 under UV-vis irradiation in a benzene solution. Reaction of 2 with Pt(COD)2 yielded the complex Pt(COD)(C24H12S2) 4 (COD=1,5-cyclooctadiene) by insertion of a Pt(COD) group into the S-S bond of 2. When heated, 4 was desulfurized and converted to 1 by elimination of a (COD)PtS grouping. Compounds 1-4 were characterized crystallographically.  相似文献   

2.
Three-component 1,2-carboamination of vinyl boronic esters with alkyl/aryl lithium reagents and N-chloro-carbamates/carboxamides is presented. Vinylboron ate complexes generated in situ from the boronic ester and an organo lithium reagent are shown to react with readily available N-chloro-carbamates/carboxamides to give valuable 1,2-aminoboronic esters. These cascades proceed in the absence of any catalyst upon simple visible light irradiation. Amidyl radicals add to the vinylboron ate complexes followed by oxidation and 1,2-alkyl/aryl migration from boron to carbon to give the corresponding carboamination products. These practical cascades show high functional group tolerance and accordingly exhibit broad substrate scope. Gram-scale reaction and diverse follow-up transformations convincingly demonstrate the synthetic potential of this method.

Three-component 1,2-carboamination of vinyl boronic esters with alkyl/aryl lithium reagents and N-chloro-carbamates/carboxamides is presented.

Alkenes are important and versatile building blocks in organic synthesis. 1,2-Difunctionalization of alkenes offers a highly valuable synthetic strategy to access 1,2-difunctionalized alkanes by sequentially forming two vicinal σ-bonds.1a–h Among these vicinal difunctionalizations, the 1,2-carboamination of alkenes, in which a C–N and a C–C bond are formed, provides an attractive route for the straightforward preparation of structurally diverse amine derivatives (Scheme 1a).2a–c Along these lines, transition-metal-catalyzed or radical 1,2-carboaminations of activated and unactivated alkenes have been reported.3a–p However, the 1,2-carboamination of vinylboron reagents, a privileged class of olefins,4a–h to form valuable 1,2-aminoboron compounds which can be readily used in diverse downstream functionalizations,5a–c,6a–d has been rarely investigated. To the best of our knowledge, there are only two reported examples, as shown in Schemes 1b and c. In 2013, Molander disclosed a Rh-catalyzed 1,2-aminoarylation of potassium vinyltrifluoroborate with benzhydroxamates via C–H activation (Scheme 1b).7 Thus, the 1,2-carboamination of vinylboron reagents is still underexplored but highly desirable.Open in a separate windowScheme 1Intermolecular 1,2-carboamination of alkenes.1,2-Alkyl/aryl migrations induced by β-addition to vinylboron ate complexes have been shown to be highly reliable for 1,2-difunctionalization of vinylboron reagents (Scheme 1c).4dh In 1967, Zweifel''s group developed 1,2-alkyl/aryl migrations of vinylboron ate complexes induced by an electrophilic halogenation.8 In 2016, the Morken group reported the electrophilic palladation-induced 1,2-alkyl/aryl migration of vinylboron ate complexes.9a–k Shortly thereafter, we,10a–c Aggarwal,11a–c and Renaud12 developed alkyl radical induced 1,2-alkyl/aryl migrations of vinylboron ate complexes. In these recent examples, the migration is induced by a C-based radical/electrophile, halogen and chalcogen electrophiles.13a,bIn contrast, N-reagent-induced migration of vinylboron ate complexes proceeding via β-amination is not well investigated. To our knowledge, as the only example the Aggarwal laboratory described the reaction of a vinylboron ate complex with an aryldiazonium salt as the electrophile, but the desired β-aminated rearrangement product was formed in only 9% NMR yield (Scheme 1c).13a No doubt, β-amino alkylboronic esters would be valuable intermediates in organic synthesis. Encouraged by our continuous work on amidyl radicals14a–i and 1,2-migrations of boron ate complexes,10a–c,15a–f we therefore decided to study the amidyl radical-induced carboamination of vinyl boronic esters for the preparation of 1,2-aminoboronic esters. N-chloroamides were chosen as N-radical precursors,16a–c as these N-chloro compounds can be easily prepared from the corresponding N–H analogues.17 Herein, we present a catalyst-free three-component 1,2-carboamination of vinyl boronic esters with N-chloroamides and readily available alkyl/aryl lithium reagents (Scheme 1d).We commenced our study by exploring the reaction of the vinylboron ate complex 2a with tert-butyl chloro(methyl)carbamate 3a applying photoredox catalysis. Complex 2a was generated in situ by addition of n-butyllithium to the boronic ester 1a in diethyl ether at 0 °C. After solvent removal, the photocatalyst fac-Ir(ppy)3 (1 mol%) and THF were added followed by the addition of 3a. Upon blue LED light irradiation, the mixture was stirred at room temperature for 16 hours. To our delight, the desired 1,2-aminoboronic ester 4a was obtained, albeit with low yield (26%,
EntryPhotocatalystSolventT (°C)Yieldb (%)
1 fac-Ir(ppy)3THFrt26
2 fac-Ir(ppy)3DMSOrt2
3 fac-Ir(ppy)3MeCNrt56
4Ru(bpy)3Cl2·6H2OMeCNrt69
5Na2Eosin YMeCNrt69
6cNa2Eosin YMeCNrt70
7cNoneMeCNrt45
8cNoneMeCN078
9cNoneMeCN−2088 (85)
10c,dNoneMeCN−202
Open in a separate windowaReaction conditions: 1a (0.20 mmol), nBuLi (0.22 mmol), in Et2O (2 mL), 0 °C to rt, 1 h, under Ar. After vinylboron ate complex formation, solvent exchange to the selected solvent (2 mL) was performed.bGC yield using n-C14H30 as an internal standard; yield of isolated product is given in parentheses.c4 mL MeCN was used.dReaction carried out in the dark.With optimal conditions in hand, we then investigated the scope of this new 1,2-carboamination protocol keeping 2a as the N-radical acceptor (Scheme 2). This transformation turned out to be compatible with various primary amine reaction partners bearing carbamate (4a, 4b and 4d–4g) or acyl protecting groups (4c) (20–85%). Notably, N-chlorolactams can be used as N-radical precursors, as shown by the successful preparation of 4h (71%). Moreover, Boc-protected ammonia was also tolerated, delivering 4i in an acceptable yield (55%).Open in a separate windowScheme 21,2-Carboamination of 1a with various amidyl radical precursors. Reaction conditions: 1a (0.20 mmol, 1.0 equiv.), nBuLi (0.22 mmol, 1.1 equiv.), in Et2O (2 mL), 0 °C to rt, 1 h, under Ar; then [N]-Cl (0.24 mmol, 1.2 equiv.), −20 °C, 16 h, in MeCN (4 mL). Yields given correspond to yields of isolated products. aA solution of [N]-Cl (0.30 mmol, 1.5 equiv.) in MeCN (1 mL) was used. See the ESI for experimental details.We continued the studies by testing a range of vinylboron ate complexes (Scheme 3). To this end, various vinylboron ate complexes were generated by reacting the vinyl boronic ester 1a with methyllithium, n-hexyllithium, isopropyllithium and tert-butyllithium. For the n-alkyl-substituted vinylboron ate complexes, the 1,2-carboamination worked smoothly to afford 4j and 4k in good yields. However, the vinylboron ate complex derived from isopropyllithium addition provided the desired products in much lower yield (4l, 18% yield). When tert-butyllithium was employed, only a trace of the targeted product was detected (see ESI). As expected, cascades comprising a 1,2-aryl migration from boron to carbon worked well. Thus, by using PhLi for vinylboron ate complex formation, the 1,2-aminoboronic esters 4m–4o were obtained in 69–73% yields with the Boc (t-BuOCONClMe), ethoxycarbonyl-(EtOCONClMe) and methoxycarbonyl (Moc)-(MeOCONClMe) protected N-chloromethylamines (for the structures of 3, see ESI) as radical amination reagents. Keeping 3b as the N-donor, other aryllithiums bearing various functional groups at the para position of the aryl moiety, such as methoxy (4p), trimethylsilyl (4q), methyl (4r), phenyl (4s), trifluoromethoxy (4t), trifluoromethyl (4u), and halides (4v–4x) all reacted well in this transformation. Aryl groups bearing meta substituents are also tolerated, as documented by the preparation of 4y (81%). To our delight, a boron ate complex generated with a 3-pyridyl lithium reagent engaged in the cascade and the carboamination product 4z was isolated in high yield (82%).Open in a separate windowScheme 3Scope of vinylboron ate complexes. Reaction conditions: 1 (0.20 mmol, 1.0 equiv.), RMLi (0.22 mmol, 1.1 or 1.3 equiv.), Et2O or THF, under Ar; then [N]-Cl (0.30 mmol, 1.5 equiv.), −20 °C, 16 h, in MeCN. Yields given correspond to yields for isolated products. See the ESI for experimental details.The reason for the dramatic reduction in yield when α-branched alkyllithium or electron-rich aryllithium reagents were used might be that the corresponding vinylboron ate complexes could be oxidized by N-chloroamides via a single-electron oxidation process.18a–e Furthermore, the α-unsubstituted vinyl boronic ester and vinyl boronic ester bearing various α-substituents are suitable N-radical acceptors and the corresponding products 4aa–4ac were obtained in 48–70% yield.To gain insights into the mechanism of this 1,2-carboamination, a control experiment was conducted. The reaction could be nearly fully suppressed when the reaction was carried out in the presence of a typical radical scavenger (2,2-6,6-tetramethyl piperidine-N-oxyl, TEMPO), indicating a radical mechanism (Scheme 4a). Further, considering an ionic process, the N-chloroamides would react as Cl+-donors that would lead to Zweifel-type products, which were not observed under the applied conditions. The proposed mechanism is shown in Scheme 4b. As chloroamides have been recently proposed to undergo homolysis under visible light irradiation,19a,b we propose that initiation proceeds via homolytic N–Cl cleavage generating the electrophilic amidyl radical A, which then adds to the electron-rich vinylboron ate complex 2a to give the adduct boronate radical B. The radical anion B then undergoes single electron transfer (SET) oxidation with 3a in an electron-catalyzed process20a,b or chloride atom transfer with 3a to provide C or D along with the amidyl radical A, thereby sustaining the radical chain. Intermediates C or D can then react via a boronate 1,2-migration10c,11c,21 to eventually give the isolated product 4a.Open in a separate windowScheme 4Control experiment and proposed mechanism.To document the synthetic utility of the method, a larger-scale reaction and various follow-up transformations were conducted. Gram-scale reaction of 2a with 3a afforded the desired product 4a in good yield, demonstrating the practicality of this transformation (Scheme 5a). Oxidation of 4a with NaBO3 provided the β-amino alcohol 5 in 89% yield (Scheme 5b). The N-Boc homoallylic amine 6 was obtained by Zweifel-olefination with a commercially available vinyl Grignard reagent and elemental iodine in good yield (79%).22 Heteroarylation of the C–B bond in 4a was realized by oxidative coupling of 4a with 2-thienyl lithium to provide 7.23Open in a separate windowScheme 5Gram-scale reaction and follow-up chemistry.In summary, we have described an efficient method for the preparation of 1,2-aminoboronic esters from vinyl boronic esters via catalyst-free three-component radical 1,2-carboamination. Readily available N-chloro-carbamates/carboxamides, which are used as the N-radical precursors, react efficiently with in situ generated vinylboron ate complexes to afford the corresponding valuable 1,2-aminoboronic esters in good yields. The reaction features broad substrate scope and high functional group tolerance. The value of the introduced method was documented by Gram-scale reaction and successful follow-up transformations.  相似文献   

3.
The vibrational analysis and torsional study of trans-1,2-dimethylcyclopropane and cis-1,2-dimethylcyclopropane     
J.R. Durig  A.B. Nease  F. Milani-Nejad 《Journal of Molecular Structure》1981
The infrared spectra of cis-1,2-dimethylcyclopropane and trans-1,2-dimethylcyclopropane have been recorded between 4000 and 200 cm?1 in the polycrystalline solid phase, and 4000 to 80 cm?1 in the gas phase. The Raman spectra of these two compounds in the gaseous and liquid phases were also recorded between 3100 and 10 cm?1. An assignment of the thirty-nine fundamental vibrations for both cis- and trans-1,2-dimethylcyclopropane is proposed, and comparisons are made with the vibrations of other similar molecules. Additionally, ten torsional transitions were observed in the far infrared and Raman spectra of cis-1,2-dimethylcyclopropane, and four transitions were observed in the spectra of the trans compound. From these spectral data, torsional barriers were determined. The effective barriers to methyl torsion are 2.92 kcal mol?1 (12.20 kJ mol?1) for cis-1,2-dimethylcyclopropane and 2.61 kcal mol?1 (11.14 kJ mol?1) for trans-1,2-dimethylcyclopronane.  相似文献   

4.
Hydroborierung von 3-methyl-1,3-butadien mit 1,2:1,2-bis(tetramethylen)diboran(6)     
Rosalinda Contreras  Bernd Wrackmeyer 《Journal of organometallic chemistry》1981,205(1):15-19
The nature of the hydroboration product obtained from 3-methyl-1,3-butadiene and 1,2 : 1,2-bis(tetramethylene)diborane(6) (1) allows for the discussion of the reaction mechanism. The hydroboration proceedsby retention of the cyclic structure in the first step, followed by exchange of BH and BC bonds and final cyclic hydroboration to give 1-(boracyclopentyl)-4-(3-methylboracyclopentyl)butane (5). The structural assignment of 5 is based on 1H, 11B and 13C NMR data.  相似文献   

5.
1,2-oxazine chemistry—IV : The conformational equilibria in 2,5- and 2,4-dimethyltetrahydro-1,2-oxazines     
F.G. Riddell 《Tetrahedron》1975,31(6):523-525
The synthesis of 2,5-dimethyltetrahydro-1,2-oxazine via 5-methyldihydro-1,2-oxazine and the preparation of a pure sample of 2,4-dimethyltetrahydro-1,2-oxazine are reported. For the 2,5-dimethyl derivative low temperature (?40° to ?45°) 1H NMR measurements show signals from the trans (95%) and cis (5%) conformations. From this result it follows that an axial 5-Me group in a tetrahydro-1,2-oxazine ring is 5·7 ± 0·4 kJ mole?1 less stable than when equatorial. Low temperature measurements on the 2,4-dimethyl derivative fail to show any sign of the conformation with an axial Me group. These results in conjunction with earlier relative free energy difference measurements, give the following conformational free energy differences for Me groups on ring C atoms; C(4) 7·1 ± 1·0; C(3) 7·9 ± 0·8; C(6) 10·1 ± 1·6 kJ mole?1.  相似文献   

6.
Conformational study of 1,2-dithiacyclononane     
《Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy》1989,45(3):321-327
The temperature dependence of the Raman spectrum of 1,2-dithiacyclononane (1,2-DTCN) in the SS stretching region has been used to infer the existence of a conformational equilibrium with ΔH0 = 5.0 ± 0.8 kJ/mol. Molecular mechanics calculations predict a (2 2 5)-C2 lowest energy conformation in equilibrium with a (2 3 4) structure. The fully decoupled 13C NMR spectrum at −80°C and the Raman spectra are consistent with this postulate. The temperature dependence of the 1H NMR spectrum of 1,2-DTCN is characteristic of the ring inversion process. A crude lineshape analysis allows us to calculate ΔG0 = 49.0 ± 1.2 kJ/mol.  相似文献   

7.
Synthesis and physicochemical studies on 1,2-bisazolylethanes     
J. Torres  J. L. Lavandera  P. Cabildo  R. M. Claramunt  J. Elguero 《Journal of heterocyclic chemistry》1988,25(3):771-782
Twenty symmetrical and asymmetrical 1,2-bisazolylethanes have been obtained from azoles and 1,2-dibromoethane or 1-chloro-2-(pyrazol-1-yl)ethane by phase transfer catalysis (PTC). The 1H and 13C nmr properties are reported and the chemical shifts of the ethylene carbon atoms discussed using an additive model.  相似文献   

8.
Herstellung von Argininpeptiden mitN 7,N 8-(1,2-Dihydroxycyclohex-1,2-ylen)-Schutz     
Ulrich Hevelke  Josef Föhles  Jean Knott  Helmut Zahn 《Monatshefte für Chemie / Chemical Monthly》1982,113(4):457-473
A method for the synthesis of arginine peptides is described, in which the side chain guanidine function is blocked through reaction with 1,2-cyclohexanedione in borate buffers.Coupling to the carboxyl group of arginine was achieved by active ester, by dicyclohexylcarbodiimide/1-hydroxybenzotriazole1, and by the mixed anhydride methods2. Neither lactam formation nor acylation of the vicinal hydroxyls of the N7, N8-(1,2-dihydroxycyclohex-1,2-ylene) guanidino group was observed.Removal of the protecting group is strongly influenced by steric factors. Some side reactions observed during modification of peptides and protein fragments with 1,2-cyclohexanedione are also described.
Herrn Professor Dr.Hermann Stetter zum 65. Geburtstag gewidmet.  相似文献   

9.
Kinetics of the pyrolysis of 1,2-diiodoethylene     
Shozo Furuyama  David M. Golden  Sidney W. Benson 《国际化学动力学杂志》1969,1(2):147-156
The spectrophotometric determination of the rate of pyrolysis of 1,2-diiodoethylene from 305.8 to 435.0° (with additional data on the addition of iodine to acetylene from 198.1 to 331.6°) has resulted in the observation of both a (in part heterogeneous) unimolecular process (A), and an iodine atom catalyzed process (B). For the homogeneous unimolecular process, log (kA/sec?1) ≈ 12.5–46/θ would appear to be reasonable, while log (kB/M?1 sec?1) = 11.8–23.9/θ, where θ = 2.303RT in kcal/mole. It is suggested that a donor–acceptor complex intermediate may explain the observed rate constant of process B and analogous reactions in other systems.  相似文献   

10.
Metallkomplexe eines 1-Oxo-1,2-diphosphacyclopentanons     
G. Bergerhoff  R. Weber  M. Cieslik  M. Fingerhuth  J. Lindner  H. J. Padberg 《无机化学与普通化学杂志》1999,625(10):1592-1598
Coordination Compounds with an 1-Oxo-1,2-diphosphacyclopentanone 1-Oxo-1,2-diphosphacyclopentanone ( 2 ) is a ligand with three different sites for typical coordination. Thus seven types of structural different coordination compounds as well as salts can be received: 2 -W(CO)5; ( 2 -SbCl3)2; 2 (–)-Cu(2+)- 2 (–); 2 (–)-Co(3+)(H2O) · (OH)- 2 (–); (CO)5W- 2 (–)-Cu(2+)- 2 (–)-W(CO)5; (CO)5W- 2 -SbCl3; Cl3Sb- 2 (–)-Cu(2+)- 2 (–)-SbCl3; Cl3Sb((CO)5W)- 2 (–)-Cu(2+)-2(–)-(W(CO)5)SbCl3; [Cu(en)2(H2O)2]2+ ( 2 (–)); [Cu(en)2(CH3OH)2]2+ [ 2 (–)-W(CO)5]2; NR4(+) 2 (–)  相似文献   

11.
Synthesis of 1H-imidazo[1,2-a]imidazole     
Frans Compernolle  Suzanne Toppet 《Journal of heterocyclic chemistry》1986,23(2):541-544
The synthesis and structure analysis of the unknown 1H-imidazo[1,2-a]imidazole ( 1 ) is described. The preparation involves alkylation of 2-aminoimidazole with bromoacetaldehyde diethyl acetal and subsequent hydrolysis and cyclization with hydrochloric acid. The structure was characterized by mass spectrometry and by 1H-, 15N- and 13C-nmr of 1 and by 1H-nmr of its 1-benzyl derivative 8 . An independent synthesis of 8 was accomplished via cyclization of 2-(N-dichloroethyl-N-benzyl)aminoimidazole ( 11 ).  相似文献   

12.
A Reductive 1,2 Transposition of Acyclic Ketones     
Gerald L. Larson  Lelia M. Fuentes 《合成通讯》2013,43(9):841-844
Earlier we noted that the hydroboration of the trimethylsilyl enol ether of an acyclic ketone results in an elimination of a trimethylsiloxyborane moiety with the subsequent formation of an olefin.1,2 The olefin formed then undergoes hydroboration giving a monoalcohol upon oxidation. (eq 1) We wish to report here on the utility of this sequence, illustrated in eq 1, in the reductive 1,2 transposition of acyclic ketones.3  相似文献   

13.
Quantum chemical approach to isomerization of 1,2-dihydro-1,2-epoxybenzene and its derivatives involving hydration     
《Journal of Molecular Structure》1998,422(1-3):1-12
The present study compares experimental values of log(104 × k0) or log kH in the isomerization of 1,2-dihydro-1,2-epoxybenzene and its derivatives with the activation parameters by the PM3 calculation in the gas phase and a simple model of hydration which includes only water molecules linking directly to the solute. The calculated results show that none of the activation parameters in the gas phase can explain the experimental rate constants, while the enthalpies in the simple model of hydration are capable of adequately reproduce their differences. Although the hydration to the lone pairs of oxygens (21.2–22.3 kcal mol−1) comprises the main part of the total hydration effect (34.1–42.8 kcal mol−1) and the weak hydration to the apolar parts comprises smaller part, the latter determines the relative rates of the isomerization.  相似文献   

14.
Bis(1,2‐diselenoquadratato)metallate     
U. Drutkowski  B. Wenzel  D. Tews  P. Strauch 《无机化学与普通化学杂志》2001,627(8):1888-1894
Bis(1,2‐diselenosquarato) Metalates A series of 1,2‐diselenosquarato metalates [M(dssq)2]2– (M = Pd2+, Pt2+, Cu2+, Ni2+, Zn2+, Cd2+, Pb2+, VO2+) was available by direct synthesis from the appropriate metal salt with dipotassium 1,2‐diselenosquarate in deoxygenized water under an argon athmosphere. The copper(II)complex, [Cu(dssq)2]2–, and the oxovanadium(IV)complex, [VO(dssq)2]2–, were identified in solution by EPR spectroscopy (parameters: [Cu(dssq)2]2–: g0 = 2.073; a = –76.0 · 10–4 cm–1, a = 47.0 · 10–4 cm–1; [VO(dssq)2]2–: g0 = 1.986; a = 74.9 · 10–4 cm–1). The complexes bis(tetraphenylphosphonium)[bis(1,2‐diselenosquarato)nickelate(II)], (Ph4P)2[Ni(dssq)2], and bis(tetraphenylphosphonium)[bis(1,2‐diselenosquarato)zincate(II)], (Ph4P)2[Zn(dssq)2], were characterized by X‐ray structure analysis. The square‐planar NiII complex (Ph4P)2[Ni(dssq)2] crystallizes in the monoclinic spacegroup P21/n with the unit cell parameters a = 11.1472(8) Å, b = 15.331(1) Å, c = 14.783(1) Å, β = 94.441(1)° and Z = 2. The ZnII‐complex (Ph4P)2[Zn(dssq)2] is tetrahedral coordinated and crystallizes in the monoclinic spacegroup P21/c with the unit cell parameters a = 9.4238(1) Å, b = 18.5823(3) Å, c = 29.5309(5) Å, β = 96.763(1)° and Z = 4.  相似文献   

15.
1,2-Bis-[(5-methyl/chloro/nitro)-2-1H-benzimidazolyl]-1,2-ethanediols and their PdCl2 complexes     
Ülküseven  Bahri  Tavman  Aydın 《Transition Metal Chemistry》2001,26(6):723-726
1,2-Bis-[(5-methyl)-2-1H-benzimidazolyl]- (L 1), 1,2-bis-[(5-chloro)-2-1H-benzimidazolyl]- (L 2), 1,2-bis-[(5-nitro)-2-1H-benzimidazolyl]-1,2-ethanediol (L 3) and their PdCl2 complexes were synthesized and characterized by elemental analysis, molar conductivity, i.r. and 1H-n.m.r. spectra. The benzene ring substituents lead to a decrease in melting point. The methyl group reduces the solubility and the acidity of L 1 and Pd(L 1)Cl2, whereas the Cl and NO2 groups increase the solubility and the acidity of L 2, L 3, Pd(L 2)Cl2 and Pd(L 3)Cl2. In Pd(L 1)Cl2 and Pd(L 2)Cl2 complexes, the ligands act as a bidentate through two nitrogen atoms. In Pd(L 3)Cl2, ligand coordination occurs through one OH group oxygen atom and one of the benzimidazole nitrogen atoms.  相似文献   

16.
Penetration of sodium cetylsulfate into monolayers of 1,2-dipalmitoyl- and 1,2-dimyristoyl-phosphatidylethanolamine     
H. -D. Dörfler  W. Rettig  H. Sackmann 《Colloid and polymer science》1985,263(7):563-569
The penetration of sodium cetylsulfate into monolayers of dipalmitoyl- and dimyristoyl-phosphatidylethanolamine was studied by the measurement of surface and penetration pressures using the vertical plate method of Wilhelmy. The penetration isotherms in two systems were investigated at different initial molecular areasA M :System I: Sodium cetylsulfate/1,2-dipalmitoyl-phosphatidylethanolamine atA M = 0.85; 0.75; 0.65; 0.55; 0.50; 0.46 and 0.44 nm2 · molecule–1.System II: Sodium cetylsulfate/1,2-dimyristoyl-phosphatidylethanolamine atA M = 0.85; 0.75; 0.60 and 0.55 nm2 · molecule–1.(T=295 K; substrate 0.1 M NaCl)The penetration isotherms (F t vs. logc s ) increase linearly atF t > 10 mN · m–1 in system I and atF t >25 mN · m–1 in system II. The isotherms of both systems are shifted to lower surfactant concentrations with decreasing molecular area of spread monolayer. A maximum of the slopes (dFt/d logc s )occurs at AM=0.50 nm2 · molecule–1. This behavior is also reflected in the dependenceG p 0 (free standard penetration enthalpy) and s (relative surface excess concentration of surfactant) onA M . These changes are related to a different packing of the compounds in the binary penetrated monolayers.In the high pressure region both system are nearly identical. Differences in the low pressure region arise from the penetration into different monolayer states.Nomenclature M effective cross sectional area of monolayer molecule - a M partial molecular area of monolayer molecule - a s partial molecular area of surfactant molecule in the penetrated film - a s 0 molecular area of surfactant molecule at definite film pressure (eq. (3)) - A M molecular area of theF/A-isotherm - A N constant in equation (2) - A K collapse area - b penetration coefficient in equations (2); (2 a) - c s bulk concentration of surfactant - logc s relative shift of penetration isotherm with regard to the adsorption isotherm at constantF t - F film pressure of monolayer component in absence of surfactant - F t total film pressure - F p film pressure change due to penetration - F p,max constant in equation (1) - G p 0 free standard penetration energy - k Boltzmann constant - K constant in equation (1) - R gas constant - T temperature - x s 0 mole fraction of surfactant in the penetrated film - M surface concentration of monolayer molecules - s relative adsorption of surfactant - w 0 surface concentration of surfactant in monolayer-free surface - factor in equation (6 a) - surface tension  相似文献   

17.
UV spectroscopy of monosubstituted derivatives of 1,2-dihydro-C60-fullerenes     
Yu. N. Biglova  V. A. Kraikin  S. A. Torosyan  V. V. Mikheev  S. V. Kolesov  A. G. Mustafin  M. S. Miftakhov 《Journal of Structural Chemistry》2012,53(6):1081-1086
UV spectroscopy is used to determine the molar absorption coefficients of C60 fullerene and monosubstituted 1,2-dihydro-C60-fullerenes in different solvents. It is found that the extinction coefficient of C60 at 330 nm (the main absorption band most frequently used for qualitative and quantitative determination of the C60 content) is independent of the nature of the solvent and is ~54400 M?1·cm?1. The molar absorption coefficients of a series of monosubstituted 1,2-dihydro-C60-fullerenes are practically independent of the chemical structure and the length of the substituent and are 35700 M?1·cm?1 (λ ~ 328 nm) and 115250 M?1·cm?1 (λ ~ 257 nm). It is shown that the substitution in fullerene proceeds via the double 6,6 bond, as evidenced by the absorption band at 424 nm in the spectra of these compounds, which is characteristic of monosubstituted methanofullerenes.  相似文献   

18.
1,2-Dipyridazinyl-?then- und-?than-1,2-diole aus Pyridazin-carbaldehyden     
Gottfried Heinisch  Ernst Luszczak  Andreas Mayrhofer 《Monatshefte für Chemie / Chemical Monthly》1976,13(2):799-808
Treatment of pyridazine-4-carboxaldehyde with catalytic amounts of KCN results in formation of pyridazine-4-carboxylic acid and two stereoisomeric diols (4,5). The configurations of4 and5 are explained on basis of the1H-NMR-spectra, the mechanism of reaction is discussed. Pyridazine-3-carboxaldehyde (10) reacts with HCN to form the addition product of10-cyanhydrine to10 (11). HCN-elimination from11 yields the enediol12 which by oxygen is oxidized to pyridazine-3-carboxylic acid or its methyl ester.  相似文献   

19.
β-Ribo- and α-arabinonucleosides containing the 1,2-benzisoxazole and 1,2-benzisothiazole rings     
Kiyotaka Yoshii  Yoshihiro Ohba  Tarozaemon Nishiwaki 《Journal of heterocyclic chemistry》1993,30(1):141-144
The reaction of the silylated base of 1,2-benzisoxazol-3(2H)-one ( 1 ) and its 7-methyl derivative 5 and 5-methyl-1,2-benzisothiazol-3(2H)-one ( 9 ), respectively, with 1-O-acetyl-2,3,5-tri-O-benzoyl-β-D-ribofuranose followed by basic deprotection gave the corresponding β-D-ribonucleosides, and the silylated base of 1 , when treated with 1-O-acetyl-2,3,5-tri-O-benzoyl-α-D-arabinofuranose in the presence of stannic chloride, afforded the corresponding α-arabinonucleoside. Structural proofs of these nucleosides are provided from elemental analyses and 1H and 13C nmr spectra.  相似文献   

20.
Molar heat capacity of binary liquid mixtures: 1,2-dichloroethane + cyclohexane and 1,2-dichloroethane + methylcyclohexane     
Emmerich Wilhelm  J.-P.E. Grolier  M.H.Karbalai Ghassemi 《Thermochimica Acta》1979,28(1):59-69
The molar heat capacity at constant pressure, CP, of the two binary liquid mixtures 1,2-dichloroethane + cyclohexane and 1,2-dichloroethane + methylcyclohexane were determined at 298.15 K from measurements of the volumetric heat capacity, CP/V, in a Picker flow microcalorimeter (V is the molar volume). For the molar excess heat capacity, CPE, the imprecision of the adopted stepwise procedure is characterized by a standard deviation of about ± 0.05 J K?1 mole?1, which amounts to ca. 3% of CPE. Literature data on ultrasonic velocities, on molar volumes, and on coefficients of thermal expansion were used to calculate the molar heat capacity at constant volume, Cv, and the isothermal compressibility, βT, of the pure substances, as well as the corresponding excess quantities CVE and (VβT)E of the binary mixture 1,2-dichloroethane + cyclohexane. A preliminary discussion of our results in terms of external and internal rotational behavior (trans-gauche equilibrium of 1,2-dichloroethane) is presented.  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