首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The electronic and geometric structures and the energetic characteristics of isolated dimeric molecules and positive ions of ferriporphyrin analogues (MC34H31O4N4)20,+ and corresponding monomers MC34H31O4N40,+ with 3d-metal atoms M = Sc-Ni in states with different multiplicities were calculated by the density functional theory B3LYP method with the Gen-1 = 6-31G*(Fe) + 6-31G(C,H,N,O) basis set. Their energies were refined with the use of the extended basis set Gen-2 = 6-311+G*(Fe) + 6-31G*(C,H,N,O). The computation results were compared with the analogous calculated data on the dimers of heme analogues (MC34H32O4N4)20,+ and MC34H32O4N40,+ with the same M atoms. The behavior of the above-mentioned properties was analyzed in going along the 3d series, upon ionization and a change in multiplicity, and in other related series. Trends were traced in the calculated energies D 1 of dissociation of the ferriporphyrin analogue dimers into monomers, (MC34H31O4N4)20,+ → MC34H31O4N40,+MC34H31O4N40, and in the energies D 2 of removal of two hydrogen atoms from heme analogue dimers resulting in ferriporphyrin analogues, (MC34H32O4N4)20,+ → (MC34H31O4N4)20,+ + 2H. It was shown that, in going along the 3d series, the D 1 energy rapidly decreases from ∼210 kcal/mol for M = Sc to a few tens of kcal/mol for M = Ni, with a small minimum for Mn and a small maximum for Fe, as the energy of the pair of broken covalent (polar) bonds M-O decreases. Conversely, the D 2 energy increases rapidly and rather monotonically from ∼45 kcal/mol for M = Sc to ∼135 kcal/mol for M = Fe, as the difference between the energies of the pair of forming covalent bonds M-O and the pair of broken donor-acceptor bonds M ← O decreases. For the positive ions, the D 1 and D 2 values are, as a rule, 5–10 kcal/mol higher than those for the parent neutral analogues. The character of the spin density distribution in states with different multiplicities was considered. Original Russian Text ? O.P. Charkin, 2009, published in Zhurnal Neorganicheskoi Khimii, 2009, Vol. 54, No. 4, pp. 655–665.  相似文献   

2.
Concerning the preparation of methyl α-D -fructopyranoside and the structure of the orthoester by-products Koenigs-Knorr glycosidation (AgClO4/Ag2CO3) of the easily accessible p-nitrobenzoylated fructosylchloride 7 yields mainly the protected glycoside 3 which was deacylated to the known title compound 1 . The orthoesters 15 and 16 were formed as by-products in this glycosidation whilst the analogous orthoesters 21 and 22 are formed as main products in the glycosidation (Ag2CO3) of the benzoylated fructosylchloride 12 . The structure of these orthoesters was deduced mainly from their spectroscopic data, from those of their derivatives 17 , 18 , 19 , 20 , 25 and 26 and by comparison of the latter with the model compounds 31 and 34 .  相似文献   

3.
Starting from D-glutamic acid ( 5 ), the bicyclic compounds 4a and 4b were synthesized via 17 (Schemes 1 and 2). The reaction leading to 4g and 4h with LiCuPh2 was not successful. But treatment of the N-protected model lactams 19 , 21 , and 22 with Li2Cu(CN)Ph2 gave the amino ketones 24 , 26 , and 27 , respectively (Scheme 3). The desired compound 23 was obtained from 20. Conversion of the unprotected lactams 28 , 31 , and 32 gave the phenyl derivative 34 in excellent yields. Ester 35 was transformed to the α -amino-γ- oxo-acid derivative 36. This conversion opens a novel access to this type of compounds.  相似文献   

4.
The pyrrolidine derivatives 3 , 4 , and 5 were prepared from the methyl ester 7 of Neu2en5Ac via lie pyrrolidine-borane adduct 33 . They inhibit Vibrio cholerae sialidase competitively with Ki = 4. 4 10?3 M, 5. 3 10?3 M, and 4. 0 10?2 M, respectively. Benzylation of 7 gave the fully O-benzylated 8 besides 9, 10 , and 11. Ozonolysis and reduction with NaBH4 of 8 and 9 gave the 1, 4-diols 12 and 15 , the hydroxy acetates 13 and 16 , and the furanoses 14 and 17 (Scheme 1), respectively. The diol 12 was selectively protected (→ 19 → 20 → 23 ) and transformed into the azide 27 by a Mitsunobu reaction. Selective base-catalysed deprotection of the diacetate 22 , obtained from 12 , was hampered by an easy acetyl-group migration. The mesylate 28 proved unstable. The azide 27 was transformed via 29 into the ketone 30 (Scheme 2). Hydrogenation of 30 gave the dihydropyrrole 31 and, hence, the pyrrole 32. The adduct 33 was obtained from 30 by a Staudinger reaction (→31) and reduction with LiBH4/HBF4. It was transformed into the pyrroudine 34 . The structure of 34 was established by X-ray analysis. Reductamination of the pyrrolidine-borane adduct with glyoxylic acid gave 40 and, hence, 3. N-Alkylation afforded 44 and, hence, the phosphonate 4. The acid 5 was obtained from 33 by acylation (→ 47 ) and deprotection (Scheme 4).  相似文献   

5.
The protected disaccharide 44 , a precursor for the synthesis of allosamidin, was prepared from the glycosyl acceptor 8 and the donors 26–28 , best yields being obtained with the trichloroacetimidate 28 (Scheme 6). Glycosidation of 8 or of 32 by the triacetylated, less reactive donors 38–40 gave the disaccharides 46 and 45 , respectively, in lower yields (Scheme 7). Regioselective glycosidation of the diol 35 by the donors 38–40 gave 42 , the axial, intramolecularly H-bonded OH? C(3) group reacting exclusively (Scheme 5). The glycosyl acceptor 8 was prepared from 9 by reductive opening of the dioxolane ring (Scheme 3). The donors 26–28 were prepared from the same precursor 9 via the hemiacetal 25 . To obtain 9 , the known 10 was de-N-acetylated (→ 18 ), treated with phthalic anhydride (→ 19 ), and benzylated, leading to 9 and 23 (Schemes 2 and 3). Saponification of 23 , followed by acetylation also gave 9 . Depending upon the conditions, acetylation of 19 yielded a mixture of 20 and 21 or exclusively 20 . Deacetylation of 20 led to the hydroxyphthalamide 22 . De-N-acetylation of the 3-O-benzylated β-D -glycosides 11 and 15 , which were both obtained from 10 , was very sluggish and accompanied by partial reduction of the O-allyl to an O-propyl group (Scheme 2). The β-D -glycoside 30 behaved very similarly to 11 and 15 . Reductive ring opening of 31 , derived from 29 , yielded the 3-O-acetylated acceptor 32 , while the analogous reaction of the β-D -anomer 20 was accompanied by a rapid 3-O→4-O acyl migration (→ 34 ; Scheme 4). Reductive ring opening of 21 gave the diol 35 . The triacetylated donors 38–40 were obtained from 20 by debenzylidenation, acetylation (→ 36 ), and deallylation (→ 37 ), followed by either acetylation (→ 38 ), treatment with Me3SiSEt (→ 39 ), or Cl3CCN (→ 40 ).  相似文献   

6.
A new approach to ketose-derived nucteosides is described. It is based upon a chain elongation of 1-deoxy-1-nitroaldoses, followed by activation of the nitro group as a leaving group, and introduction of a pyrimidine or purine base. Thus, the nitroaldose 7 was prepared from 3 by pivaloylation (→ 4 ), synthesis of the anomeric nitrones 5/6 , and ozonolysis of 6 (Scheme 1). Partial hydrolysis of 4 yielded 8/9 , which were characterized as the acetates 10/11 and transformed into the nitrones 12/13 . Ozonolysis of 12/13 gave 14/15 , which were acetylated to 16/17 . Henry reaction of 7 lead to 19 and 20 , which were acetylated to 21 and 22 (Scheme 2). Michael addition of 7 to acrylonitrile and to methyl propynoate yielded the anomers 23/24 and 25/26 , respectively. Similar reactions of 16/17 were prevented by a facile β-elimination. Therefore, the nitrodiol 15 was transformed into the orthoesters 27 and then, by Henry reaction, partial hydrolysis, and acetylation, into 28 and 29 (Scheme 2). The structure of 19 was established by X-ray analysis. It was the major product of the kinetically controlled Henry reaction of 7 . Similarly, the β-D-configurated nitroaldoses 23 and 25 were the major products of the Michael addition. This indicates a preferred ‘endo’-attack on the nitronate anion derived from 7 . AMI calculations for this anion indicate a strong pyramidalization at C(1), in agreement with an ‘endo’-attack. Nucleosidation of 21 by 31 afforded 32 and 33 . Yields depended strongly upon the nature and the amount of the promoter and reached 77% for 33 , which was transformed into 34 , 35 , and the known ‘psicouridine’ ( 36 ; Scheme 3). To probe the mechanism, the trityl-protected 30 was nucleosidated yielding 37 , or 37 and 38 , depending upon the amount of FeCl3. Nucleosidation of the nitroacetate 28 was more difficult, required SnCl2 as a promoter, and yielded 39 and 40 . The β-D-anomer 40 was transformed into 36 . Nucleosidation of 23 (SnCl4) yielded the anomers 41 and 42 , which were transformed into 43 and 44 , and hence into 45 and 46 (Scheme 4). Similarly, nucleosidation of 25 yielded 47 and 48 , which were deprotected to 49 and 50 , respectively. The nucleoside 49 was saponified to 51 . Nucleosidation of 21 by 52 (SnCl2) afforded the adenine nucleosides 53 and 54 (Scheme 5). The adenine nucleoside 53 was deprotected (→ 55 → 56 ) to ‘psicofuranine’ (1), which was also obtained from 58 , formed along with 57 by nucleosidation of 28 . The structure and particularly the conformation of the nitroaldoses, nitroketoses, and nucleosides are examined.  相似文献   

7.
Thermal decomposition of LiPF6 and LiBC4O8 (lithium bis(oxalate)borate, abbreviated as LiBOB) were studied using TG (thermogravimetry)-DTG (derivative thermogravimetry) method with different heating rate β of 5, 10, 20 and 40°C min−1 or at different constant temperature θ C (109·80, 118·79, 148·41, 176·86°C for LiPF6 and 278·51, 298·13, 317·65, 336·13 for LiBOB). From the non-isothermal kinetics we calculate that is 1·01, n LiBOB is 1·04, is 91907·61 J/mol, and E LiBOB is 205179·88 J/mol; from the isothermal kinetics we calculate that n for both LiPF6 and LiBOB are 1, ELiPF6 is 91907·61 J/mol, E LiBOB is 205179·88 J/mol, is 16·89 s−1, and lnALiBOB is 31·96 s−1. The results obtained from the two ways have minor differences and can validate each other.  相似文献   

8.
Insertion of the glycosylidene carbenes derived from the diazirines 1 , 14 , and 15 into the B−alkyl bond of the B‐alkyl‐9‐oxa‐10‐borabicyclo[3.3.2]decanes 5 , 6 , and 7 yielded the stable glycosylborinates 8 / 9 (55%, 55 : 45), 10 / 11 (31%, 65 : 35), 12 / 13 (47%, 60 : 40), 16 / 17 (55%, 55 : 45), 18 / 19 (47%, 45 : 55), and 20 / 21 (31%, 30 : 70). Crystal‐structure analysis of 17 and NOEs of 9 and 19 show that 17 , 9 , and 19 adopt similar conformations. The glycosylborinates are stable under acidic, basic and thermal conditions. The unprotected glycosylborinate 25 was obtained in 80% by hydrogenolysis of 12 . Insertion of the glycosylidene carbene derived from the diazirine 1 into a B−C bond of BEt3, BBu3, and BPh3 led to unstable glycosylboranes that were oxidised to yield the hemiacetals 29 (55%), 31 (45%), and 33 (48%), respectively, besides the glucals 30 (13%), 32 (20%), and 34 (20%), respectively. Insertion of the glycosylidene carbenes derived from 14 and 15 into a B−C bond of BEt3 led exclusively to hemiacetals; only 15 yielding traces of the glucal 40 besides the hemiacetal 39 . The glycosylidene carbene derived from 1 reacted with Al(iBu)3 and AlMe3 to generate reactive glycosylalanes that were hydrolysed, yielding the C‐glycosides 46 (21%) and 49 (30%), respectively, besides the glucals 48 (26%) and 51 (30%); deuteriolysis instead of protonolysis led to the monodeuterio analogues of 46 and 49 , respectively, which possess an equatorial 2H‐atom at the anomeric center.  相似文献   

9.
Three novel mixed ligand complexes of Ni(II), Zn(II) and Cd(II) with p-chlorobenzote and N,N-diethylnicotinamide were synthesised and characterized on the basis of elemental analysis, FTIR spectroscopic analysis, solid state UV-Vis spectrometric and magnetic susceptibility data. The thermal behavior of the complexes was studied by simultaneous TG-DTA methods in static air atmosphere and the mass spectra data were recorded. According to microanalytical results, formulas of complexes are C34H40N4O8ClNi, C34H40N4O8ClZn and C34H44N4O10ClCd. The complexes contain two moles of coordination waters, two moles p-chlorobenzoate and two mole N,N-diethylnicotinamide (dena) ligands per formula unit. In these complexes, the p-chlorobenzoate and N,N-diethylnicotinamide behave as monodentate ligand through acidic oxygen and nitrogen of pyridine ring. The decomposition pathways and the stability of the complexes are interpreted in the terms of the structural data. The final decomposition products were found to be as metal oxides.  相似文献   

10.
Cyclopentadienyl C-glycosides (= glycosyl-cyclopentadienes) have been prepared as latent fulvenes. Their reaction with nucleophiles leads to cyclopentadienes substituted with (protected) alditol moieties and, hence, to enantiomerically pure metallocenes. Treatment of 1 with cyclopentadienyl anion gave the epimeric glycosyl-cyclopentadienes 6 / 7 (Scheme 1). Each epimer consisted of a ca. 1:1 mixture of the 1, 3-and 1, 4-cyclopentadienes a and b , respectively, which were separated by prep. HPLC. Slow regioisomerisation occurred at room temperature. Diels-Alder addition of N-phenylmaleimide to 6a / b ca. 3:7 at room temperature yielded three ‘endo’-adducts, i.e., a disubstituted alkene ( 8 or 9 , 25%) and the trisubstituted alkenes 10 (45%) and 11 (13%). The structure of 10 was established by X-ray analysis. Reduction of 6 / 7 (after isolation or in situ) with LiAlH4 gave the cyclopentadienylmannitols 12a / b (80%) which were converted to the silyl ethers 13a / b (Scheme 2). Lithiation of 13a / b and reaction with FeCl2 or TiCl4 led to the symmetric ferrocene 14 (76%) and the titanocene 15 (34%), respectively. The mixed ferrocene 16 (63%) was prepared from 13a / b and pentamethylcyclopentadiene. Treatment of 6 / 7 with PhLi at ?78° gave a 5:3 mixture of the 1-C-phenylated alcohols 17a / b and 18a / b (71%) which were silylated to 19a / b and 20a / b , respectively. Lithiation of 19 / 20 and reaction with FeCl2 afforded the symmetric ferrocenes 21 and 22 and the mixed ferrocene 23 (54:15:31, 79%) which were partially separated by MPLC. The configuration at C(1) of 17–22 was assigned on the basis of a conformational analysis. The reaction of the ribofuranose 24 with cyclopentadienylsodium led to the epimeric C-glycosides 27a / b and 28a (57%, ca. 1:1, Scheme 3). The in-situ reduction of 27 / 28 with LiAlH4 followed by isopropylidenation gave 25a / b (65%) which were transformed into the ferrocene 26 (79%) using the standard method. Phenylation of 27 / 28 , desilylation, and isopropylidenation gave a 20:1 mixture of 33a / b and 34a / b (86%) which was separated by prep. HPLC. The same mixture was obtained upon phenylation of the fulvene 32 which was obtained in 36% yield from the reaction of the aldehydo-ribose 30 with cyclopentadienylsodium at ?100°. Lithiation of 33 / 34 and reaction with FeCl2 gave the symmetric ferrocene 35 (88%). Similarly, the aldehydo-arabinose 36 was transformed via the fulvene 37 (32%) into a 18:1 mixture of 38a / b and 39a / b (78%) and, hence, into the ferrocene 40 (83%). Conformational analysis allowed to assign the configuration of 33–35 , whereas an X-ray analysis of 40 established the (1S)-configuration of 38a / b and 40 . The opposite configuration at C(1) of 38a / b and 33a / b was established by chemical degradation (Scheme 4). Hydrogenation (→ 41 and 44 , resp.), deprotection (→ 42 and 45 , resp.), NaIO4 oxidation, and NaBH4 reduction yielded (+)-(S)- 43 and (?)-(R)- 43 , respectively.  相似文献   

11.
Photoinduced Vinylogous β-Cleavage of Epoxy-enones of the Ionone Series The photochemistry of the α,β-unsaturated γ,δ-epoxy-enones 1–3 is determined by: (i) C(γ)-O-scission of the epoxide (vinylogous β-cleavage of Type A); (ii) C(γ)-C(δ)-cleavage of the oxirane (vinylogous β-cleavage of Type B); (iii) (E/Z)-isomerization of the enone chromophore. In contrast, 4 with tertiary C(β) shows no Type B cleavage. Type A cleavage is induced both by n,π*- and π,π*-excitation and arises probably from the T1-state, but Type B cleavage is observed only on π,π*-excitation and represents presumably a S2-reaction. On Type A cleavage 1–4 undergo 1,2-alkyl-shifts to 1,5-dicarbonyl compounds ( 15–18, 25–28, 34 and 35 ) or rearrange to dihydrofuranes ( 7 and 30 ). The isomerization 1→7 proceeds by a stereoselective [1,3]-sigmatropic shift. On Type B cleavage 1–3 isomerize to a bicyclic enol-ether ( 8, 29 ) or to a monocyclic enol-ether ( 9 ; product of a homosigmatropic [1,5]-shift) or undergo fragmentation to isomers such as allenes 10, 22 and 31 or cyclopropenes 11 and 21 . The non-isolated, unstable (Z)-epoxy-enones 14, 19, 24 and 38 isomerize by fragmentation to the furanes 12, 23, 33 and 39 respectively, on contact with traces of acid or by heating. However, for 19 and 4 , Type B cleavage may lead to the furanes 23 and 39 . On UV. irradiation of the epoxy-enone 4 the initially formed (E/Z)-isomers 34 and 35 yield on π,π*-excitation the enones 37 and 40 by a vinylogous β-fragmentation. In addition, on n,π*-excitation 34 isomerizes to 35 , which decarbonylates exclusively to the enone 37 . The reactions of 1–4 with BF3 · O(C2H5)2 were also studied (see appendix). The epoxy-enones 1 and 2 isomerize by an 1,2-alkyl shift in good yield to the 1,4-dicarbonyl compounds 79 and 81 , whereas 3 gives the 1,4-diketone 83 , and in small amounts the 1,5-diketone 84 . On the other hand, 4 is converted to the fluorohydroxy-enone 85 and to the 1,5-dicarbonyl product 34 , the only isomer in this series which is identical with one of the photoproducts.  相似文献   

12.
The Me3Si? C(1) bond of the bis-(trimethylsilyl)ethynylated anhydroalditol 2 is selectively cleaved with BuLi to yield 3 / 4 , while AgNO2/KCN in MeOH cleaves the Me3Si? C(2′) bond, leading to 5 (Scheme 1). Both Me3Si groups are removed with NaOH in MeOH (→ 7 ), the (i-Pr)3Si group is selectively cleaved with HCl in aq. MeOH ( → 6 ); all silyl substituents are removed with Bu4NF ( → 8 ). Acetolysis transformed 9 into 13 , which was desilylated to 14 , while thiolysis of 9 led to a mixture 11 / 12 . The tetraacetate 14 has also been obtained from 9 via 10 . Oxidative dimerisation of either 3 or 5 , or of a mixture 3 / 5 yields only the homodimers 15 and 16 (Scheme 2); treatment of 16 with AgNO2/KCN yielded 17 , deprotection proceeding much more slowly than the cleavage of the Me3Si? C(2′) group of 2 . The iodoalkyne 20 , required for the cross-coupling with 5 according to Cadiot-Chodkiewicz, was prepared by deprotection of 3 / 4 to 18 , methoxymethylation (→ 19 ), and iodination. Cross-coupling yielded mostly 21 , besides the homodimer 22 . Similarly, cross-coupling of 20 and 23 (obtained from 5 ) led to 24 and 22 . The structure of 24 was established by X-ray analysis (Fig.), showing a C(6)–C(5′) distance of 5.2 Å. The conditions for deprotecting 2 were applied to 21 , and led to 25 (AgNO2/KCN), 26 (aq. NaOH), 27 (Bu4NF), and 29 (HCl/MeOH; Scheme 3). Attempted deprotection of the propargylic-ether moiety with BuLi, however, failed. The dimer 27 was further deprotected to 28 . Acetolytic (Ac2O/Me3SiOTf) debenzylation of the dimer 30 , obtained from 10 , gave 31 (83%) which was deacetylated to 32 (Scheme 4). Cross-coupling of 5 and the bromoalkyne 33 , obtained from 10 , yielded 34 ; again, acetolysis proceeded well, leading to 35 . The cellobiose derivative 38 was prepared from the lactone 36 via 37 . The glycosidic linkage of 38 proved resistant to the conditions of acetolysis, leading to 39 . Acetolysis of the benzylated thiophene 40 (from 30 with Na2S) yielded the octaacetate 41 , but proceeded in substantially lower yields (50%).  相似文献   

13.
The phosphono and the tetrazolyl analogues 4 and 5 of 4-methylumbelliferyl β-D -glucuronide (=(4-methyl-2-oxo-2H-1-benzopyran-7-yl β-D -glucopyranosid)uronic acid; 6 ) were synthesized and evaluated as substrates of β-glucuronidases. Similarly, the phenylcarbamate 7 and its phosphono analogue 8 were prepared and evaluated as inhibitors. To examine the diastereoselectivity of the phosphorylation, we also synthesized the protected L -ido-D -gluco-, and D -galacto-configurated phospha-glycopyranuronates 12, 13, 21, 22, 34 and 35 . Two strategies were followed. In the first one, the glucuronic acid 19 was decarboxylated to 11 and further transformed, via 20 , into the trichloroacetimidate 10 (Scheme 2). Phosphorylation of 10 with (MeO)3P yielded the diastereoisomers 12 and 13 , the diastereoselectivity depending on the solvent. In MeCN, 12 and 13 were obtained in a ratio of 1:1, while in non-participating solvents the L -ido 12 was by far the major diastereoisomer. The acetate 11 was inert to (MeO)3P, but reacted with (PhO)3P to the anomeric mixture 21/22 , in keeping with a stabilizing 1,3-interaction in the intermediate phosphonium salt. Similarly, the phospha-galacturonates 34 and 35 were prepared from the galactoside 23 via the enol ether 26 , the lactone 27 , and the acetates 28/29 that were also transformed into the trichloroacetimidate 33 (Scheme 3). In the second, higher-yielding strategy, phosphorylation of the pentodialdehyde 39 to 40/41 was followed by hydrolysis and acetylation to the phospha-glucuronates 43/44 (Scheme 4). Transesterification to 45/46 , selective deacetylation to 48/49 , and formation of the trichloroacetimidates 50/51 were followed by glycosidation and deprotection to 4 . The tetrazole 5 was prepared from the lactones 54/55 via the N-benzylamides 57/58 that were treated with TfN3 to give the N-benzyltetrazoles 59/60 (Scheme 4). These were transformed into the trichloroacetimidates 63/64 , glycosylated to 65 , and deprotected. The O-carbamoylhydroximo-lactone 7 derived from the glucuronate 67/68 , and the phosphonate analogue 8 were prepared by established methods. The phosphonate 4 is slowly hydrolyzed by the E. coli β-glucuronidase, but neither 4 nor the tetrazole 5 are affected by the bovine liver β-glucuronidase (Table 4). The phenylcarbamate 7 of D -glucarhydroximo-1,5-lactone, but not its phosphonate analogue 8 , is an inhibitor (KI = 8 m?M ) of the E. coli β-glucuronidase. The bovine liver β-glucuronidase is inhibited strongly by 7 (IC50 = 0.2 m?M ) and weakly by 8 (IC50 = 2mM ).  相似文献   

14.
An ab initio 3-21G study of the direct addition of HF to C2HnF(4–n), with n = 0 to 4, has been performed to investigate the effect of the substituent on the reaction. Geometry optimization of all charge-transfer complexes and transition states has been done. Standard analysis of activation energies of addition reactions, vibrational and thermodynamical analysis, as well as Morokuma energy decomposition, BSSE correction, PMO analysis, and Pauling bond orders were used to explain the results. A subset of the reactions, including that of C2H4 as reference one and the two most favorable cases, was also studied at the MP2/6–31G(d,p)//HF/6–31G(d,p) level. The barriers so obtained are in agreement with the indirectly found from experimental data. It was found that the effect of the substituent is not monotonic for the additions. Decomposition of the interaction energy is shown to be adequate to explain this nonmonotonic behavior. The implications for laser chemistry of the addition of hydrogen halides to fluorosubstituted olefins is briefly discussed.  相似文献   

15.
Selected portions of the S0 and T1 potential energy surfaces of acetaldehyde surveyed in our earlier studies have been reexamined. The assumption of the additivity of basis-set polarization and of electron correlation effects used extensively in our earlier work on acetaldehyde has been tested through explicit polarized basis-set electron-correlation calculations. The “additivity assumption” introduces average absolute errors in energy differences of only 1.9 (MP3) to 3.4 (MP2) kcal mol?1 in seven comparisons. The effects of using 6?31G** SCF optimized geometries as opposed to single-point calculations on 3?21G SCF structures (6–31G**//3–21G) as in our previous papers were examined. In six comparisons, the average absolute error in relative SCF energies introduced by the use of the 3–21G geometries rather than the fully consistent 6–31G ones was only 0.3 kcal mole?1. After a uniform scaling procedure, comparisons of the 6–31G** and 3–21G calculated vibrational frequencies with experiment for CH3CHO (S0), CH4, and CO (20 comparisons) yielded absolute differences of 41 cm?1 (6–31G**) and 57 cm?1 (3–21G). All these more elaborate calculations support for the specific case of acetaldehyde and various minima and transition states of relevance to its photochemistry, the commonly used and practically important approximations (e.g., additivity) made in our earlier studies.  相似文献   

16.
Hetero- or homocoupling of protected 1,4-cis-diethynylated 1,5-anhydroglucitols leads to two isomeric cyclotrimers and to four isomeric cyclotetramers. The C3-symmetric cyciotrimer 31 , the C4-symmetric cyclotetramer 35 , and the D2-symmetric cyclotetramer 6 have been prepared before. We have now synthesized the C1-symmetric cyciotrimer 13 , and the C1- and the C2-symmetric cyclotetramers 22 and 27 , respectively. The cyclotrimer 13 was prepared by intramolecular, oxidative homocoupling and, alternatively, by a one-pot trimerization/cyclization of the monomer 36 (Schemes 1 and 5, resp.). Oxidative homocoupling was used for the cyclization of the tetramers 19 and 25 , leading to 22 and 27. The tetramer 19 was made by sequential Cadiot-Chod-kiewicz coupling (Scheme 2)and the tetramer 25 by a combination of a Cadiot-Chodkiewicz reaction and an intermolecular, oxidalive homocoupling (Scheme 3). The acetates 34 and 38 , corresponding to 35 and 27 , respectively, were also made by a one-pot dimerization/cyclization of the dimer 37 (Scheme 5). Intramolecular oxidative heterocoupling is also feasible and results in an alternative, more convenient synthesis of the acetylated cyclotrimer 30 and the acetylated cyclotetramer 34 (corresponding to 31 and 35 , resp.; Scheme 4). The solid-state conformation of the C4-symmetric cyclotetramer 34 corresponds well to the one predicted by force-field calculations. We compared the water-solubilities of the cyclotrimers 13 and 31 and the tetramers 6, 22, 27 , and 35 , their calculated conformations (MM3*), and the D -adenosine binding properties of the cyclotetramers 6, 27 , and 35 .  相似文献   

17.
A protease from newly isolated Bacillus circulans M34 was purified by Q‐Sepharose anion exchange chromatography and Sepharose–bacitracin affinity chromatography followed by (NH4)2SO4 precipitation. The molecular mass of the purified enzyme was determined using SDS–PAGE. The optimum pH and temperature for protease activity were 11 and 50°C, respectively. The effect of various metal ions on protease activity was investigated. Alkaline protease from Bacillus circulans M34 wase activated by Zn2+, Cu2+ and Co2+ up to 31%. The purified protease was found to be stable in the organic solvents, surfactants and oxidizing agent. The substrate specificity of purified protease was investigated towards different substrates. The protease was almost completely inhibited by the serine protease inhibitor phenylmethanesulfonyl fluoride. The kinetic parameters of the purified protease, maximum rate (Vmax) and Michaelis constant (Km), were determined using a Lineweaver–Burk plot. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

18.
High-dose chemotherapy (HDCT) and autologous bone marrow/blood stem cell transplantation are an effective combination for treating a number of malignant disorders. The contamination of the autograft by malignant cells may be a reason for recurrences in spite of this treatment, for instance, in multiple myeloma. Therefore, we evaluated the use of photodynamic therapy (PDT) using the photosensitizer benzoporphyrin derivative mono-acid ring A (BPD-MA) on multiple myeloma cells in comparison to the components of the normal bone marrow (NBM) and peripheral blood apheresis product. Flow cytometry was used to measure differential BPDMA uptake of NBM components: namely lymphocytes, monocytes, granulocytes and enriched hematopoietic stem cell (CD34+) populations and also the multiple myeloma cell lines OCI-MY7 and OCI-MY4. When each population was measured individually, the order of uptake was [OCI-MY7/MY4] > [CD34+] > [granulocytes] = [monocytes] ? [lymphocytes]. Further, clonogenic assay was used to demonstrate surviving fractions for OCI-MY7, OCI-MY4 and NBM in vitro. The LD90 for OCI-MY7 and OCI-MY4 was between 10 and 20 ng/mL BPD-MA whereas this concentration did not show any significant cell kill for the colony-forming units-granulocyte/macrophage (CFU-GM) and burst-forming units—erythrocyte (BFU-E). When the NBM was “contaminated” with multiple myeloma cells in vitro, the LD90 for OCI-MY7 in this cell mixture was shifted to between 40 and 80 ng/mL BPD-MA. However, at 40 ng/ mL BPD-MA at least 50% of normal CFU-GM and BFU-E colonies survived. For CFU-GM and BFU-E derived from the enriched CD34+ cell population, BPDMA up to a concentration of 80 ng/mL did not significantly reduce the surviving fractions. We have observed a 3–4 log therapeutic window with differential cell kill when comparing multiple myeloma cell lines to the components of the NBM and apheresis product in vitro. We conclude, that BPD-MA is a molecule potentially useful as an ex vivo purging agent.  相似文献   

19.
The cis-configurated isocephem 26 as well as isodethiaoxacephems 21 and 33 were synthesized (Schemes 4 and 5). The key step involves chlorination of the corresponding carbanions of 23, 18 , and 31 with CF3SO2Cl followed by internal alkylation. β-Lactams 3, 21, 26 , and 33 were found to possess biological activity against several pathogenic microorganisms in vitro. Electronic activation of the lactam moiety in isodethiaoxacephem 33 remarkably enhanced its biological activity.  相似文献   

20.
NaSMe in toluene leads to regioselective de-C-silylation of the bis[(trimethylsilyl)ethynyl]saccharide 2 , but to decomposition of butadiynes such as 1 or 12 . We have, therefore, combined the known reagent-controlled, regioselective desilylation of 2 and of 12 (AgNO2/KCN) with a substrate-controlled regioselective de-C-silylation, based on C-silyl groups of different size. This combination was studied with the fully protected 3 which was mono-desilylated to 4 or to 5 (Scheme 1). Triethylsilylation of 5 (→ 6 ) was followed by removal of the Me3Si group (→ 7 ), introduction of a (t-Bu)Me2Si group (→ 8 ) and removal of the Et3Si group yielded 9 ; these high-yielding transformations proceed with a high degree of selectivity. Iodination of 4 gave 10 . The latter was coupled with 5 to the homodimer 11 and the heterodimer 12 , which was desilylated to 13 . The second building block for the tetramer was obtained by coupling 14 (from 7 ) with 5 , leading to 15 and 16 . Removal of the Me3Si group (→ 17 ) and iodination led to 18 which was coupled with 13 to the homotetramer 20 and the heterotetramer 19 (Scheme 2). Deprotection of 19 gave 21 , which was, on the one hand, iodinated to 22 , and, on the other hand, protected by the (t-Bu)Me2Si group (→ 23 ). Removal of the Et3Si group (→ 24 ) and coupling afforded the homooctamer 26 and the heterooctamer 25 . Yields of iodination, silylation, and desilylation were consistently high, while heterocoupling proceeded in only 50–55%. Cleavage of the (i-Pr)3SiC and MeOCH2O groups of 11 (→ 27 ), 15 (→ 28 ), 20 (→ 29 ) and 26 (→ 30 ) proceeded in high yields (Scheme 3). Complete deprotection in two steps of the heterocoupling products 16 (→ 31 → 32 ), 19 (→ 33 → 34 ), and 25 (→ 35 → 36 ) gave the unprotected dimer 32 , tetramer 34 , and octamer 36 in high yields (Scheme 4). Only the dimer 32 is soluble in H2O; the 1H-NMR spectra of 32 , 34 , and 36 in (D6)DMSO (relatively low concentration) show no signs of association.  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